Phonons in Nanostructures-Cambridge University Press (2001)
Phonons in Nanostructures-Cambridge University Press (2001)
Phonons in Nanostructures-Cambridge University Press (2001)
Phonons in nanostructures
D R D UTTA earned a Ph.D. in physics from the University of Cincinnati; she was
a research associate at Purdue University and at City College, New York, as well
as a visiting scientist at Brookhaven National Laboratory before assuming a
variety of government posts in research and development. Dr Dutta was the
Director of the Physics Division at the US Army’s Electronics Technology and
Devices Laboratory as well as at the Army Research Laboratory prior to her
appointment as the Associate Director for Electronics in the Army Research
Office’s Engineering Sciences Directorate. Dr Dutta recently assumed a senior
executive position as ARO’s Director of Research and Technology Integration.
She has over 160 publications, 170 conference presentations, 10 book chapters,
and has had 24 US patents issued. She is the joint editor of two World Scientific
books entitled Quantum-Based Electronic Devices and Systems and Advances in
Semiconductor Lasers and Applications to Optoelectronics. She is an Adjunct
Professor of the Electrical and Computer Engineering and Physics departments
of North Carolina State University and has had adjunct appointments at the
Electrical Engineering departments of Rutgers University and the University of
Maryland. Dr Dutta is a Fellow of both the Institute of Electrical and Electronics
Engineers (IEEE) and the Optical Society of America, and she was the recipient
in the year 2000 of the IEEE Harry Diamond Award.
Phonons in
Nanostructures
Michael A. Stroscio and Mitra Dutta
US Army Research Office, US Army Research Laboratory
The Pitt Building, Trumpington Street, Cambridge, United Kingdom
http://www.cambridge.org
and
Preface xi
vii
viii Contents
6.1 Non-parabolic terms in the crystal potential for ionically bonded atoms 45
6.2 Klemens’ channel for the decay process LO → LA(1) + LA(2) 46
6.3 LO phonon lifetime in bulk cubic materials 47
6.4 Phonon lifetime effects in carrier relaxation 48
6.5 Anharmonic effects in würtzite structures: the Ridley channel 50
Appendices 221
Appendix A: Huang–Born theory 221
Appendix B: Wendler’s theory 222
Appendix C: Optical phonon modes in double-heterointerface structures 225
Appendix D: Optical phonon modes in single- and double-heterointerface würtzite
structures 236
Appendix E: Fermi golden rule 250
Appendix F: Screening effects in a two-dimensional electron gas 252
References 257
Index 271
Preface
xi
xii Preface
For both these polar semiconductors, the process of LO phonon emission plays a
major role in determining the value of the saturation velocity. In non-polar materials
such as Si, which has a saturation velocity of about 107 cm s−1 , the deformation-
potential interaction results in electron energy loss through the emission of phonons.
(In Chapter 5 both the interaction between polar-optical-phonons and electrons –
known as the Fröhlich interaction – and the deformation-potential interaction will
be defined mathematically.)
Clearly, in all these cases, the electron mobility will be influenced strongly by the
interaction of the electrons with phonons. The saturation velocity of the carriers in
a semiconductor provides a measure of how fast a microelectronic device fabricated
from this semiconductor will operate. Indeed, the minimum time for the carriers to
travel through the active region of the device is given approximately by the length
of the device – that is, the length of the so-called gate – divided by the saturation
velocity. Evidently, the practical switching time of such a microelectronic device
will be limited by the saturation velocity and it is clear, therefore, that phonons play
a major role in the fundamental and practical limits of such microelectronic devices.
For modern integrated circuits, a factor of two reduction in the gate length can be
achieved in many cases only through building a new fabrication facility. In some
cases, such a building project might cost a billion dollars or more. The importance
of phonons in microelectronics is clear!
A second example of the importance of carrier–phonon interactions in modern
semiconductor devices is given by the dynamics of carrier capture in the active
quantum-well region of a polar semiconductor quantum-well laser. Consider the
case where a current of electrons is injected over a barrier into the quantum-well
region of such a laser. For the laser to operate, an electron must lose enough energy
to be ‘captured’ by the quasi-bound state which it must occupy to participate in
the lasing process. For many quantum-well semiconductor lasers this means that
the electron must lose an energy of the order of a 100 meV or more. The energy
loss rate of a carrier – also known as the thermalization rate of the carrier – in
a polar-semiconductor quantum well is determined by both the rate at which the
carrier’s energy is lost by optical-phonon emission and the rate at which the carrier
gains energy from optical-phonon absorption. This latter rate can be significant
in quantum wells since the phonons emitted by energetic carriers can accumulate
in these structures. Since the phonon densities in many dimensionally confined
semiconductor devices are typically well above those of the equilibrium phonon
population, there is an appreciable probability that these non-equilibrium – or ‘hot’
– phonons will be reabsorbed. Clearly, the net loss of energy by an electron in such
a situation depends on the rates for both phonon absorption and phonon emission.
Moreover, the lifetimes of the optical phonons are also important in determining the
total energy loss rate for such carriers. Indeed, as will be discussed in Chapter 6, the
longitudinal optical (LO) phonons in GaAs and many other polar materials decay
into acoustic phonons through the Klemens’ channel. Furthermore, over a wide
1.2 Devices with nanostructure components 3
where r and k are the position vector and wavevector components in a plane
parallel to the interfaces, k z = nπ/L z , and n = 1, 2, 3, . . . labels the energy
eigenstates, whose energies are
h̄ 2 (k )2 h̄ 2 π 2 n 2
E n (k ) = + . (1.2)
2m 2m L 2z
face-centered cubic (fcc) lattices depicted in Figure 2.1 have no excess or deficit
of charge relative to the neutral situation. This changes dramatically for polar
semiconductors like gallium arsenide, since here the ionic bonding results in charge
transfer from the Group V arsenic atoms to the Group III gallium atoms: Since
Group V atoms have five electrons in the outer shell and Group III atoms have
three electrons in the outer shell, it is not surprising that the gallium sites acquire
a net negative charge and the arsenic sites a net positive charge. In binary polar
semiconductors, the two atoms participating in the ionic bonding carry opposite
charges, e∗ and −e∗ , respectively, as a result of the redistribution of the charge
associated with polar bonding. In polar materials such ionic bonding is characterized
by values of e∗ within an order of magnitude of unity. In the remaining sections of
this chapter, it will become clear that e∗ is related to the readily measurable or known
ionic masses, phonon optical frequencies, and high-frequency dielectric constant of
the polar semiconductor.
since, for many semiconductors, its frequency is in the terahertz range, which
happens to coincide with the infrared portion of the electromagnetic spectrum.
The lower-frequency solution is known as the acoustic mode. More precisely, since
only longitudinal displacements have been modeled, these two solutions correspond
to the longitudinal optical (LO) and longitudinal acoustic (LA) modes of the
linear-chain lattice. Clearly, the displacements along this chain can be described
in terms of wavevectors q in the range from −π/2a to π/2a. From the solution
for ω, it is evident that over this Brillouin zone the LO modes have a maximum
frequency [2α(1/m + 1/M)]1/2 at the center of the Brillouin zone and a minimum
frequency (2α/m)1/2 at the edge of the Brillouin zone. Likewise, the LA modes have
a maximum frequency (2α/M)1/2 at the edge of the Brillouin zone and a minimum
frequency equal to zero at the center of the Brillouin zone.
In polar semiconductors, the masses m and M carry opposite charges, e∗ and
−e∗ , respectively, as a result of the redistribution of the charge associated with
polar bonding. In polar materials such ionic bonding is characterized by values of
e∗ equal to 1, to an order-of-magnitude. When there is an electric field E present
in the semiconductor, it is necessary to augment the previous force equation with
terms describing the interaction with the charge. In the long-wavelength limit of the
electric field E, the force equations then become
and
and
and
m
−Mω u 2 = 2α u 1 + u 1 − e∗ E;
2
(2.13)
M
accordingly,
and
where ω02 = 2α(1/m + 1/M) is the resonant frequency squared, in the absence of
Coulomb effects; that is, for e∗ = 0. The role of e∗ in shifting the phonon frequency
will be discussed further in the next section.
Clearly, the electric polarization P produced by such a polar diatomic lattice is
given by
N e∗ u N e∗ (u 1 − u 2 ) 1 N e∗2 1 1
P= = = + E, (2.16)
(∞) (∞) (∞) (ω02 − ω2 ) m M
A2 2α cos qa 2α − mω2
= = , (2.18)
A1 2α − Mω 2 2α cos qa
and it is clear that the ratio of the displacement amplitudes is approximately equal
to unity for acoustic phonons near the center of the Brillouin zone. Thus, in contrast
to the optical modes, the acoustic modes are characterized by in-phase motion of
2.3 Linear-chain model and macroscopic models 11
the different masses m and M. Typical mode patterns for zone-center acoustic
and optical modes are depicted in Figures 2.3(a), (b). The transverse modes are
illustrated here since the longitudinal modes are more difficult to depict graphically.
The higher-frequency optical modes involve out-of-plane oscillations of adjacent
ions, while the lower-frequency acoustic modes are characterized by motion of
adjacent ions on the same sinusoidal curve.
2.3.3 Polaritons
In the presence of a transverse electric field, transverse optical (TO) phonons of
a polar medium couple strongly to the electric field. When the wavevectors and
frequencies of the electric field are in resonance with those of the TO phonon, a
coupled phonon–photon field is necessary to describe the system. The quantum of
this coupled field is known as the polariton. The analysis of subsection 2.3.1 may
be generalized to apply to the case of transverse displacements. In particular, for a
transverse field E, the oscillator equation takes the form
N e∗2 1 1
(ωTO
2
− ω2 )P = + E, (2.19)
(∞) m M
Figure 2.3. Transverse displacements of heavy ions (large disks) and light ions
(small disks) for (a) transverse acoustic modes, and (b) transverse optical modes
propagating in the q-direction.
12 2 Phonons in bulk cubic crystals
the transverse optical frequency. As will become apparent later in this section, the
LO phonon frequency squared differs from the TO phonon frequency squared by an
amount proportional to e∗2 .
According to the electromagnetic wave equation, ∂ 2 D/∂t 2 = c2 ∇ 2 E, where
D = E + 4π P, the dispersion relation describing the coupling of the field E of the
electromagnetic wave to the electric polarization P of the TO phonon is
c2 q 2 E = ω2 (E + 4π P) (2.20)
or, alternatively,
where waves of the form ei(qr −ωt) have been assumed. The driven oscillator
equation and the electromagnetic wave equation have a joint solution when the
determinant of the coefficients of the fields E and P vanishes,
ω2 − c2 q 2 4π ω2
N e∗2 1 1 =0 (2.22)
−(ωTO − ω )
(∞) m + M
2 2
where the polarization due to the electronic contribution, Pe (ω), has been included
as well as the polarization associated with the ionic contribution, P(ω).
As is customary, the dielectric constant due to the electronic response is denoted
by
4π Pe (ω)
(∞) = 1 + , (2.25)
E(ω)
and it follows that
4π N e∗2 1 1
(ω) = (∞) + 2 + . (2.26)
(ωTO − ω2 ) (∞) m M
4π N e∗2 1 1
(0) = (∞) + 2 + . (2.27)
ωTO (∞) m M
From these last two results it follows straightforwardly that
[(0) − (∞)]ωTO
2
(ω) = (∞) +
(ωTO
2 − ω2 )
(0) − (∞)
= (∞) + . (2.28)
1 − ω2 /ωTO
2
From electromagnetic theory it is known that the dielectric function (ω) must
vanish for any longitudinal electromagnetic disturbance to propagate. Accordingly,
the frequency of the LO phonons, ωLO , must be such that (ωLO ) = 0; from the last
equation, this condition implies that
(0) − (∞)
(ωLO ) = 0 = (∞) + (2.29)
1 − ωLO
2 /ω2
TO
or, equivalently,
1/2
(0)
ωLO = ωTO . (2.30)
(∞)
It then follows that
(0) − (∞) (ωLO /ωTO )2 (∞) − (∞)
(ω) = (∞) + = (∞) +
1 − ω2 /ωTO
2 1 − ω2 /ωTO
2
(ωLO /ωTO )2 − 1
= (∞) 1 +
1 − ω2 /ωTO
2
2 − ω2
ωTO ωLO
2 − ω2
TO
= (∞) + 2
ωTO
2 − ω2 ωTO − ω2
2 − ω2
ωLO
= (∞) , (2.31)
ωTO
2 − ω2
or alternatively
(ω) ω2 − ω2
= LO . (2.32)
(∞) ωTO
2 − ω2
In the special case where ω = 0, this relation reduces to the celebrated Lyddane–
Sachs–Teller relationship
(0) ω2
= LO . (2.33)
(∞) ωTO
2
1 ∂ 2E 1 ∂ 2P
−4π∇(∇ · P) − ∇ 2 E + + 4π = 0. (2.36)
c2 ∂ 2 t c2 ∂ 2 t
Assuming that P and E both have spatial and time dependences of the form
ei(q·r−ωt) ,
this last result takes the form
−4πω2 P/c2
E= . (2.38)
q 2 − ω2 /c2
From Appendix A, E and P are also related through
1 [(0) − (∞)]ωTO 2
P= + [(∞) − 1] E; (2.39)
4π ωTO
2 − ω2
thus
or, equivalently,
q 2 c2 2 (0) − ω2 (∞)
ωTO
= . (2.41)
ω2 ωTO
2 − ω2
4π ω2 P/c2 4πq Pq
E= − 2
q 2 − ω2 /c2 q − ω2 /c2
4π ω2 P/c2 4πq 2 P
= 2 −
q − ω2 /c2 q 2 − ω2 /c2
4π ω2
= 2 −q P
2
q − ω2 /c2 c2
= −4π P. (2.42)
Then
1 [(0) − (∞)]ωTO 2
P= + [(∞) − 1] E
4π ωTO
2 − ω2
[(0) − (∞)]ωTO
2
=− + [(∞) − 1] P (2.43)
ωTO
2 − ω2
or, equivalently,
1/2
(0)
ω = ωTO = ωLO ,
(∞)
and the Lyddane–Sachs–Teller relation is recovered once again! In Chapter 3, we
shall return to the Loudon model to describe uniaxial crystals of the würtzite type.
Chapter 3
Phonons in bulk würtzite crystals
Next when I cast mine eyes and see that brave vibration, each
way free; O how that glittering taketh me.
Robert Herrick, 1648
16
3.1 Basic properties of phonons in würtzite structure 17
The numbers of the various long-wavelength modes are summarized in Table 3.1.
For the zincblende case, s = 2 and there are six modes: one LA, two TA, one
LO and two TO. For the würtzite case, s = 4 and there are 12 modes: one LA,
two TA, three LO and six TO. In the long-wavelength limit the acoustic modes are
simple translational modes. The optical modes for a würtzite structure are depicted
in Figure 3.2.
From Figure 3.2 it is clear that the A1 and E 1 modes will produce large electric
polarization fields when the bonding is ionic. Such large polarization fields result
in strong carrier–optical-phonon scattering. These phonon modes are known as
infrared active. As we shall see in Chapter 5, the fields associated with these infrared
modes may be derived from a potential describing the carrier–phonon interaction of
Figure 3.2. Optical phonons in würtzite structure. From Gorczyca et al. (1995),
American Physical Society, with permission.
3.2 Loudon model of uniaxial crystals 19
⊥ (ω) 0 0
(ω) = 0 ⊥ (ω) 0 (3.1)
0 0 (ω)
with
ω2 − ωLO,⊥
2 ω2 − ωLO,
2
⊥ (ω) = ⊥ (∞) and (ω) = (∞) ,
ω2 − ωTO,⊥
2 ω2 − ωTO,
2
(3.2)
80
Energy (meV)
60
40
20
0
Reduced wavevector
Figure 3.3. Phonon dispersion curves for GaN crystal of würtzite structure. From
Nipko et al. (1998), American Institute of Physics, with permission.
20 3 Phonons in bulk würtzite crystals
1/2
µN ⊥ (∞) − 1
P⊥ = ⊥ (0) − ⊥ (∞) ωTO,⊥ u⊥ + E⊥ , (3.5)
4π V 4π
µN 1/2 (∞) − 1
P = (0) − (∞) ωTO, u + E , (3.6)
4π V 4π
Figure 3.4. Dielectric constants for GaN, 1⊥ (GaN) and 1z (GaN), and for AlN,
2⊥ (AlN) and 2z (AlN). From Lee et al. (1998), American Physical Society, with
permission.
3.2 Loudon model of uniaxial crystals 21
ωLO,⊥
2 − ω2
A⊥ = ⊥ (∞) − 1, (3.11)
ωTO,⊥
2 − ω2
ωLO,
2 − ω2
A = (∞) − 1, (3.12)
ωTO,
2 − ω2
q 2 c2 ωTO,⊥
2 ⊥ (0) − ω2 ⊥ (∞)
= . (3.14)
ω2 ωTO,⊥
2 − ω2
Thus,
22 3 Phonons in bulk würtzite crystals
In the limit where retardation effects are neglected, c → ∞ and it follows that
and
−4π [q (q · P) − ω2 P /c2 ]
E = → −4π(P⊥ sin θ cos θ + P cos2 θ)
q 2 − ω2 /c2
= − sin θ cos θ A⊥ E ⊥ − cos2 θ A E . (3.18)
ωLO,
2 − ω2
+ (∞) cos2 θ
ωTO,
2 − ω2
or equivalently
⊥ (ω)q⊥
2
+ (ω)q2 = 0. (3.21)
ωLO,⊥
2 − ω2 ωLO,
2 − ω2
sin2 θ + cos2 θ = 0 (3.22)
ωTO,⊥
2 − ω2 ωTO,
2 − ω2
or equivalently
where
ω12 = ωTO,
2
sin2 θ + ωTO,⊥
2
cos2 θ,
(3.24)
ω22 = ωLO,
2
cos2 θ + ωLO,⊥
2
sin2 θ.
When ωTO, − ωTO,⊥ is very much less than ωLO, − ωTO, and ωLO,⊥ − ωTO,⊥
this equation has roots
1
ω2 = (ω12 + ω22 ) ± [(ω12 − ω22 ) + 2ω2 (θ)] , (3.25)
2
where
(ωLO,
2 − ωLO,⊥
2 )(ωTO,
2 − ωTO,⊥
2 )
ω2 (θ) = 2 sin2 θ cos2 θ; (3.26)
ω22 − ω12
thus
ω2 = ωTO,
2
sin2 θ + ωTO,⊥
2
cos2 θ
(ωLO,
2 − ωLO,⊥
2 )(ωTO,
2 − ωTO,⊥
2 )
− sin2 θ cos2 θ
ω22 − ω12
≈ ωTO,
2
sin2 θ + ωTO,⊥
2
cos2 θ (3.27)
and
ω2 = ωLO,
2
cos2 θ + ωLO,⊥
2
sin2 θ
(ωLO,
2 − ωLO,⊥
2 )(ωTO,
2 − ωTO,⊥
2 )
+ sin2 θ cos2 θ
ω22 − ω12
≈ ωLO,
2
cos2 θ + ωLO,⊥
2
sin2 θ. (3.28)
infrared-active modes in these III-V nitrides are the A1 (LO), A1 (TO), E 1 (LO), and
E 1 (TO) modes and the frequencies associated with these modes, ω A1 (LO) , ω A1 (TO) ,
ω E 1 (LO) , and ω E 1 (TO) are given by ωLO, , ωTO, , ωLO,⊥ , and ωTO,⊥ , respectively.
Let us consider the case of GaN in more detail. From the results of Section 3.2, it
follows immediately that
E⊥ sin θ cos θ A sin θ cos θ A sin θ
=− = = , (3.29)
E 1 + sin2 θ A⊥ cos2 θ A cos θ
and
1/2
u⊥ ωTO,
2 − ω2 ⊥ (0) − ⊥ (∞) ωTO,⊥ E⊥
= 2
u ωTO,⊥ − ω2 (0) − (∞) ωTO, E
1/2
ωTO,
2 − ω2 ⊥ (0) − ⊥ (∞) ωTO,⊥ sin θ
= . (3.30)
ωTO,⊥
2 − ω2 (0) − (∞) ωTO, cos θ
Since q = (q⊥ , q ) = (q sin θ, q cos θ), the first of these relations illustrates the
fact that E q, as expected from q 2 E = −4πq(q · P); this last equality follows from
∇ ·(E+4πP) = 0. The ratio u ⊥ /u may be estimated for GaN for the transverse-like
modes, with ω2 = ωTO, 2 sin2 θ + ωTO,⊥
2 cos2 θ, as
u⊥ ⊥ (0) − ⊥ (∞) 1/2 ωTO,⊥ cos θ cos θ
=− ≈ −0.95 , (3.31)
u (0) − (∞) ωTO, sin θ sin θ
and for the longitudinal-like modes, with ω2 = ωLO,
2 cos2 θ + ωLO,⊥
2 sin2 θ, as
u⊥ ωTO,
2 − ω2 ⊥ (0) − ⊥ (∞) 1/2 ωTO,⊥ sin θ
= 2
u ωTO,⊥ − ω2 (0) − (∞) ωTO, cos θ
ωTO,
2 − ωLO2
⊥ (0) − ⊥ (∞) 1/2 ωTO,⊥ sin θ
= 2
ωTO,⊥ − ωLO 2 (0) − (∞) ωTO, cos θ
sin θ
= 1.07 , (3.32)
cos θ
Table 3.2. Difference frequencies in cm−1 for GaN and AlN as well
as for other würtzite crystals.
Würtzite ωTO, − ωTO,⊥ ωLO, − ωTO, ωLO,⊥ − ωTO,⊥
GaN 27 211 186
AlN 59 279 243
AgI 0 18 18
BeO 44 403 375
CdS 9 71 64
ZnO 33 199 178
ZnS 0 76 76
3.3 Application of Loudon model to III-V nitrides 25
where ωLO
2 is taken to be equal to both ω2
LO, and ωLO,⊥ since ωLO, ≈ ωLO,⊥ .
2 2 2
The properties of uniaxial crystals derived in this section and in Section 3.2 will
be used extensively in Chapter 7 to determine the Fröhlich potentials in würtzite
nanostructures.
Chapter 4
When you measure what you are speaking about and express it
in numbers, you know something about it; but when you
cannot measure it, when you cannot express it in numbers, your
knowledge is of a meagre and unsatisfactory kind; it may be the
beginning of knowledge but you have scarcely in your thoughts
advanced to the stage of science, whatever the matter may be.
Lord Kelvin, 1889
This chapter deals with the application of Raman scattering techniques to measure
basic properties of phonons in dimensionally confined systems. It is, however,
appropriate at this point to emphasize that non-Raman techniques such as neutron
scattering (Waugh and Dolling, 1963) have been used for many years to determine
the phonon dispersion relations for bulk semiconductors. Indeed, for thermal
neutrons the de Broglie wavelengths are comparable to the phonon wavelengths.
For bulk samples, neutron scattering cross sections are large enough to facilitate
the measurement of phonon dispersion relations. This is generally not the case for
quantum wells, quantum wires, and quantum dots, where Raman and micro-Raman
techniques are needed to make accurate measurements of dispersion relations
in structures of such small volume. Further comparisons of neutron and Raman
scattering measurements of phonon dispersion relations are found in Section 7.5.
Raman scattering has been a very effective experimental technique for observing
phonons; it involves measuring the frequency shift between the incident and
26
4.2 Raman scattering for bulk zincblende and würtzite structures 27
Since the momenta of the incident and scattered photons are small compared with
the reciprocal lattice vectors, only excitations with q 0 take part in the Raman
process illustrated in Figure 4.1. In the case of Raman scattering in semiconductors,
the absorption of photons gives rise to electron–hole pairs; hence the intensity of the
Raman scattering and the resonances reflect the underlying electronic structure of
the material. The Raman intensity, I (ωi ), is given by
1
I (ωi ) ∝ ωs4 |εε s Tεε i |2 (4.3)
α,β
(E α − h̄ωi )(E β − h̄ωs )
where the ωi and ωs are the frequencies of the incoming photon and of the
scattered photon respectively, E α and E β are the energies of the intermediate states,
T the Raman tensor, and ε i and ε s are the incident and scattered polarization
vectors. The summation is over all possible intermediate states. In general, for
semiconductors there may be the following real intermediate states: Bloch states,
which form the conduction or valence bands, exciton states and in-gap impurity
states. In equation (4.3), the second factor gives the Raman selection rules, which
come about from symmetry considerations of the interactions involved in a Raman
process. The selection rules are conveniently summarized in the form of Raman
tensors. These selection rules are essential tools for determining crystal orientation
and quality.
Details of the theoretical description of Raman scattering and these effects in
the vicinity of the critical points of the semiconductor are given in excellent books
and reviews elsewhere (Loudon, 1964; Hayes and Loudon, 1978; Cardona, 1975;
Cardona and Güntherodt, 1982a, b, 1984, 1989, 1991) and will not be repeated here.
Instead we will summarize key results in zincblende and würtzite crystals both for
the bulk case and, in Chapter 7, for quantum wells and superlattices. While first-rate
articles and book chapters exist for the results of the zincblende structures, the work
on the nitrides, with their würtzite structure, is more recent and hence in this book
we will cover the latter results in more detail.
0 0 0
0 0 d R(x) mode,
0 d 0
0 0 d
0 0 0 R(y) mode, (4.4)
d 0 0
0 d 0
d 0 0 R(z) mode.
0 0 0
Raman scattering has been used now for several decades as a characterization
tool in understanding, for example, crystal structure and quality, impurity content,
strain, interface disorder, and the effects of alloying and sample preparation. Much
work has been done in this class of cubic zincblende crystals since the first laser
measurements of Hobden and Russell (1964) in zincblende GaP. The prototypical
system that has been studied extensively is GaAs, and comprehensive reviews
are available (Loudon and Hayes, 1978; Cardona, 1975; Cardona and Güntherodt,
1982a, b, 1984, 1989, 1991). Frequencies of the LO and TO modes, ωLO and ωTO
respectively, for some of these systems are listed in Table 4.1.
4.2 Raman scattering for bulk zincblende and würtzite structures 29
A1 , 2B1 , E 1 and 2E 2 . The A1 and E 1 modes and the two E 2 modes are Raman
active while the B modes are silent. The A and E modes are polar, resulting in
a splitting of the LO and the TO modes (Hayes and Loudon, 1978). The Raman
tensors for the würtzite structure are as follows:
a 0 0
0 a 0 A1 (z) mode
0 0 b
0 0 c
0 0 0 E 1 (x) mode
c 0 0
(4.5)
0 0 0
0 0 c E 1 (y) mode
0 c 0
f 0 0 0 −f 0
0 −f 0 −f 0 0 E 2 mode.
0 0 0 0 0 0
The vibrational modes in würtzite structures are given in Figure 3.3. Details of the
frequencies are given in Table 4.2.
Following some early work (Manchon et al., 1970; Lemos et al., 1972; Burns
et al., 1973) there has been a number of more recent experiments (Murugkar et al.,
1995; Cingolani et al., 1986; Azuhata et al., 1995) identifying the Raman modes
in these nitride materials. The early work was mainly on crystals in the form of
needles and platelets and the more recent work has been on epitaxial layers grown
on sapphire, on 6H-SiC, and on ZnO as well as some more unusual substrates. Table
4.2 gives the Raman modes as well as the scattering geometry in which they were
observed in the experiments of Azuhata et al. (1995). Experiments on AlN and InN
crystallites and films, particularly for the latter material, are more scarce, reflecting
the difficulties in achieving good growth qualities for these materials. In uniaxial
materials, when the long-range electrostatic field interactions of the polar phonons
dominate the short-range field of the vibrational force constants, phonons of mixed
symmetry can be observed (Loudon, 1964) under specific conditions of propagation
direction and polarization. They have been seen in the case of AlN (Bergman et al.,
1999).
of the optical modes, which involves the anharmonic effects mentioned previously,
will be discussed here. Other processes that give experimental information on the
electron–phonon interaction include the generation of optical phonons by high-
energy carriers, intervalley scattering between different minima in the conduction
band, and carrier–carrier scattering; these are reported by Kash and Tsang (1991)
for the prototypical system of GaAs.
Measurements of phonon linewidths for Raman and infrared measurements in
GaAs, ZnSe, and GaP give phonon lifetimes of 2–10 ps (von der Linde, 1980;
Menéndez and Cardona, 1984). For systems that are not far from equilibrium, the
lifetimes of the phonons can be described by anharmonic processes. The decay of
an optical phonon is frequently via pairs of acoustic phonons or via one acoustic
phonon and one optical phonon of appropriate energies and momenta (Cowley,
1963; Klemens, 1966). The first measurements with continuous-wave pumping
(Shah et al., 1970) of highly non-equilibrium LO phonons in GaAs yielded estimates
of LO-phonon lifetimes of approximately 5 ps at room temperature. This was
consistent with values obtained from linewidth studies. von der Linde (1980) used
time-resolved Raman scattering to obtain directly the time decay of non-equilibrium
LO phonons. They obtained a value of 7 ps for GaAs LO phonons at 77 K.
Subsequent experiments by Kash et al. (1985) led to the conclusion that the LO
phonon lifetime in GaAs was limited by its anharmonic decay into two acoustic
phonons.
Kash et al. (1987, 1988) and Tsen and Morkoç (1988a, b) used time-resolved
Raman scattering for the alloy system AlGaAs. The results for the lifetimes are
similar to those for pure GaAs; here, though, the phonon linewidths are broadened
owing to the disorder of the alloys and these inhomogeneous broadening effects need
to be considered. Secondly, although the dispersion relations of AlAs are different
from those of GaAs there is a similarity in decay times that is interesting and
unexpected. Tsen (1992) and Tsen et al. (1989) reported on the use of time-resolved
Raman studies of non-equilibrium LO phonons in GaAs-based structures.
Tsen et al. (1996, 1997, 1998) have studied the electron–phonon interactions in
GaN of würtzite structure via picosecond and sub-picosecond Raman spectroscopy.
Results on undoped GaN with an electron density of n = 5 × 1016 cm−3 showed
that the relaxation mechanism of the hot electrons is via the emission of LO phonons
and that the Fröhlich interaction is much stronger than the deformation-potential
interaction in that material. The measured lifetime was found to be 3 ps at 300 K and
5 ps at 5–25 K (Tsen et al., 1996, 1997, 1998). The electron–LO-phonon scattering
rate was seen to be an order of magnitude larger than that for GaAs and was
attributed to the much larger ionicity in GaN. These experiments also indicated that
the longitudinal phonons decay into a TO and an LO phonon or two TO phonons.
Raman investigations of phonon lifetimes have been reported by Bergman et al.
(1999) in GaN, AlN, and ZnO würtzite crystals. These lifetimes were obtained
from measured Raman linewidths using the uncertainty relation, after correcting
32 4 Raman properties of bulk phonons
for instrument broadening (Di Bartolo, 1969). These results demonstrate that the E 21
mode has a lifetime of 10 ps, an order of magnitude greater than that of the E 22 ,
E 1 (TO), A1 (TO) and A1 (LO) modes. This result was found to be true for samples
of high-quality GaN, AlN, ZnO as well as for AlN with a high level of impurities.
An explanation of the relative long lifetime of the E 22 phonons was given in terms
of factors including energy conservation constraints, density of final states, and
anharmonic interaction coefficients. The E 22 mode lies at the lowest energy of the
optical phonon modes in the würtzite dispersion curves (Nipko et al., 1998; Nipko
and Loong, 1998; Hewat, 1970) and only the acoustic phonons provide channels of
decay. At the zone edges, the acoustic phonons are equal to or larger than those of
the E 22 mode. Thus, for energy conservation to hold, the E 22 phonons have to decay
to acoustic phonons at the zone center, where their density is low.
A2 (TO) mode were inconclusive. Studies by Demangeot et al. (1998) concluded that
the A1 (LO), A1 (TO), and E 1 (TO) modes all exhibit one-mode behavior. However,
infrared reflectance experiments indicate that the E 1 (TO) mode displays two-mode
behavior (Wisniewski et al., 1998).
Yu et al. (1998) extended the modified random-element isodisplacement model
developed for zincblende structures by including the additional phonon modes and
the anisotropy of the würtzite structure. According to this model, the E 1 and the
A1 phonon modes should show one-mode behavior. In zincblende AlGaN crystals,
the results of Harima et al. (1999) indicate that the LO-phonon shows one-mode
behavior while the TO mode shows two-mode behavior.
Raman experiments using two ultraviolet wavelengths were performed by Alex-
son et al. (2000) on InGaN in the range 0 < x < 0.5. They investigated the A1 and
the E 2 phonons. These studies show a one-mode behavior of the A1 (LO) phonon
while the E 2 phonon demonstrates a two-mode characteristic.
The fact that the E 2 mode behaves differently from the E 1 and A1 modes is
not surprising, when one considers the specific atoms giving rise to the vibrations.
This in fact emerges from experiments on GaN würtzite films from natural GaN as
well as from GaN containing the isotope 15 N reported by Zhang et al. (1997). All
the A1 and the E 1 modes observed in the 15 N isotope were seen to shift to lower
frequencies. Niether the E 21 nor E 22 mode showed a similar shift; the E 21 mode was
essentially unaffected by the different isotopic mass. The authors thus concluded
that the E 21 vibration is due to the motion of the Ga atoms, which are heavy, alone
and, thus, that there is no frequency response to an isotropic change in the nitrogen
mass, which is considerably lighter.
(Scott et al., 1969), SiC (Klein et al., 1972) and, more recently, the nitrides discussed
in the next paragraph.
Here, we summarize the experimental Raman scattering results of phonon–
plasmon coupled modes in GaN, which as grown tends to be an n-type material
(Edgar, 1994); a higher carrier concentration is achieved via intentionally doping
the material. Due to this fairly high carrier concentration, the plasmons in GaN
are considered to be overdamped, similarly to those in SiC (Klein et al., 1972).
Experiments by Kozawa et al. (1998) in GaN films with a relatively low concentra-
tion of carriers show a broadening and a weakening of the intensity of the Raman
features as well as a shift to higher frequencies. The carrier concentrations obtained
by fitting the Raman lineshape as in Klein et al. (1972), Hon and Faust (1973), and
Irmer et al. (1983) give values similar to those obtained from Hall measurements
(Kozawa et al., 1998). Other experiments on GaN to study these effects at lower
free-carrier concentrations have been carried out by Wetzel et al. (1996) and Ponce
et al. (1996) and show similar results. Kirillov et al. (1996) studied the effect on
the Raman modes in the high-carrier-concentration limit. This study found that with
these higher carrier concentrations the Raman features corresponding to the upper
branch of the phonon–plasmon coupled modes are too broad to extract meaningful
information. Demangeot et al. (1997) carried out a study of the lower branch of
the coupled phonon–plasmon mode for GaN for higher carrier concentrations. The
broad Raman peak observed could be fitted by the models referenced above (Klein
et al., 1972; Hon and Faust, 1973).
Chapter 5
Occupation number representation
Oh mighty-mouthed inventor of harmonies.
Alfred, Lord Tennyson, 1863
35
36 5 Occupation number representation
pq2 1
Hq = + mωq2 u q2 , (5.1)
2m 2
where m is the mass of the oscillator, ωq is the frequency of the phonon, u q is the
displacement associated with it, and pq is its momentum. Introducing the operators,
aq and aq† ,
mωq 1
aq = uq + i pq (5.2)
2h̄ 2h̄mωq
and
mωq 1
aq† = uq − i pq , (5.3)
2h̄ 2h̄mωq
and
Thus, aq† acting on |Nq gives a new eigenstate with eigenvalue increased by 1
and aq acting on |Nq gives a new eigenstate with eigenvalue decreased by 1; that
is,
5.1 Phonon mode amplitudes and occupation numbers 37
aq† |Nq = n q + 1|Nq + 1 (5.8)
and
√
aq |Nq = n q |Nq − 1, (5.9)
where the eigenvalues are consistent with the relation Nq |Nq = n q |Nq with Nq =
aq† aq . The eigenstates |Nq are orthonormal, so that
Likewise,
√
Nq |aq |Nq = n q δ Nq ,Nq −1 . (5.12)
incoming waves and by ê∗q, j for outgoing waves. The factor e−iωt is common to
both incoming and outgoing phonons and it is generally included – along with
factors of E/h̄ associated with the carrier phases – in the integral over time that
appears in the Fermi golden rule. As a means of including the phase factors to
ensure proper accounting of energy and momentum as well as the appropriate unit
polarization vectors, (5.17) will now be written as a sum over all wavevectors q; the
appropriate non-temporal phase factors appear as multipliers of the corresponding
phonon operators:
1 h̄ iq·r
u(r) = √ aq e êq, j + aq† e−iq·r ê∗q, j
N q j=1,2,3 2mω q
1 h̄ †
=√ êq, j (aq + a−q ) eiq·r ≡ u(q) eiq·r .
N q j=1,2,3 2mωq q
(5.18)
where q is summed over all wavevectors in the Brillouin zone and N is the
number of unit cells in the sample. Chapter 7 will treat the role of dimensional
confinement in modifying the optical and acoustic phonon modes in nanostructures.
Two major modifications result: the phase space is restricted and the plane-wave
nature of the phonon is modified. Both of these effects will be treated by appropriate
modifications of equation (5.18): the changes in the phase space due to dimensional
confinement will be addressed by modifying the sum over q and the dimensional
confinement of the plane wave will be described by introducing suitable envelope
functions. Equations (5.8)–(5.18) will find many applications here and in Chapters
8, 9 and 10.
The factor h̄/2mωq in the final expression for u(r) ensures that the desired
Hamiltonian is consistent with the phonon mode amplitude. It is convenient in
practical calculations of u(r) to cast the amplitude constraints implied by [u q , pq ] =
i h̄ in the form of an integral:
h̄
∗ 1
u (r)u(r)d r =
3
√ êq · êq e−i(q −q)·r d 3 r
q q 2N m ω ω
q q
1 h̄
= , (5.19)
nm q 2ωq
In the literature, quantized fields – such as the displacement field u(r) – are
expressed in terms of aq and aq† in a number of different ways. Indeed, it is possible
to perform canonical transformations that essentially exchange the roles of u q and
pq . Specifically, from the definitions of aq and aq† in terms of u q and pq , it follows
that
h̄ †
pq h̄
uq = (aq + aq ) and = −i (a − aq† ),
2mωq mωq 2mωq q
(5.22)
where the primes have been added in preparation for the following substitutions:
aq → −iaq and aq† → +iaq† . With these substitutions it follows that
pq pq
u q → and → −u q . (5.23)
mωq mωq
Thus, the canonical transformation aq → −iaq has the result that the roles of
u q and pq /mωq are interchanged. Clearly, this canonical transformation leaves
the harmonic-oscillator Hamiltonian unchanged. This is, of course, true for this
Hamiltonian whether expressed in terms of u q and pq or in terms of aq and aq† .
As a result of this canonical transformation, the quantized fields – such as the
displacement field u(r) – may be expressed in terms of aq and aq† in a number
of different ways. Indeed, the literature testifies to the fact that all the different
forms are used widely. Evidently, factors of ±i as multipliers of aq and aq† do
not change the matrix element of any such field times its complex conjugate, since
(±i)(±i)∗ = 1; thus, quantum-mechanical transition rates are not affected by this
transformation, as should clearly be the case. There is one other invariance that
is used frequently to rewrite fields expressed in terms of aq and aq† in the most
convenient form for a particular application. Namely, in expressions where a sum
or integral over positive and negative values of q is present, substitutions of the
type aq eiq·r → a−q e−iq·r leave the expression unchanged if all other multiplicative
factors contain even powers of q. Again, the quantum-mechanical transition rates
are not affected by such a change of variable.
after H. Fröhlich, who formulated the first qualitatively correct formal description.
In this book, the potential energy associated with the Fröhlich interaction will be
denoted by φFr (r). Clearly the polarization P associated with polar-optical phonons
and the potential energy associated with the Fröhlich interaction, φFr (r), are related
by
In terms of the phonon creation and annihilation operators of Section 5.1, P(r)
may be written as
d 3 q iq·r
P(r) = ζ aq e eq, j + aq† e−iq·r e∗q, j (5.25)
j=1,2,3
(2π ) 3
where eq, j represents the polarization vector associated with P(r) and q is the
phonon wavevector; then, it follows that
d 3 q iq·r
4π∇ · P(r) = 4πiζζ aq e q · eq, j − aq† e−iq·r q · e∗q, j .
j=1,2,3
(2π ) 3
(5.26)
Consider the case of a polar crystal with two atoms per unit cell, such as GaAs.
Clearly, the dominant contribution to P(r) results from the phonon modes in which
the normal distance between the planes of positive and negative charge varies. Such
modes are obviously the LO modes since in the case of LO modes eq, j is parallel
to q. However, TO phonons produce displacements of the planes of charge such
that they remain at fixed distances from each other; that is, the charge planes ‘slide’
by each other but the normal distance between planes of opposite charge does not
change. So, TO modes make negligible contributions to P(r). For TO phonons,
eq, j · q = 0. Accordingly,
d 3 q iq·r
4π ∇ · P(r) = 4πiζζ aq e q − aq† e−iq·r q , (5.27)
(2π ) 3
and the potential energy associated with the Fröhlich interaction, φFr (r), is given by
d 3 q 1 iq·r
HFr = φFr (r) = −4πieζζ aq e − aq† e−iq·r , (5.28)
(2π ) 3 q
where φFr (r), has been denoted by HFr , the Fröhlich interaction Hamiltonian since
φFr (r) is the only term contributing to it.
The dependence of φFr (r) on q −1 is familiar from the Coulomb interaction; the
coupling constant ζ remains to be determined.
42 5 Occupation number representation
N e∗
P(r) = uq (r)
(∞)
N e∗ 1 h̄
= √
(∞) N q j=1,2,3 mM
2 ωLO
m+M
× aq eiq·r eq, j + aq† e−iq·r e∗q, j ), (5.29)
where the division by (∞) accounts for screening, and the normal-mode expression
(5.18) has been used for u(r). By noticing that
1 d 3q
√ ↔ (5.30)
V q (2π )3
so that
2
N e∗ 1
h̄ h̄ ωLO 1 1
ζ = √ = − ,
(∞) N mM 2ωLO 4π (∞) (0)
2 ωLO
m+M
(5.33)
and
d 3 q 1 iq·r
HFr = φFr (r) = −4πieζζ aq e − aq† e−iq·r
(2π ) q
3
2π e2 h̄ωLO 1 1 1 iq·r
= −i − aq e − aq† e−iq·r . (5.34)
V (∞) (0) q q
5.3 Acoustic phonons and deformation-potential interaction 43
The crystal potential may be expanded in powers of the displacements of the ions
from their equilibrium positions to yield a sum over quadratic and higher-order
terms. The quadratic terms, of course, represent the harmonic modes considered
at length in Chapter 5. The cubic and higher-order terms, containing products of
three or more displacements, are generally known as the anharmonic terms. These
anharmonic terms lead to modifications of the harmonic modes in the quadratic
approximation of the harmonic oscillator. Indeed, while the harmonic approximation
may be used to describe the phonon dispersion relations it cannot describe the decay
of phonon modes: the anharmonic interaction is necessary to describe the decay of
a phonon into other phonons. The leading term in the anharmonic interaction is the
cubic term and it may be written as
1
Hq, j;q , j ;q , j = √ P(q, j; q , j ; q , j )uq, j uq , j uq , j , (6.1)
N
where q, q , and q are the phonon wavevectors and j, j , and j designate the
polarizations of the phonon modes participating in the anharmonic interaction. N
is the number of unit cells in the crystal, and P(q, j; q , j ; q , j ) describes the
anharmonic coupling. The normal process where q − q − q = 0 is taken into
account in this analysis but the umklapp process where q − q − q = b = 0
is ignored since the normal process is expected to dominate for small phonon
wavevectors. From Section 5.1, the displacement is given by
45
46 6 Anharmonic coupling of phonons
1 h̄ iq·r
u(r) = √ aq e êq, j + aq† e−iq·r ê∗q, j
N q j=1,2,3 2mωq, j
1
=√ uq, j (6.2)
N q j=1,2,3
where, as defined in Chapter 5, the unit polarization vectors are denoted by êq, j
for incoming waves and by ê∗q, j for outgoing waves, and the phonon creation and
annihilation operators are given by aq† and aq , respectively.
(6.5)
1
nq = . (6.8)
e−h̄ωq /k B T − 1
To calculate the net rate of phonon decay through the Klemens channel it is
necessary to consider not only the decay rate but also the generation rate for the
reverse process, G. The total loss rate for the LO phonons is then given by − + G.
Writing as n q (n q + 1)(n q + 1) it follows that G is given by (n q + 1)n q n q
and the inverse lifetime takes the form (Bhatt et al., 1994; Klemens, 1966; Ferry,
1991)
1 π h̄ 2 U 3 1
a
= (1 + n q + n q )
τ q ,q
4N m 3 ωq ωq ωq
× δq,q +q δ(h̄ωq − h̄ωq − h̄ωq ). (6.9)
48 6 Anharmonic coupling of phonons
In the general case the sum over final states must include a sum over j and j
as well. Bhatt et al. (1994) applied this result to calculate the lifetime of the bulk
LO phonon in GaAs as a function of temperature and as a function of phonon
wavevector along the 100 and 111 directions at selected temperatures; these
results are shown in Figures 6.2 and 6.3.
the rates for both phonon absorption and emission. Clearly, the lifetimes of the
optical phonons are important in determining the total energy loss rate for such
carriers. As discussed in the last section, the LO phonons in GaAs and many
other polar materials decay into acoustic phonons through the Klemens channel.
Over a wide range of temperatures and phonon wavevectors, the lifetimes of LO
phonons in GaAs vary from a few picoseconds to about 10 ps (Bhatt et al., 1994), as
illustrated by Figures 6.2 and 6.3. Longitudinal-optical phonon lifetimes for other
polar semiconductors are also of this magnitude. As a result of the Klemens channel,
the ‘hot’ LO phonons decay into acoustic phonons in times of the order of 10 ps.
These LO phonons undergoing decay into acoustic phonons are not available for
absorption by the carriers and, as a result of the Klemens channel, the carrier energy
relaxation rate is influenced strongly by the LO phonon lifetime. Indeed, it follows
that the time required to modulate a semiconductor quantum-well laser through
the switching of the electronic current being used to pump the laser is limited by
the time it takes the carrier to thermalize in the active quantum-well region of the
laser.
Figure 6.3. Wavevector dependence of the lifetime of bulk LO phonons in GaAs for
wavevectors along the 100 (squares) and 111 (circles) directions at 77 K (upper
set of points) and 300 K (lower set of points). From Bhatt et al. (1994), American
Institute of Physics, with permission.
50 6 Anharmonic coupling of phonons
1 2
= h̄ωTO (ωLO − ωTO )3
τGaN 2πρLO ωLO vLA
3
where ρLO is the reduced-mass density, vLA is the group velocity, and qLA is the LA
wavevector associated with the LA frequency ωLO −ωTO ; see also Ridley and Gupta
(1991). By taking ≈ 108 cm−1 as for GaAs, Ridley obtained a zero-temperature
rate of 5.0 × 1011 s−1 . A full understanding of phonon decay in the wide-bandgap
III-V nitride materials is yet to be obtained. However, it is likely that the optical-
phonon lifetimes in the III-V nitrides will play an important role in the design of
electronic and optoelectronic devices fabricated from them. The Raman analysis of
anharmonic phonon decay rates given in Chapter 4 provides essential information
for the systematic design of such devices.
Chapter 7
Moreover, E(r) and P(r), in medium n are related through the dielectric suscepti-
bility, n (ω):
where
52
7.1 Dielectric continuum model of phonons 53
ω2 − ωLO,n
2
n (ω) = n (∞) , (7.5)
ω2 − ωTO,n
2
ω2 − ωLO,n,a
2 ω2 − ωLO,n,b
2
n (ω) = n (∞) , (7.6)
ω2 − ωTO,n,a
2 ω2 − ωTO,n,b
2
for a ternary polar material A y B1−y C, where the subscript a denotes frequencies
associated with the dipole pairs AC and the subscript b denotes frequencies
associated with the dipole pairs BC. As in subsection 2.3.1, the displacement field is
related to the fields E(r) and P(r) through the driven oscillator equation and through
the effective charge, en∗ : for a binary medium n,
An alternative and useful form of these equations for the case of a binary material
results straightforwardly from the relations of Appendix A. Indeed, it is shown in
Appendix A, equations (A.8) and (A.9), for a diatomic polar material that
1/2
V
ü = −ωTO
2
u+ (0) − (∞) ωTO E,
4π µN
(7.10)
µN 1/2 (∞) − 1
P= (0) − (∞) ωTO u + E.
4π V 4π
In the first equation it has been assumed that ü has a general form for the
time dependence and may not be simply sinusoidal in ω. As will become evident,
this pair of equations is well suited as the basis for an alternative method for
performing the calculations of subsection 7.3.1. Moreover, it provides a convenient
starting point for the derivation of the macroscopic equations describing optical
phonons in polar uniaxial materials (Loudon, 1964). Uniaxial materials such as
54 7 Continuum models for phonons
the hexagonal würtzite structures GaN, AlN, and Gax Al1−x N have relatively wide
bandgaps and are suited for high-temperature electronics and short-wavelength
optoelectronic devices. Loudon (1964) introduced a useful model for describing
the macroscopic equations of a uniaxial polar crystal by introducing one dielectric
constant associated with the direction parallel to the c-axis, , and another dielectric
constant associated with the direction perpendicular to the c-axis, ⊥ . In Loudon’s
model a separate set of Huang–Born equations is necessary for the phonon mode
displacements parallel to the c-axis, u , and perpendicular to it, u⊥ . For a medium
denoted by n it then follows straightforwardly that
V
ü⊥,n = −ωTO,⊥,n u⊥,n +
2
(0)⊥,n − (∞)⊥,n ωTO,⊥,n E⊥,n ,
4π µn N
µn N (∞)⊥,n − 1
P⊥,n = (0)⊥,n − (∞)⊥,n ωTO,⊥,n u⊥,n + E⊥,n ,
4π V 4π
2
ω − ωLO,⊥,n
2
⊥,n (ω) = ⊥,n (∞) , (7.11)
ω2 − ωTO,⊥,n
2
and
V
ü,n = −ωTO,,n u,n +
2
(0),n − (∞),n ωTO,,n E,n ,
4π µn N
µn N (∞),n − 1
P,n = (0),n − (∞),n ωTO,,n u,n + E,n ,
4π V 4π
2
ω − ωLO,,n
2
,n (ω) = ,n (∞) . (7.12)
ω2 − ωTO,,n
2
E(r) = −∇φ(r),
D(r) = E(r) + 4π P(r)
(7.13)
ρ + (ω)E (r)ẑ,
= ⊥ (ω)E ⊥ (r)ρ̂
∇ · D(r) = 0,
1 h̄ †
u(r)⊥() =√ êq, j,⊥() (aq + a−q ) eiq·r , (7.14)
N q j=1,2,3 2mωq
with q⊥ = q sin θ, where θ is the angle between q and the c-axis, which is taken as
the z-axis. Moreover, q = q cos θ and ê2q, j,⊥ + ê2q, j, = 1. Hence
2π h̄ †
êq, j,⊥() (aq + a−q )
V ωq
(0)⊥(),n − (∞)⊥(),n ωTO,⊥(),n
= (−i)q⊥() φ(q), (7.18)
ωTO,⊥(),n
2 − ωq2
so that
2π h̄ † 2
(aq + a−q )
V ωq
[(0)⊥ − (∞)⊥ ] ωTO,⊥ 2
= −q 2
sin θ φ 2 (q),
2
(ω⊥,TO
2 − ωq2 )2
[(0) − (∞) ] ωTO,
2
−q 2
cos θ φ 2 (q),
2
(7.19)
(ωTO,
2 − ωq2 )2
and
2π h̄ †
φ(q) = −i (aq + a−q )(ωTO,⊥
2
− ωq2 )(ωTO,
2
− ωq2 )
V q 2 ωq
× [(0)⊥ − (∞)⊥ ]ωTO,⊥
2
(ωTO,
2
− ωq2 )2 sin2 θ
−1/2
+ [(0) − (∞) ]ωTO,
2
(ωTO,⊥
2
− ωq2 )2 cos2 θ . (7.20)
56 7 Continuum models for phonons
Thus
†
H= (−e)φ(q) eiq·r (aq + a−q )
q
2π e2 h̄ 1 iq·r †
=i e (aq + a−q )(ω⊥,TO
2
− ωq2 )(ωTO,
2
− ωq2 )
q V ωq q
× [(0)⊥ − (∞)⊥ ]ωTO,⊥
2
(ωTO,
2
− ωq2 )2 sin2 θ
−1/2
+ [(0) − (∞) ]ωTO,
2
(ω⊥,TO
2
− ωq2 )2 cos2 θ
1/2
4π e2 h̄V −1 1 iq·r †
=i e (aq + a−q ).
q (∂/∂ω)[(ω)⊥ sin θ + (ω) cos θ]
2 2 q
(7.21)
In the isotropic case, (ω)⊥ = (ω) and this result must reduce to the
expression, (5.34), obtained in Section 5.2 for the interaction Hamiltonian de-
scribing the carrier–LO-phonon interactions. Indeed, since the general form of the
Lyddane–Sachs–Teller relation implies that
2 − ω2
ωTO LO 1 1 1/2
√ = −ωLO − , (7.22)
ωTO (0) − (∞) (∞) (0)
the Hamiltonian for the uniaxial case reduces to (5.34) upon taking (ω)⊥ = (ω)
and ωq = ωLO .
T = Y e, (7.23)
∂ 2u ∂ Tx x ∂ Tyx ∂ Tzx ∂
ρ = + + = (λ + µ) + µ∇ 2 u,
∂t 2 ∂x ∂y ∂z ∂x
∂ 2v ∂ Tx y ∂ Tyy ∂ Tzy ∂
ρ 2 = + + = (λ + µ) + µ∇ 2 v, (7.33)
∂t ∂x ∂y ∂z ∂y
∂ 2w ∂ Tx z ∂ Tyz ∂ Tzz ∂
ρ 2 = + + = (λ + µ) + µ∇ 2 w,
∂t ∂x ∂y ∂z ∂z
where, as usual,
∂2 ∂2 ∂2
∇2 ≡ + + . (7.34)
∂x2 ∂ y2 ∂z 2
Two alternative forms of the three-dimensional force equations are encountered
frequently in the literature. The first of these is derived by writing the components
of u(x, y, z) = (u 1 , u 2 , u 3 ) as u α , α = 1, 2, 3; it then follows that the three force
equations may be rewritten as
∂ 2uα ∂ Tαβ
ρ = , (7.35)
∂t 2 ∂rβ
where
Tαβ = λSαα δαβ + 2µSαβ . (7.36)
In these equations, the subscripts α and β run over 1, 2, 3 (corresponding to
x, y, z). A repeated index in a term implies summation. δαβ is the Kronecker
delta function. In a second alternative form the three force equations are written
straightforwardly as the single vector equation
∂ 2u
= ct2 ∇ 2 u + (cl2 − ct2 ) grad , (7.37)
∂t 2
where it is now clear that
= ∇ · u = div u. (7.38)
Here, ct and cl are the transverse and longitudinal sound speeds and we have
λ λ + 2µ
ct2 = and cl2 = . (7.39)
ρ ρ
In physical acoustics the solutions for the displacement fields are frequently
specified in terms of two potential functions, a scalar potential φ and a vector
potential = (x , y , z ), through
∂φ ∂x ∂ y
u= + − ,
∂x ∂y ∂z
∂φ ∂x ∂z
v= + − , (7.40)
∂y ∂z ∂x
∂φ ∂ y ∂x
w= + − ,
∂z ∂x ∂y
60 7 Continuum models for phonons
1 ∂ 2φ λ + 2µ
∇ 2φ = , cl2 = ,
cl2 ∂t 2 ρ
(7.41)
1 ∂ 2 i λ
∇ i = 2
2
, i = x, y, z, ct2 = .
ct ∂t 2 ρ
The scalar potential φ corresponds to the ‘irrotational’ part of the solution and
the vector potential corresponds to any remaining ‘rotational’ fields. In the literature
the irrotational solutions are also referred to as the longitudinal, compressional, or
dilatational solutions. Moreover, seismologists frequently refer to these solutions as
P waves. Likewise, the rotational vector-potential solutions based on i are iden-
tified as transverse, shear, distortional, or equivoluminal solutions. In seismology
these solutions are commonly identified as S waves.
Herein, the irrotational solutions will generally be referred to as longitudinal
modes and the corresponding sound speed will be denoted by cl . Likewise, the
rotational fields will be denoted as transverse modes and the associated sound speed
will be denoted by ct . The principal interest in this book is on using the longitudinal
and transverse solutions of the elastic continuum model to describe the longitudinal
acoustic (LA) and transverse acoustic (TA) phonons in nanostructures. Hence,
the notation adopted herein is that corresponding most naturally to the solid-state
community’s descriptions of phonons as longitudinal and transverse.
Every year, experimental observations of acoustic modes in nanostructures are
being reported in the literature in increasing numbers. Examples of such experi-
mental studies include the works of Nabity and Wybourne (1990a, b), Seyler and
Wybourne (1992), and Sun et al. (1999). Wybourne and his coworkers report
measurements of confined acoustic phonons in very thin metallic foils and wires
on dielectric substrates. Sun et al. (1999) presented data on folded acoustic modes
in semiconductor superlattices. In view of such observations, Section 7.6 presents
extensions of the elastic continuum theory of this section to the case of acoustic
modes in dimensionally confined structures and Chapter 9 focuses on carrier–
acoustic-phonon scattering rates in both bulk materials and dimensionally confined
nanostructures.
i (r) = i (q, z) e−iq·ρρ , (7.42)
q
where ρ ≡ (x, y) and q is the two-dimensional wavevector in the x y-plane, that is,
q = qx x̂ + q y ŷ, where x̂ and ŷ are unit vectors. Then
Ei (r) = −∇i (r) = Ei (q, z) e−iq·ρρ ,
q
(7.43)
Pi (r) = i (ω)Ei (r) = Pi (q, z) e−iq·ρρ .
q
Following the concepts of Section 5.1, the mode normalization condition requires
that the energy of a phonon of mode q is h̄ωq ; for the case of a single interface at
z = 0 separating two layers n and m, this condition is
62 7 Continuum models for phonons
∞ ∗
L2 (µn nn )un (q, z) · (µn nn )un (q, z) dz
0
0 ∗
+y L2 (µm nm )um,a (q, z) · (µm nm )um,a (q, z) dz
−∞
0 ∗
+ (1 − y) L2 (µm nm )um,b (q, z) · (µm nm )um,b (q, z) dz
−∞
h̄
= . (7.44)
2ωq
so that
!
c(iqe−qz q̂ − qe−qz ẑ) z ≥ 0,
Ei (q, z) =
c(iqeqz q̂ + qeqz ẑ) z ≤ 0,
! (7.46)
i c(iqe−qz q̂ − qe−qz ẑ) z ≥ 0,
Pi (q, z) =
i c(iqeqz q̂ + qeqz ẑ) z ≤ 0.
region 1 region 2
Figure 7.1. A ternary–binary structure. From Kim and Stroscio (1990), American
Institute of Physics, with permission.
7.3 Optical modes in dimensionally confined structures 63
where q̂ is the unit vector specifying the direction of q ≡ (qx , q y ). Let material n
be a binary layer filling the space z ≥ 0, region 1, and material m be a ternary layer
filling the space z ≤ 0, region 2, as illustrated in Figure 7.1. Then, for the right-hand
medium, material 1, using the notation of Appendix B, it follows that
thus
1 c(iqe−qz q̂ − qeqz ẑ)
u1 (q, z) = . (7.48)
n1 e1∗ [1 + α1 µ1 (ω01
2 − ω2 )/e∗2 ]
1
For material 2, there are two driven-oscillator equations, one for the AC pair,
denoted by a, and one for the BC pair, denoted by b:
Then
2 c(iqeqz q̂ + qeqz ẑ)
Elocal (q, z) = ∗2 ∗2 ,
ye2a (1 − y)e2b
n2 + + α2
µ2,a (ω02,a
2 − ω2 ) µ2,b (ω02,b2 − ω2 )
∗
e2,a(b)
u2,a(b) (q, z) =
µ2,a(b) (ω02,a(b)
2 − ω2 )
2 c(iqeqz q̂ + qeqz ẑ)
× ∗2 ∗2 .
ye2a (1 − y)e2b
n2 + + α2
µ2,a (ω02,a
2 − ω2 ) µ2,b (ω02,b2 − ω2 )
(7.51)
64 7 Continuum models for phonons
With these expressions for u1 (q, z) and u2,a(b) (q, z), the normalization condition,
with n = 1 and m = 2, yields
h̄ 1 µ1 n1 χ12 q
=
2ω L 2 c2 {n1 e1∗ [1 + α1 µ1 e1∗−2 (ω01
2 − ω2 )]}2
∗ 2
e
+ µ2a n2 y 2a (ω02,a 2
− ω2 ) q
µ2,a
! 2 ∗2 ∗2 2 "
2 ye2a (1 − y)e2b
× + + α2
n22 µ2,a (ω02,a
2 − ω2 ) µ2,b (ω02,b
2 − ω2 )
∗ 2
e
+ µ2b n2 (1 − y) 2b (ω02,b 2
− ω2 ) q
µ2,b
! 2 ∗2 ∗2 2 "
2 ye2a (1 − y)e2b
× + + α2 ,
n22 µ2,a (ω02,a
2 − ω2 ) µ2,b (ω02,b
2 − ω2 )
(7.52)
where the integral in the first term has been performed using
∞ ∞
dz(iqe−qz q̂ − qe−qz z)∗ · (iqe−qz q̂ − qe−qz ẑ) = dz 2q 2 e−2qz = q,
0 0
(7.53)
and the second and third integrals have been performed using
0 ∞
dz(iqeqz q̂ + qeqz ẑ)∗ · (iqeqz q̂ + qeqz ẑ) = dz 2q 2 e−2qz = q. (7.54)
−∞ 0
1
ωTO,n
2
= ω0,n
2
+ 13 ωplasma,n
2
, (7.55)
1 − 3 π nn αn
4
1 1
ωLO,n
2
− ωTO,n
2
= 23 ωplasma,n
2
+ 13 ωplasma,n
2
1 + 3 π n n αn
8
1 − 3 π n n αn
4
ωplasma,n
2
= .
(1 + 83 π nn αn )(1 − 43 π nn αn )
Here the subscript n represents either material 1 or material 2. In these relations, the
plasma frequency squared, ωplasma,n
2 , is given by
7.3 Optical modes in dimensionally confined structures 65
∗2
ωplasma,n,a(b)
2
= 4π nn en,a(b) /µn,a(b) . (7.56)
Thus,
(r) = (q, z) e−iq·ρρ = ce−q|z| e−iq·ρρ
q q
4π h̄ 1/2 ∂ E 1 (ω) ∂ E 2 (ω)
−1/2
=− + eq|z| e−iq·ρρ , (7.58)
q L 2q ∂ω ∂ω
2 2 (7.59)
ω − ωLO,2,a
2 ω − ωLO,2,b
2
2 (ω) = 2 (∞) .
ω2 − ωTO,2,a
2 ω2 − ωTO,2,b
2
where êq, j and ê∗q, j of equation (6.2) have been taken as unit vectors in the
longitudinal direction, since the IF phonon modes considered here are longitudinal
optical (LO) phonons. The dispersion relation for this optical phonon mode is
given from the requirement that the normal components
of the electric
displacement
field be continuous at z = 0, that is, 2 (ω)E 2,z z=0 = 1 (ω)E 1,z z=0 . From this
condition, it follows immediately that the frequencies of the IF optical phonons must
satisfy 1 (ω) + 2 (ω) = 0.
This result is similar to that for a bulk semiconductor, where the optical phonon
frequencies must satisfy (ω) = 0. Moreover, since this is the condition necessary
for the propagation of any longitudinal electromagnetic disturbance, it was expected
66 7 Continuum models for phonons
that the frequencies of longitudinal optical phonons should satisfy this dispersion
relation.
In the case of the two-region, single-heterointerface structure, the IF longitudinal
optical phonon frequencies depend on both 1 (ω) and 2 (ω). Therefore, the IF
optical phonon mode is a joint mode of both materials. Indeed, the electrostatic
phonon potential for this mode falls off exponentially with distance from the
heterointerface in both materials and, of course, the IF phonon electrostatic potential
has only one common value at the interface. Hence, such an IF optical phonon mode
must have one common electrostatic phonon potential throughout the two-region
heterostructure. It is therefore not surprising that the frequency of such an interface
LO phonon mode depends on the dielectric constants of both materials. Clearly,
the IF optical phonon mode described by HIF is a joint mode of both materials. In
general, IF optical phonon modes are joint modes of all the materials in a given
heterostructure. This property is manifest throughout this book.
Clearly, the IF optical phonon modes do not form a complete set of optical
phonon modes for the case of two semi-infinite regions joined at a single hetero-
interface. Indeed, these IF modes vanish exponentially as |z| → ∞ and it is
clear that bulk-like optical phonons must exist in regions significantly removed
from the heterointerface. These additional modes are known as half-space modes.
For a structure with a single heterointerface these half-space modes have been
given by Mori and Ando (1989) for the case where the two semi-infinite regions
are composed of binary semiconductors. Mori and Ando also gave the full set
of optical phonon modes for double-heterointerface structures, where one type of
binary semiconductor layer with interface planes situated at −a/2 and +a/2 is
bounded by two semi-infinite regions of a different binary semiconductor. Appendix
C provides a summary of the phonon modes of the double-heterointerface structure
for three models, the slab modes of the dielectric continuum model with electrostatic
boundary conditions, as in Mori and Ando (1989), and two other models discussed
widely in the literature. Appendix C also discusses Raman measurements useful
in understanding the behavior of these modes (Sood et al., 1985). The modes
arising from the second and third models are known as the guided modes and the
reformulated (or Huang–Zhu) modes. As discussed in Appendix C, all these sets
of phonon modes predict the same intrasubband and intersubband scattering rates
provided that each set is composed of a complete, orthogonal set of phonon modes.
inside the slab i.e., in the range (−a, +a). #Outside the slab, where = 1, the
solutions have the form φ(z) = φ± exp(± qx2 + q y2 z), where the positive sign
applies for z ≤ −a and the negative sign applies for z ≥ +a. The constants φ1 ,
φ2 , φ+ , and φ− are determined by the usual boundary conditions that the tangential
component of E and the normal component of D are continuous at z = ±a. From
these conditions it is seen that φ± = 0 and it is thus clear that for this mode
φ(z), E(r), and D(r) are zero in the regions surrounding the slab; in particular φ(z)
vanishes at the surfaces of the layer, where z = ±a. For z in the range (−a, +a),
the boundary conditions may be satisfied by taking either φ1 = 0 or φ2 = 0, so that
there are two solutions corresponding to the two polarization vectors:
φ2 iq ·ρρ $ mπ mπ mπ %
+ (r) =
Pm e i iq a cos z − ẑ sin z
4πa 2a 2 2a
m = 1, 3, 5, . . . ,
φ $ mπ mπ mπ %
(7.63)
1 iq ·ρρ
Pm
− (r) = e i iq a sin z + ẑ cos z
4πa 2a 2 2a
m = 2, 4, 6, . . . ,
where ẑ is the unit vector in the z-direction. Of course, ∇ · D(r) = 0 implies that
∇ 2 φ(r) = −∇ · E(r) = +4π ∇ · P(r). These standing modes are now widely known
as the confined optical phonon modes in a slab. They exist for m running from 1 to
some maximum number N2a ; the values of m must terminate at N2a , the number of
68 7 Continuum models for phonons
unit cells in thickness 2a, since the continuum model adopted here must fail when
the number of half-wavelengths in 2a becomes equal to or greater than the number
of unit cells in the same thickness.
The remaining solution corresponds to the case where (ω) = 0 inside the slab
# #
φ(x) = φ1 exp + qx2 + q y2 z + φ2 exp − qx2 + q y2 z .
#
As before, when |z| ≥ a, the solution is φ(z) = φ± exp(± qx2 + q y2 z). The
boundary conditions then restrict the modes to the forms
# 1 − iq ·ρρ
P0+ (r) = −φ1 qx2 + q y2 e i
$ # 4π # %
× i q̂ cosh qx2 + q y2 z + ẑ sinh qx2 + q y2 z ,
(7.64)
# 1 − iq ·ρρ
P0− (r) = −φ1 qx2 + q y2 e i
$ # 4π # %
× i q̂ sinh qx2 + q y2 z + ẑ cosh qx2 + q y2 z ,
where q̂ is the unit two-dimensional wavevector. These last two modes describe the
so-called IF optical phonon modes in the polar semiconductor slab of thickness 2a.
The boundary conditions imply that the frequencies for these modes are solutions of
1 + (ω) #
= ± exp −2 qx2 + q y2 a , (7.65)
1 − (ω)
#
[(0) + 1] ∓ [(0) − 1] exp − qx2 + q y2 a
ω± = ωTO # ,
2 2
(7.66)
[(∞) + 1] ∓ [(∞) − 1] exp − qx2 + q y2 a
where the plus sign corresponds to the even mode, the minus sign to the odd mode.
As pointed out by Licari and Evrard (1977), this continuum model is capable
of predicting both the confined LO phonons and the interface IF optical phonons
because for both of these modes there exists a polarization charge density. In
particular, both ρ = −∇ · P, the volume charge density, and σ = −P · n̂, the
surface charge density, contribute to the confined LO modes; here, n̂ is the unit
vector normal to the surface and pointing into the vacuum. For the IF modes, only
σ makes a contribution. Clearly, in this model the polarization charge acts as the
source of the fields associated with these phonon modes. Transverse modes are not
predicted by this continuum approach since for such modes ∇ · P = 0 and P · n̂ = 0.
Licari and Evrard (1977) used this model to study the effects of electronic
polarizability on the phonon modes and they derive conditions for the slab which
7.3 Optical modes in dimensionally confined structures 69
are equivalent to Wendler’s conditions for the two-layer system described in Ap-
pendix B. Licari and Evrard also used their model to construct the normalized polar-
ization eigenvectors and frequencies for the phonon modes of the dielectric slab.
Moreover, they constructed the Hamiltonian for the electron–polar-optical-phonon
interaction and showed that the correct harmonic oscillator energy is recovered when
the eigenvectors of the slab are used to evaluate the Hamiltonian; in particular, it
can be shown that the normal modes are consistent with the harmonic oscillator
energy of Section 5.1. Finally, Licari and Evrard presented a very enlightening
physical derivation of the electron–phonon interaction Hamiltonian for a slab by
applying boundary conditions to the electron–phonon interaction Hamiltonian for
a bulk semiconductor. Specifically, starting with expression for the bulk Fröhlich
interaction, which we take as the expression (5.34) derived in Section 5.2,
! "1/2
2π e2 h̄ωLO 1 1 1
HFr = −i − †
(aq + a−q ) e−iq·r
V (∞) (0) q q
! 2π e2 h̄ωLO 1 1
"1/2
= −i − †
(aq + a−q ) e−iq·r
q V q 2 (∞) (0)
= −i Vq (aq + a−q †
) e−iq·r , (7.67)
q
Licari and Evrard took q = (q , qz ) and split the sum over q into a sum over q and
a sum over qz > 0:
HFr = Vq e−iq ·r e−iqz z (aq ,qz + a−q
†
,−qz
)
q ,qz >0
†
+ eiqz z (aq ,−qz + a−q ,qz
). (7.68)
Then, using eiθ = cos θ + i sin θ to write e±iqz z in terms of sines and cosines,
√
HFr = 2 Vq e−iq ·r
q ,qz >0
† †
× {cos qz z[a+ (q ) + a+ (−q )] + sin qz z [a− (q ) + a− (−q )]},
(7.69)
where
1 −i
a+ (q ) = √ (aq ,qz + aq ,−qz ), a− (q ) = √ (aq ,qz − aq ,−qz ).
2 2
(7.70)
† †
The operators a+ (−q ) and a− (−q ) are given by taking the adjoints. These oper-
ators describe phonons which propagate as plane waves in the x- and y-directions
but as standing modes in the z-direction. Indeed, since qz = mπ/2a the Fröhlich
Hamiltonian for the two-dimensional slab takes the form
70 7 Continuum models for phonons
1/2
4π e2 h̄ωLO 1 1
HFr = − e−iq ·r
V (∞) (0) q
!
cos(mπ/2a)z †
× [a (q ) + am+
2 + (mπ/2a)2 ]1/2 m+
(−q )]
m=1,3,5...
[q 2
x + q y
"
sin(mπ/2a)z †
+ [am− (q ) + am− (−q )] .
m=2,4,6...
[qx2 + q y2 + (mπ/2a)2 ]1/2
(7.71)
This Hamiltonian vanishes for z = ±a, as it must since the Fröhlich interaction
Hamiltonian is given by −eφ, as explained in Section 5.2, and since φ(±a) = 0 for
the potential describing the fields associated with phonon modes in the dielectric
slab. This heuristic derivation makes manifest the fact that the confined phonon
modes in the slab located between −a and +a are standing modes with an integer
number of half-wavelengths confined within the slab. This Hamiltonian does not
contain the contributions of the IF optical phonons in the slab since it satisfies only
the boundary conditions for the confined optical phonon modes at z = ±a, namely
HFr (a) = −eφ(±a) = 0. As shown by Licari and Evrard (1977), the Fröhlich
interaction Hamiltonian for the IF optical phonon modes in the dielectric slab is
! "1/2
2π e2 h̄ωTO
HFr =− [(0) − (∞)]
L2
# 1/2
#
sinh 2 qx2 + q y2 a
e− qx +q y a
2 2
× e−iq ·r #
q qx2 + q y2
& #
× G + qx2 + q y2 , z [a0+ (q ) + a0+ †
(−q )]
# '
+ G − qx2 + q y2 , z [a0− (q ) + a0−
†
(−q )] , (7.72)
where
# ( #
# cosh qx2 + q y2 z cosh qx2 + q y2 a
G + qx2 + q y2 , z = #
−2 qx2 +q y2 a
[(∞) + 1] − [(∞) − 1] e
# 1/4
[(∞) + 1] − [(∞) − 1] e−2 qx2 +q y2 a
× # ,
−2 qx2 +q y2 a
[(0) + 1] − [(0) − 1] e
(7.73)
and
# ( #
# sinh qx2 + q y2 z sinh qx2 + q y2 a
G − qx2 + q y2 , z = #
−2 qx2 +q y2 a
[(∞) + 1] + [(∞) − 1] e
# 1/4
[(∞) + 1] + [(∞) − 1] e−2 qx2 +q y2 a
× # .
−2 qx2 +q y2 a
[(0) + 1] + [(0) − 1] e
(7.74)
7.3 Optical modes in dimensionally confined structures 71
In the result (7.72) the phonon creation and annihilation operators are not
summed over m; since there are just two IF optical phonon modes, the subscript
m on the creation and annihilation operators for the confined optical phonon modes
is replaced by 0 and the plus sign in the subscript corresponds to the even mode
while the minus sign corresponds to the odd mode. The modes discussed by Licari
and Evrard are recognized as the optical modes of a double-interface heterostructure
in the special case where = 1 outside the region bounded by the two heteroint-
erfaces. As mentioned previously, Appendix C compares the three frequently used
complete sets of optical phonon modes for a double-interface heterostructure for
the case where all the material layers are polar semiconductors. As may be seen
straightforwardly, the modes considered by Licari and Evrard correspond to the slab
modes derived with electrostatic boundary condition for the special cases where the
quantum well is bounded by a vacuum, so that = 1.
it follows that
√
nn µn u⊥,n (q)2 + √nn µn u,n (q)2 = h̄ 1 , (7.77)
2ω V
provides the necessary generalization to the case of a uniaxial crystal. From
Section 7.1, the equations governing u⊥,n and u,n may be written as
1 (0)⊥(),n − (∞)⊥(),n
u⊥(),n = √ ωTO,⊥(),n E⊥,n , (7.78)
4π µn nn ωTO,⊥(),n
2 − ω2
72 7 Continuum models for phonons
where the time dependence of each displacement has been assumed to be of the form
eiωt . Then, it follows that
√
nn µn u⊥,n (r)2 + √nn µn u,n (r)2
so that
√
nn µn u⊥,n (r)2 + √nn µn u,n (r)2
1 1 ∂(ω)⊥,n 2 1 1 ∂(ω),n 2
= E⊥,n + E,n . (7.82)
4π 2ω ∂ω 4π 2ω ∂ω
Using this identity, the normalization condition becomes
1 1 ∂(ω)⊥,n 2
1 1 ∂(ω),n 2
h̄
E⊥,n + E,n dr = .
4π 2ω ∂ω 4π 2ω ∂ω 2ω
(7.83)
As discussed in Section 7.3, the normalization condition for the case where there is
dimensional confinement in only the z-direction is then
1 1 ∂(ω)⊥,n 2
1 1 ∂(ω),n 2
h̄
L 2
E⊥,n + E,n dz = ,
4π 2ω ∂ω 4π 2ω ∂ω 2ω
(7.84)
where the potential, (q, z), is associated with E⊥,n and E,n . The form of
normalization (7.84) is particularly convenient for optical modes in the dielectric
7.3 Optical modes in dimensionally confined structures 73
continuum model since the phonons may be described in terms of the associated
electric fields and potentials (Kim and Stroscio, 1990; Lee et al., 1998; Komirenko
et al., 2000a).
In this subsection, the quantization condition for a uniaxial crystal will be
applied to determine the electron–optical-phonon interaction Hamiltonian for in-
terface phonons in a single-heterointerface structure. Appendix D summarizes the
electron–optical-phonon interaction Hamiltonians for all the optical phonon modes
in single- and double-heterointerface uniaxial crystals for the case where the c-axis
is perpendicular to the heterointerface(s).
As an illustration, consider the interface optical mode in a würtzite structure
composed of two semi-infinite regions separated by a single heterointerface situated
at z = 0. The c-axis is taken to be normal to the heterointerface. In the region
z < 0 the dielectric functions are (ω)⊥(),2 and in the region z > 0 they
are (ω)⊥(),1 . Each of the four functions, (ω)⊥(),1(2) obeys the generalized
Lyddane–Sachs–Teller relation; in particular, each is less than zero for frequencies
selected from one range of interest from among the four ranges ωTO,⊥(),1(2) <
ω < ωLO,⊥(),1(2) . In addition, for ωLO,⊥(),1(2) < ω < ωTO,⊥(),1(2) , it is clear
that (ω)⊥(),1(2) > 0 for the range corresponding to the one dielectric function of
interest. For each of these four ranges, the positivity or negativity properties are as
for zincblende crystals.
However, for uniaxial crystals the products (ω)⊥,1 (ω),1 and (ω)⊥,2 (ω),2
may be either positive or negative since, for a material n, (ω)⊥,n and (ω),n may
have different signs depending on the overlap of the two regions ωTO,⊥,n < ω <
ωLO,⊥,n and ωTO,,n < ω < ωLO,,n .
For binary zincblende heterostructures such as GaAs/AlAs, the IF modes exist
for the frequencies in the two ranges ωTO,AlAs < ω < ωLO,AlAs and ωTO,GaAs <
ω < ωLO,GaAs . Since these two ranges do not overlap, the frequency condition
for the existence of IF modes in such zincblende heterostructures is typified by
GaAs (ω)AlAs (ω) < 0 for all allowed IF mode frequencies ω. Such a simple
characterization is not possible for uniaxial crystals.
As will become obvious, this situation leads to significant differences in the
optical phonon modes in würtzite and zincblende structures. For the zincblende case
!
ρ Aeκ2 z z<0
φ(r ) = e iq·ρ
× −κ (7.86)
Be 1 z z>0
From the continuity of the tangential component of the electric field at z = 0, it then
follows that A = B. From the continuity of the normal component of the electric
74 7 Continuum models for phonons
(ω)⊥,1 2
κ12 = q (ω),1 (ω)⊥,1 > 0,
(ω),1
(7.88)
(ω)⊥,2 2
κ22 = q (ω),2 (ω)⊥,2 > 0,
(ω),2
so that
(ω)⊥,1 (ω)⊥,2
,1 κ1 + ,2 κ2 = (ω),1 q + ,2 (ω) = 0. (7.89)
(ω),1 (ω),2
Accordingly,
− (ω)⊥,1 (ω),1 + (ω)⊥,2 (ω),2 = 0 (ω),1 < 0 and (ω),2 > 0,
+ (ω)⊥,1 (ω),1 − (ω)⊥,2 (ω),2 = 0 (ω),1 > 0 and (ω),2 < 0,
(7.90)
and it follows that (ω)⊥,1 (ω),1 = (ω)⊥,2 (ω),2 . Thus
!
Aeκ2 z z<0
φ(r ) = φ0 eiq·ρρ ×
Be−κ1 z z>0
(7.91)
ρ
exp( (ω)⊥,2 /(ω),2 qz) z<0
= φ0 e iq·ρ
×
exp(− (ω)⊥,1 /(ω),1 qz) z > 0.
Now E⊥ and E are given by the appropriate gradients of φ0 , and the integrals
2 2
needed to calculate the normalization condition are related to E⊥,n and E,n
through
0 0
2 q 2 1 (ω)⊥,2
E⊥,2 dz = φ q2 2
e 2 dz =
2κ z
φ0 =
2
qφ02 ,
0
−∞ −∞ 2κ 2 2 (ω),2
∞ ∞
q2 2 1 (ω)⊥,1
E⊥,1 2 dz = φ 2 q 2 e −2κ1 z
dz = φ0 = qφ02 ,
0
0 0 2κ1 2 (ω),1
0 0
2 κ 1 (ω)⊥,2
E,2 dz = φ 2 κ 2 2
e2κ2 z dz = φ02 = qφ02 ,
0 2
−∞ −∞ 2 2 (ω),2
∞ ∞
2 κ 1 (ω)⊥,1
E,1 dz = φ 2 κ 2 e−2κ1 z dz = φ02 =
1
qφ02 .
0 1
0 0 2 2 (ω),1
(7.92)
7.3 Optical modes in dimensionally confined structures 75
Then we have
1 h̄
L 22ω
1 1 ∂(ω)⊥,n 2
1 1 ∂(ω),n 2
= E⊥,n + E,n dz
4π 2ω ∂ω 4π 2ω ∂ω
0 ∞
1 1 ∂(ω)⊥,2 2 1 1 ∂(ω)⊥,1 2
= E⊥,2 dz + E⊥,1 dz
4π 2ω −∞ ∂ω 4π 2ω 0 ∂ω
0 ∞
1 1 ∂(ω),2 1 1 ∂(ω),1 2
E,2 dz + E,1 dz
2
+
4π 2ω −∞ ∂ω 4π 2ω 0 ∂ω
1 1 q ∂⊥,1 q ∂,1 κ1 ∂⊥,2 q ∂,2 κ2
= + + + φ02 ,
4π 2ω 2 ∂ω κ1 ∂ω q ∂ω κ2 ∂ω q
(7.93)
and it follows that
−1
4π h̄ 2 ∂⊥,1 q ∂,1 κ1 ∂⊥,2 q ∂,2 κ2
φ02 = 2 + + +
L q ∂ω κ1 ∂ω q ∂ω κ2 ∂ω q
(7.94)
and
HIF = −(e(q, z)) eiq·ρρ (aq + a−q
†
)
q
4π e2 h̄
= (2q −1 )1/2
q L2
−1/2
∂⊥,1 q ∂,1 κ1 ∂⊥,2 q ∂,2 κ2
× + + +
∂ω κ1 ∂ω q ∂ω κ2 ∂ω q
× eiq·ρρ (aq + a−q
†
)
√
− (ω)⊥,1 /(ω),1 qz
e√ z > 0,
× (7.95)
e (ω)⊥,2 /(ω),2 qz z < 0,
to the case of würtzite superlattices. For the specific case of a GaN/AlN superlattice
with the c-axis normal to the heterointerfaces and with a superlattice period, d, as
shown in Figure 7.2, the requirement of periodicity and the boundary conditions
imposed on the fields for the quantum-well case lead to dispersion relations of the
form
a1 (ω) cosh[γ1 (ω)d/2] + a2 (ω) sinh[γ1 (ω)d/2] = 0, (7.97)
the antisymmetric modes and
a1 (ω) sinh[γ1 (ω)d/2] + a2 (ω) cosh[γ1 (ω)d/2] = 0, (7.98)
for the symmetric modes. In these dispersion relations we have
# (
γ1 (ω) = q⊥ (ω)⊥,1 (ω),1 ,
a1 (ω) = sign[(ω),1 ] (ω)⊥,1 (ω),1 , (7.99)
a2 (ω) = sign[(ω),2 ] (ω)⊥,2 (ω),2 .
The dispersion relations (7.97), (7.98) are depicted in Figures 7.3 and 7.4 for the
AlN(5 nm)/GaN(5 nm) superlattice along with quasi-confined modes. In the limit
q⊥ d → ∞, these dispersion relations reduce to the condition (ω)⊥,1 (ω),1 =
(ω)⊥,2 (ω),2 , as they must since in the short-wavelength limit the frequencies
of these modes cannot depend on d and, in fact, should be given by the dispersion
relation for a single heterointerface between GaN and AlN.
Let us consider in more detail the Hamiltonian for the single-heterointerface
structure. For the case when (ω)⊥,1 = (ω),1 and (ω)⊥,2 = (ω),2 , q 2 =
κ12 = κ22 and the Hamiltonian reduces to
4πe2 h̄ L 2 1 −q|z| iq·ρρ †
HIF = e e (aq + a−q ),
q ∂ 1 (ω)/∂ω + ∂ 2 (ω)/∂ω q
(7.100)
d2 d1
z
z=0
Figure 7.2. Würtzite superlattice considered by Gleize et al. (1999). From Gleize
et al. (1999). American Physical Society, with permission.
7.3 Optical modes in dimensionally confined structures 77
which can be recognized as the Hamiltonian for the IF optical phonon modes in a
zincblende single-heterointerface system.
In this subsection the interface optical phonon modes in a single-heterointerface
würtzite structure have been normalized to construct the Fröhlich-like electron–
Figure 7.3. Dispersion of interface and quasi-confined modes for an infinite and
unstrained AlN(5 nm)/GaN(5 nm) superlattice in the transverse optical (TO)
frequency range. The shaded areas depict the bands for all values of Q z ,
−π/d < Q z < π/d, lying in the first Brillouin zone of the superlattice: - — -,
Q z = 0; ——, Q z = π/d. The S ( j) and the AS ( j) are the symmetric and
antisymmetric modes with respect to the middle plane of any layer. j = 1 for GaN
and j = 2 for AlN. The quasi-confined modes are identified by their order m, an
integer following a comma. From Gleize et al. (1999), American Physical Society,
with permission.
78 7 Continuum models for phonons
Figure 7.4. Dispersion of interface and quasi-confined modes for an infinite and
unstrained AlN(5 nm)/GaN(5 nm) superlattice in the longitudinal optical (LO)
frequency range. The shaded areas depict the bands for all values of Q z ,
−π/d < Q z < π/d, lying in the first Brillouin zone of the superlattice: - — -,
Q z = 0; ——, Q z = π/d. The S ( j) and the AS ( j) are the symmetric and
antisymmetric modes with respect to the middle plane of any layer. j = 1 for GaN
and j = 2 for AlN. The quasi-confined modes are identified by their order m, an
integer following a comma. From Gleize et al. (1999), American Physical Society,
with permission.
7.3 Optical modes in dimensionally confined structures 79
with
i (q, z) = ci− e−qz + ci+ e+qz ≡ ci− φi− + ci+ φi+ , (7.102)
where the z-axis is taken to be normal to the heterointerfaces and where, as usual,
ρ = (x, y) and q denote the position and wavevector in two dimensions. ci− and ci+
are the relative amplitudes of the exponentially decaying and growing potentials,
respectively, in layer i; as will become clear, these relative amplitudes are related
through a transfer matrix. Figure 7.5 depicts a generic potential i (z) for regions
R0 , R1 , . . . , Rn .
According to the electrostatic boundary conditions the electrostatic poten-
tial i (q, z) and the normal component of the electric displacement, i Ei =
−i ∂i (q, z)/∂z, must be continuous at each heterointerface; thus, at the hetero-
interface located at z i ,
∂i (q, z i ) ∂i−1 (q, z i )
i (q, z i ) = i−1 (q, z i ) and i = i−1
∂z ∂z
(7.103)
so that
Then with the matrix C0 for region R0 , the column vector Ci and therefore
the electrostatic potential i (q, z i ) can be determined in any region through the
sequence
Qi (z i ) = Mi (z i )−1 Mi−1 (z i ).
Moreover, in each region (ω) and (∞) are related through the generalized
Lyddane–Sachs–Teller relations, subsection 2.3.3 and Section 7.1. For interface
optical phonons, the potentials must decrease exponentially as z → ±∞ so that,
for an n-region heterostructure, cn+ = 0 and c0− = 0. Thus, the dispersion relation
for this interface mode is obtained by setting the (2, 2) component of the transfer
matrix equal to zero; that is,
dispersion relation goes as (ω2 )2n . Thus, for such an n-interface structure with only
binary layers, there are 2n interface optical phonons. By extending this argument,
it follows that such a heterostructure with alternating layers of binary and ternary
semiconductors has 3n interface optical phonon modes, since each ternary layer has
two binary-like optical phonon modes.
The normalization condition for these modes is a straightforward generalization
of the normalization condition for optical phonon modes in simple heterostructures,
namely,
h̄
L2 dz |µi ni ui (q, z)|2 = . (7.108)
i Ri 2ω
∂i (q, z)
−χi (ω) iqi (q, z)q̂ + ẑ
∂z
= . (7.110)
ni ei∗ [1 + αi µi ei∗−2 (ω0i
2 − ω2 )]
Generalizing our previous expression for Pi (q, z) to the corresponding result for
layer i,
/ 0
∗ ei∗ 4π
= ni ei + n α
i i E i (q, z) + Pi (q, z) . (7.112)
µi (ω0i
2 − ω2 ) 3
82 7 Continuum models for phonons
Thus
Pi (q, z) = χi (ω)Ei (q, z)
ei∗
ni ei∗ + ni αi
µi (ω0i
2 − ω2 )
= / 0 Ei (q, z), (7.113)
ei∗
1− 3π
4
ni ei∗ + ni αi
µi (ω0i
2 − ω2 )
/ 0−2
ei∗ ei∗
2 2
4
= ni 2 −1 + π nn + αi
(ω0i − ω2 )2 3 µi (ω0i
2 − ω2 )
/ 0−2
1 ω2pi 3 ω pi
1 2
= −(ω0i − ω ) +
2 2
, (7.114)
4π (1 − 4π 3 nn αn )
2 1 − 4π 3 nn α n
where as defined previously ω2pi = 4π ni ei∗ /µi is the plasma frequency squared.
2
so that
4πe2 h̄ L −2 1 −iq·ρρ −q|z| †
HIF = e e (a−q + aq ),
q ∂0 (ω)/∂ω + ∂1 (ω)/∂ω q
(7.121)
which is identical to the result obtained previously.
A second illustrative example is given by the case of a layer of one material
situated in the region from z = −d/2 to z = +d/2 and bounded by two semi-infinite
regions of another material; for example, we might consider a GaAs quantum well
of thickness d embedded in AlAs barriers; the center of the quantum well is at
z = 0. For this case, the phonon potential must decrease exponentially for z → ±∞
and the phonon potential in the quantum well must be a combination of increasing
and decreasing exponentials. Consider the case where the phonon potential in the
quantum well is even. Let the dielectric constant in the quantum well be 1 (ω) and
that of the barriers be 0 (ω). Since the barriers are taken to be the same material,
2 (ω) = 0 (ω). Then, it is clear that 0 (q, z) = e+q(z+d/2) for z ≤ −d/2,
84 7 Continuum models for phonons
1 (q, z) = (cosh qz)/(cosh qd/2) for |z| < d/2, and 2 (q, z) = e−q(z−d/2) for
z ≥ d/2 define an admissible envelope for the phonon potential. It follows that
−d/2
∂0 (q, z) 2
dz q |0 (q, z)| +
2 2 = dz 2q 2 e2qz eqd = q,
∂z
R0 −∞
∂1 (q, z) 2
dz q |1 (q, z)| +
2 2
R1 ∂z
+d/2
cosh2 qz + sinh2 qz
= dz q 2 = 2q tanh qd/2,
−d/2 cosh2 qd/2
∞
∂2 (q, z) 2
dz q |2 (q, z)| +
2 2 =
dz 2q 2 e−2qz eqd = q,
R2 ∂z d/2
(7.122)
and, accordingly,
4π h̄ 1/2 1 ∂0 (ω) 1 ∂1 (ω) qd −1/2
A= 2q + 2q tanh (7.123)
2ωL 2 2ω ∂ω 2ω ∂ω 2
so that for the symmetric case
4π e2 h̄ L −2 1 −iq·ρρ
HIF,S = e †
f S (q, z)(a−q + aq ),
∂0 (ω) ∂1 (ω) 2q
q + tanh qd/2
∂ω ∂ω
(7.124)
where f S (q, z) = i (q, z). The dispersion relation for this optical phonon mode is
given from the requirement that the normal components of the electric displacement
field be continuous at the heterointerfaces. At z = −d/2, 0 (ω)E 0,z z=−d/2 =
1 (ω)E 1,z z=−d/2 .
From this condition it follows immediately that the frequencies of the IF optical
phonons must satisfy 0 (ω) + 1 (ω) tanh qd/2 = 0. This same dispersion relation is
obtained from the continuity of the normal component of the electric displacement
field at z = d/2. Recall that in a bulk semiconductor the optical phonon frequencies
must satisfy (ω) = 0; indeed, since this is the condition necessary for the
propagation of any longitudinal electromagnetic disturbance, it is expected that
the frequencies of longitudinal optical phonons should be given by this dispersion
relation.
In the case of a two-material, double-heterointerface structure, the IF longitudinal
optical phonon frequencies depend on both 0 (ω) and 1 (ω). The expression for
HIF,S is identical to that of Kim and Stroscio (1990) and can be rewritten to be in
the form given by Mori and Ando (1989). The mode described by this Hamiltonian
is the symmetric IF optical phonon for the quantum well being considered. As a
7.3 Optical modes in dimensionally confined structures 85
where f A (q, z) = i (q, z). As before, the dispersion relation for this mode follows
from the requirement that the normal component of the electric displacement field
be continuous at the heterointerface. In this case, 0 (ω) + 1 (ω) coth qd/2 = 0.
This result reproduces the Hamiltonian derived by Kim and Stroscio (1990) for the
antisymmetric IF optical phonon of the quantum-well system in question.
The transfer-matrix approach of Yu et al. (1997) may be used to gain insights
into the nature of phonons in superlattices. Indeed, application of the transfer
matrix method to a multiple-barrier AlAs/GaAs structure, Figure 7.6(a), leads to
the dispersion relations such as those depicted in Figure 7.7 for various AlAs/GaAs
heterostructures.
As will become evident in Chapter 10, the five-interface heterostructure of Figure
7.6(b) is of importance in narrow-well semiconductor lasers. The transfer-matrix
method of Yu et al. (1997) may be applied to determine the IF phonon dispersion
relations and the associated IF phonon potentials. For the case where the two
barriers in Figure 7.6(b) are Al0.6 Ga0.4 As and the shallow barrier to the far left is
Al0.25 Ga0.75 As, the dispersion relations are as in Figure 7.8 and the five AlAs-like
interface modes are as shown in Figure 7.9. As discussed previously, there are in
total 15 IF modes in such a five-interface binary–ternary heterostructure. Indeed, as
Figure 7.7. Longitudinal optical IF phonon dispersion relations for (a) one-barrier,
(b) two-barrier, (c) three-barrier, (d) four-barrier structures of the type shown in
Figure 7.6(a). From Yu et al. (1997), American Institute of Physics, with permission.
Figure 7.9. The phonon potentials for the five AlAs-like interface phonon modes of
Figure 7.8. The line codes used in Figure 7.8 are employed here also to indicate
which phonon potentials correspond to which. From Yu et al. (1997), American
Institute of Physics, with permission.
The antisymmetric modes a, b, c and the dispersion relation are obtained from these
results by substituting coth qd1 /2 for tanh qd1 /2.
The dispersion relations determined by Kim et al. (1992) are displayed in
Figures 7.11 and 7.12 for the case of a 60-ångstrom-wide GaAs quantum well
with 60-ångstrom-wide AlAs barriers. There are four symmetric (S) and four
antisymmetric (A) IF optical phonon modes for this heterostructure. The AlAs-like
modes are denoted by the subscript 2 and the GaAs-like modes by the subscript 1.
Figure 7.11. Dispersion relation for the four symmetric (S) LO phonon interface
modes for the heterostructure shown in Figure 7.10 for the case of a
60-ångstrom-wide GaAs quantum well with 60-ångstrom-wide AlAs barriers. From
Kim et al. (1992), American Institute of Physics, with permission.
90 7 Continuum models for phonons
The ± subscripts are used to distinguish the two different roots of the dispersion
relation for each of the modes.
As will become clear in Chapter 10 the interface LO phonon modes contribute
significantly to the valley current in such a double-barrier quantum-well structure.
There it will be explained how the mode labeled by ω S2− in Figure 7.11 makes
a major contribution to the valley current in certain double-barrier quantum-well
structures, through phonon-assisted tunneling of carriers into the quantum well.
Figure 7.12. Dispersion for the four antisymmetric (A) LO phonon interface modes
for the heterostructure shown in Figure 7.10 for the case of a 60-ångstrom-wide
GaAs quantum well with 60-ångstrom-wide AlAs barriers. From Kim et al. (1992),
American Institute of Physics, with permission.
7.4 Comparison of continuum and microscopic models for phonons 91
(Bhatt et al., 1995). Molinari et al. (1993) applied a microscopic model to calculate
the atomic displacement amplitudes and the Fröhlich potentials for quantum wells
and quantum wires. These calculations show that none of the macroscopic models –
including the so-called ‘slab’ model – gives a completely accurate representation of
the microscopic situation. However, as discussed in Appendix C the intersubband
and intrasubband scattering rates computed with the phonon modes of all these
models are in good agreement as long as the macroscopic model selected is
based on a complete set of orthogonal modes. Further discussions concerning the
comparisons of these macroscopic models are given in Appendix C. In this section,
the intersubband and intrasubband scattering rates calculated with slab modes will
be compared with the results of the microscopic models (Bhatt et al., 1993a). Figure
7.13 presents a comparison of the scattering rates for intersubband and intrasubband
electron–phonon scattering for a GaAs quantum well embedded in AlAs barriers.
For well widths in the 2 to 10 nanometer range and for a temperature of 300 K,
the scattering rates calculated from the microscopic model (Bhatt et al., 1993a) and
from the slab modes are in excellent agreement for an electron energy of 50 meV.
transitions for three different types of phonon model: macroscopic models with slab
modes (solid lines), the ab initio model (broken and dotted lines) of Molinari et al.
(1992), and a microscopic model (Bhatt et al., 1993a) with empirical force constants
(broken lines). These results indicated that the slab model and the simplified
microscopic model provide good approximations to the scattering rates predicted
by the fully microscopic model.
vary non-monotonically with the increase in supercell size. The material quality in
these structures, however, is not good enough yet for these effects to be observed
experimentally.
Resonant Raman studies on (AlGa)N/GaN quantum wells have allowed the
observation of the A1 (LO) phonons in the quantum wells (Behr et al., 1997; Gleize
et al., 2000), but the observation of a series of confined phonons similar to those
observed in the zincblende structures has yet to be made in the würtzite nitride
system.
The first really significant observation of the impact of heterostructures on light
scattering was the observation of doublets in the acoustic phonons in superlattices
by Colvard et al. (1980). A number of other studies on different materials have since
been done. These have been reviewed extensively by Jusserand and Cardona (1991).
As yet no similar observations have been reported for the würtzite nitride system,
although recently Göppert et al. (1998) have reported that confined optical and
folded acoustic phonons have been observed in the würtzite CdSe/CdS superlattices.
As a final example of where Raman measurements provide insights into the prop-
erties of confined optical phonon modes in polar semiconductor heterostructures,
Figure 7.16. Raman spectra for a 20-period GaN(9 nm)/AlN(8.5 nm) würtzite
superlattice (lower line) and a 40-period GaN(3 nm)/AlN(3 nm) würtzite superlattice
(upper line). The incident laser wavelength was 244 nm. From Dutta et al. (2000), to
be published.
96 7 Continuum models for phonons
Figure 7.17 (Fasol et al., 1988) illustrates that the confined phonon energies and
wavevectors for superlattices and quantum wells fall on the dispersion curve (solid
line) measured by neutron scattering at 10 K for bulk GaAs (Richter and Strauch,
1987). The neutron scattering data are in agreement with the earlier results of Waugh
and Dolling (1963).
The Raman data in Figure 7.17 are from works as follows: circles, Klein (1986);
diagonal crosses, Worlock (1985); squares, Castro and Cardona (1987); diamonds,
Colvard et al. (1980); and upright crosses, Sood et al. (1985), Jusserand and Paguet
(1986), and Sood et al. (1986).
The agreement between the Raman and neutron scattering measurements is
extremely enlightening. These results tell us – at least for the GaAs/AlAs system –
that phonon confinement effects serve to restrict the phase space of the phonons but
not to alter substantially the basic energy–wavevector relationship for the phonons.
This may be understood by considering the dominant role of the nearest-neighbor
coupling of adjacent ions in a polar semiconductor. Indeed, the presence of a hetero-
interface will not influence the local LO phonon frequency in the region between two
ions located about two or more monolayers away from the heterointerface. Thus, the
basic energy–wavevector relationship – the dispersion relationship – is expected to
be approximately that of the bulk phonons. The phase space, however, is altered
dramatically by the effects of dimensional confinement: only wavevectors corre-
sponding to multiples of half-wavelengths, as discussed previously, are allowed.
where
but the normal components of the stress tensor – the traction force – must vanish;
with the expressions for stress given in Section 7.2 it follows that at z = ±a/2
(Bannov et al., 1994a, b, 1995),
∂u x ∂u z
Tx,z = µ + = 0,
∂z ∂x
∂u y ∂u z
Ty,z =µ + = 0, (7.132)
∂z ∂y
∂u z
Tz,z = λ + 2µ = 0.
∂z
Taking the displacement eigenmodes to be of the form
2 ∞
1
u(r, t) = d 2 q eiq ·r un (q , z), (7.133)
2π n −∞
where
2
2 d d
t dz 2 − cl qx (cl2 − ct2 )iqx
2 2
c 0
dz
d2
D=
0 ct2 − ct2 qx2 0 ,
(7.135)
dz 2
2
d d
(cl2 − ct2 )iqx 0 cl 2 − ct qx
2 2 2
dz dz
with boundary conditions corresponding to zero traction force at z = ±a/2
du x du y du z c2 − 2c2
= −iqx u z , = 0, and = −iqx l 2 t u x .
dz dz dz cl
(7.136)
From classical acoustics (Auld, 1973) it is known that the problem at hand admits
to three types of solution: shear waves, dilatational waves, and flexural waves. First,
7.6 Continuum model for acoustic modes in dimensionally confined structures 99
consider the shear waves. For these modes, the only component of the displacement
is parallel to the surfaces z = ±a/2; taking this non-zero component to be in the
y-direction, we have un (q , z) = (0, u y , 0) with
!
cos qz,n z for n = 0, 2, 4, . . .
uy = (7.138)
sin qz,n z for n = 1, 3, 5, . . .
and qz,n = nπ/a. These transverse modes are designated as rotational modes. The
frequency–wavevector relation for these shear waves is
#
ωn = ct qz,n
2 + q 2.
x (7.139)
Clearly, these shear modes have wavelengths such that an integral number n of
half wavelengths fits into the confinement region of length a. In nanoscale crystalline
layers the number of half-wavelengths n is limited to the number of unit cells in the
thickness a. Thus, for an elemental semiconductor layer with Nm monolayers in the
thickness a, n takes on integer values from 0 to Nm .
The second class of solutions is associated with so-called dilatational modes.
These dilatational modes are irrotational modes and they are associated with
compressional distortions of the medium. The compressional character of these
modes leads to local changes in the volume of the medium. They have two non-zero
components: un (q , z) = (u x , 0, u z ) with
$ qa ql a %
t
u x = iqx (qx2 − qt2 ) sin cos ql z + 2ql qt sin cos qt z ,
2 2
2 q a q a 3 (7.140)
t l
u z = ql −(qx2 − qt2 ) sin sin ql z + 2qx2 sin sin qt z ,
2 2
where ql and qt are solutions of
For each value of qx this pair of equations has either pure imaginary or real
solutions, denoted by ql,n (qx ) and qt,n (qx ); here, the label n is used to denote the
different branches of the solutions ql,n (qx ) and qt,n (qx ). The dilatational modes have
frequencies ωn satisfying
# #
ωn = cl ql,n
2 + q2 = c
x t qt,n + q x .
2 2 (7.142)
Numerical solutions of these dispersion relations were given by Bannov et al. (1995)
for a 100-ångstrom-wide GaAs slab under the assumption that cl = 5.7×105 cm s−1
and ct = 3.35 × 105 cm s−1 .
100 7 Continuum models for phonons
Finally, the third class of solutions is referred to as the flexural modes. These
flexural modes are of the form un (q , z) = (u x , 0, u z ) with
2 qt a ql a 3
u x = iqx (qx2 − qt2 ) cos sin ql z + 2ql qt cos sin qt z ,
2 2
2 q a q a 3 (7.143)
t l
u z = ql (qx2 − qt2 ) cos cos ql z − 2qx2 cos cos qt z ,
2 2
ql and qt being determined as solutions of the pair of equations
Just as for the dilatational modes, this pair of equations for the flexural modes admits
solutions of the form ql,n (qx ) and qt,n (qx ), where n labels the different branches of
the solutions. These modes are normalized, according to the procedures of Section
5.1, in terms of wn (q , z) instead of un (q , z) since, as mentioned above, the
√
considerations of Appendix A make it clear that it is convenient to use wn = ρun :
h̄
u(r) = †
(an,q + an,−q )wn (q , z) eiq ·r . (7.145)
q ,n 2L 2 ρωn (q )
This last result is, of course, consistent with the normalization condition of subsec-
tion 7.3.1,
√ √ h̄
L 2 dz{ µnu(q, z)}∗ · { µnu(q, z)} = ; (7.146)
2ω(q)
force, must be continuous at z = ±a/2. The Young’s modulus, E α , and the Poisson
ratio, να , for medium α may be expressed in terms of the Lamé constants as
µα (3λα + 2µα ) λα
Eα = and να = (7.147)
λα + µα 2(λα + µα )
where α = 1 for the slab and α = 2 for the embedding materials. Then, the inverse
relations are
E α να Eα
λα = and µα = , (7.148)
(1 − 2να )(1 + να ) 2(1 + να )
and it follows that
Eα λα
λ α + µα = and να = . (7.149)
2(1 − 2να )(1 + να ) 2(λα + µα )
Writing the displacement field u as the sum of its longitudinal part, ul , satisfying
∇ × ul = 0, and its transverse part, ut , satisfying ∇ · ut = 0, it follows from the
wave equation (7.130) of subsection 7.6.1 that
∂2 ∂2
ul − clα
2
ul = 0 and ut − ctα
2
ut = 0, (7.150)
∂t 2 ∂t 2
where
E α (1 − να ) Eα
2
clα = and 2
ctα = . (7.151)
ρα (1 + να )(1 − 2να ) 2ρα (1 + να )
Following Wendler and Grigoryan (1988), the media are assumed to be isotropic
and, without loss of generality, we may consider acoustic modes propagating in
the x-direction with wavevector q . Wendler and Grigoryan (1988) classified the
acoustic modes for such an embedded quantum well as symmetric shear vertical
waves, antisymmetric shear vertical waves, symmetric shear horizontal waves, and
antisymmetric shear horizontal waves. Defining u(z) through the relationship
u(x, y, z) = u(z) · exp i(q x − ωt) , (7.152)
The localized modes for the embedded slab under consideration must satisfy the
boundary conditions
u(z)|z=±∞ = 0. (7.157)
Then from the wave equations for the displacements, the symmetric shear vertical
(SSV) modes must have the form
A S exp (−ηl2 z) + B2S exp (−ηt2 z) z > a/2,
2
u 1S (z) = A1S cosh ηl1 z + B1S cosh ηt1 z a/2 > z > −a/2,
A S exp η z + B S exp η z z < −a/2,
2 l2 2 t2
ηl2 S q S
i A exp (−ηl2 z) + B exp (−ηt2 z) z > a/2,
q 2 ηt2 2
ηl1 S q S
u 3 (z) =
S
i A sinh ηl1 z − B sinh ηt1 z a/2 > z > −a/2,
q 1 ηt1 1
ηl2 S q S
i A exp (ηl2 z) − B exp (ηt2 z) z < −a/2.
q 2 ηt2 2
(7.158)
and the antisymmetric shear vertical (ASV) modes have the form
A A exp (−ηl2 z) + B2A exp (−ηt2 z) z > a/2,
2
u 1 (z) =
A
A1 sinh ηl1 z + B1 sinh ηt1 z
A A
a/2 > z > −a/2,
−A A exp η z − B A exp η z z < −a/2,
2 l2 2 t2
ηl2 A q A
i A exp (−η z) + B exp (−η z) z > a/2,
q 2
l2
ηt2 2
t2
ηl1 A q A
u 3 (z) =
A
i − A cosh ηl1 z − B cosh ηt1 z a/2 > z > −a/2,
q 1 ηt1 1
ηl2 A q A
i A exp ηl2 z − B exp ηt2 z z < −a/2.
q 2 ηt2 2
(7.159)
The functions ηlα and ηtα are defined by
where α = 1 for the slab and α = 2 for the embedding materials. The conditions
ηlα = 0 and ηtα = 0 are again recognized as the bulk dispersion relations for
7.6 Continuum model for acoustic modes in dimensionally confined structures 103
medium α for the longitudinal and transverse acoustic modes, respectively. From
the definitions of the stress, Ti j , and strain, Si j , of Section 7.2, it follows that for
medium α
%
(α) Eα (α) να $ (α) (α) (α)
Ti j = S + S + S22 + S33 δi j . (7.161)
1 + να i j 1 − 2να 11
From the expressions (7.158) and (7.159) for u 1S (z), u 3S (z), u 1A (z), and u 3A (z) it then
follows that
(α) (α)
T13 = 2ρa ctα
2
S13 ,
(α)
T23 = 0, (7.162)
(α) (α) 2 (α)
T33 = ρa (clα
2
− ctα
2
)S11 + ρa clα S33 .
(α)
Requiring that Ti3 and u be continuous at z ± a/2 yields four equations. The
determinant of these equations, det dmn , then yields the dispersion relations for the
SSV modes. In dmn , m and n take on the the values 1, 2, 3, and 4, and the dmn are
given by
ηl2 a ηt2 a
d11 = exp − , d12 = exp − ,
2 2
ηl1 a ηt1 a
d13 = − cosh , d14 = − cosh ,
2 2
ηl2 a q2 ηt2 a
d21 = −ηl2 exp − , d22 = − exp − ,
2 ηt2 2
ηl1 a q2 ηt1 a
d23 = −ηl1 sinh , d24 = − sinh ,
2 ηt1 2
ηl2 a η2
2 t2
+ q2 ηt2 a
d31 = −2ρ2 ct2 ηl2 exp −
2
, d32 = −ρ2 ct2 exp − ,
2 ηt2 2
ηl1 a η2
2 t1
+ q2 ηl1 a
d33 = −2ρ1 ct1
2
ηl1 sinh , d34 = −ρ1 ct1 sinh,
2 ηt1 2
$ %
ηl2 a ηt2 a
d41 = ρ2 ct2
2
ηt2
2
+ q2 exp − , d42 = 2ρ2 ct2
2 2
q exp − ,
2 2
$ %
ηl1 a ηt1 a
d43 = −ρ1 ct1 ηt1 + q cosh
2 2 2
, d44 = −2ρ1 ct1 q cosh
2 2
.
2 2
(7.163)
For the ASV modes, the corresponding elements dmn are given in terms of these
results by making the replacements cosh sinh, d14 → −d14 , d24 → −d24 , d31 →
−d31 , d32 → −d32 , d33 → −d33 , and d44 → −d44 . As in Wendler and Grigoryan
104 7 Continuum models for phonons
(1988), A1S,A , A2S,A , B1S,A , and B2S,A are then given in terms of the new elements
dmn by
−d14 d12 d13
1
A1S,A = −d24 d22 d23 ,
det dmn
−d34 d32 d33
d11 d12 −d14
1
A2S,A = d21 d22 −d24 , (7.164)
det dmn
d31 d32 −d34
d11 −d14 d13
1
B1S,A = d21 −d24 d23 ,
det dmn
d31 −d d
34 33
and
B2S,A = 1, (7.165)
By procedures analogous to those for the SSV modes, it follows that the
dispersion relation for the SSH modes is given by
2 η
ρ2 ct2 t2 θt1 a
− tan = 0. (7.169)
ρ1 ct1
2 θ
t1 2
7.6 Continuum model for acoustic modes in dimensionally confined structures 105
Here c is the phase velocity and γ is the wavevector in the z-direction. Morse
considered the wave equations
∂2 ∂2 ∂2 ∂2
+ 2 + ql φ = 0,
2
+ 2 + ql ψi = 0,
2
(7.173)
∂x2 ∂y ∂x2 ∂y
where
where
ρ is the density of the elastic medium, and the longitudinal and transverse sound
speeds are given by
cl = (λ + 2µ)/ρ and ct = µ/ρ. (7.179)
The modes associated with this case are known as the ‘thickness modes’, as
designated by Morse, who showed that h = h 1 = h 2 leads to an adequate
description of the experimentally determined modes when d ≥ 2a (Morse, 1948).
Using the expressions (7.177) for u, v, and w to evaluate Tx x = Tyx = Tzx = 0 at
x = ±a, it follows that
2h sin q1 a h sin q2 a q2 sin q2 a
A
−(γ 2 + h 2 − q 2 ) cos q1 a 2q1 q2 cos q2 a 0 B
2
2(h 2 + γ 2 ) sin q1 a (γ 2 + h2 − q22 ) sin q2 a 0 C
= 0. (7.180)
The dispersion relation for q2 = 0 is given by the expression resulting from the
condition that the determinant of the coefficients vanishes, that is,
tan q2 a 4q1 q2 (h 2 + γ 2 )
=− , (7.181)
tan q1 a (h 2 + γ 2 − q22 )2
c 2 $ a %2 2 c 2 $ a %2 ω 2
χ 2 = s2 = γ = ,
ct π ct π ct
2 $ %
ah a 2 2
ψ =s +
2 2
= (γ + h 2 ), (7.183)
π π
$ π %2 $π % 2
r̃
γ2 = ψ 2 − h2 = ψ2 −
a a 2
with r̃ = a/d.
Solving (7.180) for B and C in terms of A, it follows that
where
sin q1 a 2(h 2 + γ 2 )
B = αA = − A,
sin q2 a (γ 2 + h 2 − q22 )
(7.185)
q2 h sin q1 a 2q2 h
C =− B= A = β A.
h2 + γ 2 sin q2 a (γ 2 + h 2 − q22 )
a d
1 h̄
dx dy (u ∗ u + v ∗ v + w ∗ w) = , (7.186)
4ad −a −d 2Mωγ
where ωγ is the angular frequency of the mode with wavevector γ . Performing the
indicated integrations, it follows that
7.6 Continuum model for acoustic modes in dimensionally confined structures 109
! 2 3
A2
f 1 (h, d) f 2 (q1 , a) + 2αg1 (q1 , q2 , a) + α 2 f 2 (q2 , a)
4ad
/ 0
h2 2βh
− f 1 (h, d) 2 f 1 (q1 , a) + g2 (q1 , q2 , a) + β 2 f 1 (q2 , a)
q1 q1
2
γ 2
+ f 1 (h, d) 2 f 1 (q1 , a) − (q2 α + hβ)g2 (q1 , q2 , a)
q1 q 1
q1 α + hβ 2
+ f 1 (q2 , a)
γ2
/ 0"
h2 2βh
+ 2d f 1 (q1 , a) + g2 (q1 , q2 , a) + β f 1 (q2 , a)
2
q12 q1
h̄
= , (7.187)
2Mωγ
with
sin 2hd
f 1 (h, d) = d 1 + , f 2 (h, d) = 2d − f 1 (h, d),
2hd
sin(q1 − q2 )a sin(q1 + q2 )a
g1 (q1 , q2 , a) = − , (7.188)
q1 − q2 q1 + q2
sin(q1 − q2 )a sin(q1 + q2 )a
g2 (q1 , q2 , a) = + .
q1 − q2 q1 + q2
It is convenient to define a new normalization constant Bγ , through
2h̄
A2 = ≡ A2γ . (7.189)
Mωγ Bγ
The principal propagation mode has no nodal surfaces parallel to the length of
the quantum wire; this corresponds to the case n = 0. Morse found close agreement
between theory and experiment for a/d = 1/8 and as expected less agreement
for a/d = 1/2. In addition to the ‘thickness modes’ another set of modes was
observed experimentally by Morse (1948, 1950). These modes are known as ‘width
modes’ and are determined by a procedure used to analyze the ‘thickness modes’.
Specifically, Morse took q1 = q2 = q and obtained a set of equations similar to
those for u 1 , v1 , and w1 but with x and y interchanged. By imposing the boundary
conditions at y = ±d, the ‘width modes’ were found to have a dispersion relation
110 7 Continuum models for phonons
identical in form to that for the ‘thickness modes.’ The dispersion curves for selected
acoustic modes are shown in Figure 7.19 for a 28.3 Å × 56.6 Å GaAs quantum wire
and in Figure 7.20 for a 50 Å × 200 Å GaAs quantum wire.
For carriers at the non-degenerate point in band α, E α (k), the deformation-
potential interaction Hamiltonian Hdef α is given in terms of the displacement operator
û(r) by
α
Hdef = E 1α ∇ · û(r). (7.191)
At such a symmetry point, only the irrotational – that is, the longitudinal – compo-
α . Accordingly, only the the potential φ contributes
nents of u(r) contribute to Hdef
α . Since there are multiple modes for a given value of n, another index, m,
to Hdef
is needed to describe the phonon spectrum at each value of γ . For the case of a
quantum wire, the quantization of the acoustic phonons may be performed by taking
1
û(r) = √ [u(γ , x, y)aγ + c.c.] eiγ z , (7.192)
L n,m,γ
where the components of u(γ , x, y) = (u, v, w) were normalized previously over
the area 4ad. The deformation potential is then given by
α E 1α ∂u ∂v
Hdef = √ [an,m (γ ) + an,m (−γ )] + + iγ w eiγ z ,
L n,m,γ ∂x ∂y
(7.193)
Figure 7.19. The six lowest-order (m = 1, 2, . . . , 6) width modes (solid lines) and
thickness modes (broken lines) for a 28.3 Å × 56.6 Å GaAs quantum wire. From Yu
et al. (1994), American Physical Society, with permission.
7.6 Continuum model for acoustic modes in dimensionally confined structures 111
and, upon applying the Fermi golden rule, these combinations lead to conditions
enforcing the conservation of energy.
The Hamiltonian is independent of time. In Chapters 8 and 9, such time-
independent carrier–phonon Hamiltonians will be used in applying the Fermi golden
rule to calculate carrier–phonon scattering rates. The carrier–phonon interaction also
has a time dependence of the form eiωγ t , where ωγ is the phonon frequency. As will
become obvious in Chapters 8 and 9, such time-dependent factors are combined
with the time-independent factors of carrier wavefunctions. Since
∂u
= A(q1 cos q1 x + αq2 cos q2 x) cos hy,
∂x
∂v h2
=A cos q1 x + β cos q2 x cos hy, (7.194)
∂y q1
/ 0
γ2
−iγ w = A − cos q1 x + (q2 α + hβ) cos q2 x cos hy,
q1
it follows that
∂u ∂v h2 γ2
+ + iγ w = A q1 + + cos q1 x cos hy
∂x ∂y q1 q1
1 $ 2 %
=A q1 + h 2 + γ 2 cos q1 x cos hy
q1
ωγ
= A 2 cos q1 x cos hy (7.195)
cl q1
and, accordingly,
α Eα ωγ
Hdef = √1 A 2 cos q1 x cos hy[an,m (γ ) + an,m (−γ )] eiγ z .
L n,m,γ cl q1
(7.196)
and also because it resembles the microtubuline structure found in many parts
of the human body. As discussed previously in this section and in Section 7.3,
the elastic continuum model provides an approximate description of the acoustic
phonon modes in such dimensionally confined nanostructures. The force equations
for a cylindrical elastic medium may be written as (Auld, 1973; Sirenko et al., 1995)
∂ 2 u zϕ ∂ Tr z 1 ∂ Tϕz ∂ Tzz Tr z
ρ = + + + ,
∂t 2 ∂r r ∂y ∂z r
where the axis of the cylinder is oriented along the z-direction, ϕ is the azimuthal
angle, and r is the radial coordinate of the cylindrical structure. As before, the stress
tensor T is related to the strain tensor S through the Hooke’s law relationship
in this stress–strain relation, λ and µ are the Lamé constants. Alternatively, these
force equations are frequently written in the form
Figure 7.20. The six lowest-order (m = 1, 2, . . . , 6) width modes (solid lines) and
thickness modes (broken lines) for a 50 Å × 200 Å GaAs quantum wire. From Yu
et al. (1994), American Physical Society, with permission.
7.6 Continuum model for acoustic modes in dimensionally confined structures 113
∂ ∂u k (r, t)
ρ(r)ü i (r, t) = λi jkl (r) , (7.199)
∂ xj ∂ xl
where the elastic stiffness tensor for a particular isotropic medium is expressed as
These equations are more complicated than their counterparts in rectilinear co-
ordinates, Section 7.2. Indeed, the additional complexity of the force equations
in cylindrical coordinates is a direct consequence of the fact that in curvilinear
coordinates the basis vectors are coordinate dependent.
Consider the acoustic phonon modes in a cylindrical waveguide of radius a
embedded in an elastic medium. Both of these media are taken to be isotropic. From
the normalization procedures of Section 5.1, the modes are normalized in terms
of w instead of u since the considerations of Appendix A make it clear that it is
√
convenient to use w = ρu; the displacement operator û(r) is then given by
h̄
û(r) = [wmn,q (r )amn,q + w∗mn,−q (r )amn,−q
†
] eimϕ+iqz/a .
q,mn 2Lρωmn (q)
(7.201)
The quantum number n labels modes with the same m and q in the set wmn,q (r ),
where q represents the z-component of the wavevector qz . In determining the
normalization constants for the normal modes wmn,q (r ), it is convenient to write
1
wmn,q (r) = wmn,q (r ) eimϕ+iqz/a = √ u(r ) eimϕ+iqz/a , (7.202)
πa 2 N
where u(r ) is the classical displacement given by the elastic continuum model and
the normalization constant N is then determined by the normalization condition
d 2r ρ(r)w∗n,m,q (r) · wn ,m ,q (r) = δn,m,q;n ,m ,q , (7.203)
and q ≡ aqz .
Let the density and Lamé constants of the cylindrical waveguide be ρ1 , λ1
and µ1 respectively, and those of the surrounding material be ρ2 , λ2 and µ2 . The
general solution of the classical elastic continuum equations for such a cylindrical
structure may be written (Beltzer, 1988; Stroscio et al., 1996) in terms of three scalar
potentials φ, ψ, and χ as
where êz is a unit vector along the z-direction. The second and third terms in this
last result correspond to the usual irrotational contribution to u, expressed as a sum
114 7 Continuum models for phonons
of two mutually normal vectors. The potentials φ, ψ, and χ satisfy scalar wave
equations with longitudinal and transverse sound speeds given by
clξ = (λξ + 2µξ )/ρξ and ctξ = µξ /ρξ ; (7.205)
the subscript ξ takes on the value 1 to designate the material parameters of the
cylinder and the subscript 2 to designate those of the surrounding material. Solutions
of the classical elastic continuum equations are sought with vibration frequency ω,
wavevector qz = q/a, and azimuthal quantum number m. Seeking acoustic modes
confined near the cylindrical waveguide, the scalar potentials for r < a are taken to
be
φ ibl1 Jm (kl r/a)
1
ψ = Bt1 Jm (kt r/a) eimϕ+iqz/a−iωt . (7.206)
a
χ bt1 Jm (kt r/a)
Outside the cylindrical waveguide, where r > a, the solutions are taken to be
φ ibl2 K m (κl r/a)
1
ψ = Bt2 K m (κt r/a) eimϕ+iqz/a−iωt , (7.207)
a
χ bt2 K m (κt r/a)
where kl,t and κl are defined by
2
kl,t = q 2 − ω2 a 2 /c(l,t)1
2
and κl,t
2
= ω2 a 2 /c(l,t)2
2
− q2 (7.208)
and the bl1 etc. are normalization constants, to be determined. In the expressions for
φ, ψ, and χ, it is assumed that kl,t 2 > 0 and κ 2 > 0, since confined acoustic modes
l,t
are desired. Substituting these potentials into the general expression for u it follows
that
kl r a kt r kt r
−iu r (r ) = bl1 kl Jm + Bt1 m Jm + bt1 qkt Jm ,
a r a a
a kl r kt r a kt r
−u ϕ (r ) = bl1 m Jm + Bt1 kt Jm + bt1 mq Jm ,
r a a r a
kl r kt r
−u z (r ) = bl1 q Jm − bt1 kt Jm
2
a a
(7.209)
for r < a and
a
−iu r (r ) = bl2 kl K m (κl r/a) + Bt2 m K m (κt r/a) + bt2 qκt K m (κt r/a),
r
a κl r κt r a κt r
−u ϕ (r ) = bl2 m K m + Bt1 κt K m + bt1 mq K m ,
r a a r a
(7.210)
7.6 Continuum model for acoustic modes in dimensionally confined structures 115
κl r κt r
−u z (r ) = bl2 q J K m − bt1 κt2 K m
a a
for r > a. By applying the boundary conditions of continuity of displacement and
continuity of the normal components of the stress tensor at r = a, it follows that
U1 −U2 B1
= 0, (7.211)
µ1 F1 −µ2 F2 B2
where Bξ = [blξ , Btξ , btξ ]T and where the ‘displacement’ matrices Uξ are evaluated
at r = a and are given by
Gξ myξ qYξ
−mgξ −Yξ −mqyξ ; (7.212)
−qgξ 0 ktξ yξ
(7.213)
=1 = k(l,t) , k(l,t)ξ =2 = −κ(l,t) , g1 = Jm (kl ), G 1 = kl Jm (kl ), y1 =
2
where k(l,t)ξ 2 2 2
wϕ = 0,
1 bl1 q J0 (kl r/a) − bt1 kt2 J0 (kt r/a) r < a,
−wz = √
πa 2 N r z bl2 q K 0 (κl r/a) + bt2 κt2 K 0 (κt r/a) r > a.
(7.219)
For the axisymmetric radial–axial modes, the normalization condition requires
that
2 2 2
Nr z = ρ1 b1t {q [J0 (kl ) + J12 (kl )] + kl2 [J12 (kl ) − J0 (kl )J1 (kl )]}
+ b1t kt {kt [J0 (kt ) + J12 (kt )] + q 2 [J12 (kt ) − J0 (kt )J2 (kt )]}
2 2 2 2
− 4b1l b1t qkt J0 (kl )J1 (kt )
2 2 2
+ ρ2 b2l {q [K 1 (κl ) − K 02 (κl )] + κl2 [K 0 (κl )K 2 (κl ) − K 12 (κl )]}
+ b2t κt {κt [K 12 (κt ) − K 02 (κt )] + q 2 [K 0 (κt )K 2 (κt ) − K 12 (κt )]}
2 2 2
− 4b2l b2t qκt K 0 (κt )K 1 (κt ) . (7.220)
As for the case of rectangular quantum wires, subsection 7.6.3, for carriers at
the non-degenerate point in band α, E α (k), the deformation-potential interaction
α is given in terms of the displacement operator û(r) by
Hamiltonian Hdef
7.6 Continuum model for acoustic modes in dimensionally confined structures 117
α
Hdef = E 1α ∇ · û(r). (7.221)
Again, at such a symmetry point only the irrotational – that is, the longitudinal
– components of u(r) contribute to Hdef α . Accordingly, only the the potential φ,
α
(7.204), contributes to Hdef . Indeed, from the normalized components wmn,q (r)
(7.202), û(r) is obtained readily and by using the relation ∇ 2 φ = −(ω/cl )2 φ it
follows that
ωmn (q) 2 h̄
α α
Hdef = −E 1
q,mn acl 2π LρN ωmn (q)
× mn,q (r )amn,q + ∗mn,−q (r )amn,−q
†
eimϕ+iqz/a , (7.222)
Let us consider the case of a thin cylindrical shell. For a cylindrical shell, the
boundary conditions on the inner and outer surfaces are
Pµ = Tµν n ν , (7.224)
where P represents an external pressure that would be present, for example, in the
case where the cylindrical shell is in contact with a liquid, Tµν is the stress tensor,
and n ν is the normal to the surface of the shell. In particular, P = ±P in êr and
out
n = ∓êr where êr is the unit vector in the r -direction. The subscripts ‘in’ and ‘out’
are alternatives. For a cylindrical shell of infinite length in the z-direction and of
thickness h and radius R such that h R, the boundary conditions are
Tr ϕ R∓h/2 = Tr z | R∓h/2 = 0, Trr | R∓h/2 = P in . (7.225)
out
Assuming that all quantities except Trr are nearly constant with respect to r over
the interval from R − h/2 to R + h/2, it is possible to show that
These results follow straightforwardly by integrating the right- and left-hand sides
of each force equation over the interval from R − h/2 to R + h/2, invoking the
118 7 Continuum models for phonons
boundary conditions in the radial force equation, and cancelling factors of h. From
the stress–strain relation (7.198), it follows that
1 ∂u ϕ ∂u z 1 ∂u z ∂u ϕ
Sϕϕ = + ur , Szz = , and Sϕz = + ,
r ∂ϕ ∂z r ∂ϕ ∂z
(7.233)
(7.239)
2
−1 m νq
det D = m 2 − m 2 − ν− q 2 −ν+ mq = 0.
νq −ν+ mq 2 − ν− m 2 − νq 2
(7.240)
Hence
6 − [(ν− + 1)(m 2 + q 2 ) + 1]4
+ (m 2 + ν− q 2 )(ν− m 2 + q 2 ) + (1 + ν− )(m 2 + q 2 )
− ν 2 q 2 − m 2 − ν+ m q 2 − ν− (1 − ν 2 )q 4 = 0
2 2 2
(7.241)
or
Aq 4 − Bq 2 + C = 0, (7.242)
where
A = ν− (1 + ν 2 ),
√
it follows that the longitudinal mode has frequency longitudinal ≡ II m = ν− m.
Since m ≈ 0 was obtained in the lowest order in q it is necessary to find the first
III
2
non-vanishing term. For m = 0, by making the assumption that III m=0 = αq 2
and collecting terms up to order q 4 in the dispersion equation, it follows that
−α 2 q 4 + αq 2 (1 + ν− − ν 2 )q 2 − ν− (1 − ν 2 )q 4 = 0. (7.247)
α[ν− m 4 + (1 + ν− )m 2 − m 2 ]q 4 − ν− (1 − ν 2 )q 4 = 0, (7.248)
so that α = (1 − ν 2 )/(m 4 + m 2 ) and IIIm ≈ (1 − ν 2 )/(m 2 + 1)(q 2 /m). In the
limit of large q the leading terms of the dispersion relation imply that
6 − (1 + ν− )q 2 4 + ν− q 4 2 − ν− (1 − ν 2 )q 4 = 0 (7.249)
α 3 − (1 + ν− )α 2 + ν− α = 0, (7.250)
m (q)
and III 1 − ν 2 . Analysis of the coefficients cr , cϕ , and cz (Sirenko et al.,
1996a, b) reveals that in the limit of large q the I, II, and III modes correspond to
pure longitudinal, torsional, and radial vibrations, respectively.
122 7 Continuum models for phonons
(7.252)
where the term describing the coupling between the shell and the fluid is given by
α Im (Q) K m (Q)
Wm Q = − . (7.253)
Q Im (Q) K m (Q)
Here
2
ω2 c
Q =R
2 2
qz2 − 2 ≡ q2 − 2 , (7.254)
sf sf
In the case of a thin cylindrical shell immersed in fluid, interface modes in the fluid
are localized near the cylindrical surface and correspond to the region Q 2 > 0. In
the case where ω > s f qz the acoustic disturbances are radiated from the cylindrical
shell. Indeed, the ωqz -plane is divided into two regions by the curve ω > s f qz . As
just indicated, the region defined by Q 2 > 0, or ω < s f qz , is the region where
interface modes are localized on the scale of R/Q from the cylindrical shell. When
ω > s f qz , Q 2 < 0 and the relation
π (1)
K m (−i|Q|) = i m+1 H (|Q|),
2 m
makes manifest the radiation of cylindrical waves from the cylindrical shell into the
(1)
region surrounding it, since the Hankel function of the first kind, Hm , represents
outgoing cylindrical waves.
The case of a cylindrical shell immersed in fluid was solved numerically by
Sirenko et al. (1996b) in order to model the vibrational behavior of microtubules
(MTs) immersed in water. These microtubules are of great interest in biology as
they are present in many biological structures including cytoskeleton and eukariotic
cells. In these numerical calculations, the inner and outer radii of the MTs are taken
to be 11.5 nm and 14.2 nm respectively, the length of the MTs is taken to be 8 nm,
their mass is taken to be 1.83 × 10−19 g, the Poisson ratio, ν, is taken to be 0.3, and
the Young’s modulus is taken to be 0.5±0.1 GPa. These parameters give a thin-plate
longitudinal sound speed c of 610 m s−1 and a density of 1.47 g cm−3 . The sound
7.6 Continuum model for acoustic modes in dimensionally confined structures 123
speed and density of water are taken to be 1.50 km s−1 and 1 g cm−3 . The calculated
dispersion relations, im (q), for m = 1, 2, and 3 are shown in Figures 7.21, 7.22,
and 7.23, respectively.
From Figures 7.21–7.23, it is apparent that for qz m/R the mode frequencies
of the immersed MTs tend to those of the free-standing MTs and do not depend on
m. These modes are seen to have maximum frequencies of the order of tens of GHz.
Moreover, the sound speeds of the axisymmetric acoustic modes are in the range
200–600 m s−1 .
Sirenko et al. (1996b) also considered the dynamical behavior of cytoskeletal
filaments, by using the elastic continuum model to determine the mode structure
for the vibrations of a solid cylinder. Particular attention was given to (a) the
axisymmetric torsional mode, (b) the axisymmetric radial–longitudinal mode, and
(c) the flexural mode, as depicted in Figure 7.24.
Figure 7.21. Dispersion relations im for m = 1 modes of the MT sample discussed
in the text. I, pure longitudinal; II, torsional; III, radial. The frequency is in units of
= ω R/c and q = qz R. = 1 yields ω = 7.6 GHz for the parameters given in the
text. The dotted line separating the regions of interface and radiated waves
corresponds to the condition ω = s f qz , as discussed in the text. The solid and broken
lines correspond to MTs immersed in water and free standing respectively. From
Sirenko et al. (1996a), American Physical Society, with permission.
124 7 Continuum models for phonons
and normalizing the acoustic phonon Fourier amplitude, u(q, r, ϕ, z), according to
Figure 7.22. Dispersion relations im for m = 2 modes of the MT sample discussed
in the text. I, pure longitudinal; II, torsional; III, radial. The frequency is in units
of ω R/c and q = qz R. = 1 yields ω = 7.6 GHz for the parameters given in the
text. The dotted line separating the regions of interface and radiated waves
corresponds to the condition ω = s f qz , as discussed in the text. The solid and broken
lines correspond to MTs immersed in water and free standing respectively. From
Sirenko et al. (1996a), American Physical Society, with permission.
7.6 Continuum model for acoustic modes in dimensionally confined structures 125
2π π a
3 h̄
dϕ dθ sin θ dr r 2 u(q, r, θ, ϕ) · u∗ (q, r, θ, ϕ) = .
4πa 3 0 0 0 2Mωq
(7.257)
Here, a is the radius of the quantum dot, N is the number of unit cells in the
normalization volume V , aq is the phonon annihilation operator, q is the phonon
wavevector, ωq is the angular frequency of the phonon mode, M is the mass of the
ions in the unit cell, and r, θ, φ are the usual spherical coordinates. For a quantum
dot with rectangular faces the normalization condition for the acoustic phonon mode
amplitude is given by
a/2 b/2 c/2
1 h̄
dx dy dz u(q, x, y, z) · u∗ (q, x, y, z) = .
abc −a/2 −b/2 −c/2 2Mωq
(7.258)
Figure 7.23. Dispersion relations im for m = 3 modes of the MT sample discussed
in the text. I, pure longitudinal; II, torsional; III, radial. The frequency is in units of
= ω R/c and q = qz R. = 1 yields ω = 7.6 GHz for the parameters given in the
text. The dotted line separating the regions of interface and radiated waves
corresponds to the condition ω = s f qz , as discussed in the text. The solid and broken
lines correspond to MT immersed in water and free standing, respectively. From
Sirenko et al. (1996a), American Physical Society, with permission.
126 7 Continuum models for phonons
The classical acoustic modes in an isotropic elastic medium have been analyzed
previously, and many of the most useful known results summarized, by Auld (1973).
The lowest-order pure-compressional mode is referred to frequently as the breathing
mode. The displacement field associated with this lowest order compressional mode
of a sphere of radius a is given by
ωq r
u(q, r ) = r̂γ j1 e−iωq t , (7.259)
cl
where γ is the normalization constant, r̂ is the unit vector in the radial direction, j1
is the spherical Bessel function of order unity, j1 (x) = sin x/x 2 − cos x/x, ωq is the
√
mode frequency, and the longitudinal sound speed cl is equal to (λ + 2µ)/ρ. The
frequency for a free-standing sphere is determined by the condition that the normal
component of the traction force at the surface of the sphere vanishes; that is, Trr = 0
at r = a:
d2 ωq r 2λ d ωq r
(λ + 2µ) 2 j0 + j0 = 0, (7.260)
dr cl r dr cl r =a
where j0 (x) = sin x/x is the spherical Bessel function of order zero. This last result
implies that
ωq r 4µωq a/c
tan = . (7.261)
cl 4µ − (λ + 2µ)(ωq a/c)2
The normalization condition for this lowest-order breathing mode is
ω̃l
3γ 2 h̄
dr r 2 j12 (r ) = , (7.262)
ω̃l3 0 2Mωq
with ω̃l representing the quantity ωq a/cl . This integral may be performed analyti-
cally and it follows that
1 h̄
γ =# , (7.263)
j 2 (ω̃ ) − j (ω̃ ) j (ω̃ ) 3Mωq
1 l 0 l 2 l
ωq r
ϕ τ cos θ j1
u(q, r, θ) = ϕ̂ e−iωq t , (7.264)
ct
ϕ is the unit vector in the ϕ -direction, τ is the normalization constant to
where ϕ̂
be determined from the phonon normalization condition, and the transverse sound
√
speed of the shear wave is given by ct = µ/ρ. This mode is depicted in Figure
7.25(b). The normalization condition for this mode is
τ 2 ω̃t h̄
dr r 2 j12 (r ) = , (7.265)
ω̃t 0
3 2Mω q
with ω̃t = ωq a/ct . Thus, the normalization constant τ may be evaluated in terms of
the same integral used to calculate γ ; indeed,
1 h̄
γ =# . (7.266)
j 2 (ω̃ ) − j (ω̃ ) j (ω̃ ) Mω q
1 t 0 t 2 t
Following the same procedure as for the breathing mode, it follows that the
dispersion relation for the lowest-order torsional mode is
ωq a 3ωq a/ct
tan = . (7.267)
ct 3 − (ωq a/ct )2
Krauss and Wise (1997) have recently observed the coherent acoustic phonons in
spherical quantum dots. The damping of the lowest-order acoustic phonon modes
observed by Krauss and Wise has been described in terms of the elastic continuum
model by Stroscio and Dutta (1999).
McSkimin (1944) gave approximate classical flexural thickness modes for a
structure with rectangular faces. The structure considered in this section has faces
joining each other at right angles, and the faces in the x y-, yz-, and x z-planes are
rectangles such that the width of the structure – in the x-direction – is a, the height
of the structure – in the y-direction – is b, and the length of the structure – in the
z-direction – is c. The approximate flexural thickness modes given by McSkimin are
∂u ∂v ∂w
(x, y, z) = + +
∂x ∂y ∂z
l12 n2
= A cos mx m sin l1 y + sin l1 y + sin l1 y cos nz. (7.269)
m m
The flexural acoustic modes of McSkimin are used in this section to illustrate the
quantization of modes for such a structure. Obviously, these modes do not form a
complete set. Indeed, in addition to compressional modes there are flexural modes
corresponding to the width and length modes. The normalization condition for
McSkimin’s flexural thickness modes,
1 a/2 b/2 c/2 h̄
dx dy dz uu ∗ + vv ∗ + ww ∗ = , (7.270)
abc −a/2 −b/2 −c/2 2Mω
reduces to
1
f 2 (m, a/2) f 1 (n, c/2)[ f 2 (l1 , b/2) + 2αg1 (l1 , l2 , b/2)
abc
+ 2βg1 (l1 , l3 , b/2) + α 2 f 2 (l2 , b/2) + 2αβg1 (l2 , l3 , b/2) + β 2 f 2 (l3 , b/2)]
/
l2 2αl1l2
+ f 1 (m, a/2) f 1 (n, c/2) 12 f 1 (l1 , b/2) + g2 (l1 , l2 , b/2)
m m2
2βl1 (m 2 + n 2 ) α 2l22 2αβl2 (m 2 + n 2 )
− g2 (l 1 , l 3 , b/2) + f 1 (l 2 , b/2) −
l3 m 2 m2 l3 m 2
0
2β (m + n )
2 2 2
× g2 (l2 , l3 , b/2) + f 1 (l3 , b/2)
l32 m 2
/
n2 2αn(m 2 + l22 )
+ f 1 (m, a/2) f 2 (n, c/2) f 2 (l 1 , b/2) − g1 (l1 , l2 , b/2)
m2 nm 2
2βn 2 α 2 (m 2 + l22 )
+ g1 (l 1 , l 3 , b/2) + f 2 (l2 , b/2)
m2 n2m 2 0
2αβn(m 2 + l22 ) β 2n2
− g1 (l2 , l3 , b/2) + f 2 (l3 , b/2)
nm 2 m2
h̄
= , (7.271)
2Mω
where
130 7 Continuum models for phonons
+a/2
a sin ha
f 1 (h, a/2) = dy cos hy =
2
1+ ,
−a/2 2 ha
+a/2
f 2 (h, a/2) = dy sin2 hy = a − f 1 (h, a/2),
−a/2
+b/2
g1 (li , l j , b/2) = dy sin li y sin l j y
−b/2
(7.272)
sin(li − l j )b/2 sin(li + l j )b/2
= − ,
li − l j li + l j
+b/2
g2 (li , l j , b/2) = dy cos li y cos l j y
−b/2
The deformation potential for these flexural thickness modes has an especially
simple form at a non-degenerate -point, namely,
Hdef = E 1 (x, y, z)
E1 A 2
=√ (m + l1 + n ) (cos mx sin l1 y cos nz) aq + c.c. .
2 2
N q m
(7.273)
131
132 8 Carrier–LO-phonon scattering
phonon absorption processes the initial and final states are chosen in a similar
manner. In the energy-conserving delta function in S {e,a} (k, k ), the upper sign
corresponds to phonon emission and the lower sign to phonon absorption. In this
expression,
7 the electron and phonon states represent different spaces so the ‘product’
7
|k Nq ≡ k, Nq ; that is, calculating the matrix elements of electron states involves
only integration over coordinate space and calculating matrix elements of phonon
states involves only taking matrix elements of the phonon creation and annihilation
operators as in Section 5.1. Clearly these two types of matrix element may 7 be
|k Nq ≡
calculated
7
independently of each other. For this reason, the ‘product’
k, Nq should not cause any confusion. The phonon states are as defined in Section
5.1 and the Fröhlich Hamiltonian is that of Section 5.2,
2π e2 h̄ωLO 1 1 1 iq·r
HFr = −i − aq e − aq† e−iq·r
V (∞) (0) q q
C
= aq eiq·r − aq† e−iq·r , (8.3)
q q
where
2πe2 h̄ωLO 1 1
C = −i − . (8.4)
V (∞) (0)
where the phonon matrix elements have been evaluated using the expressions of
Section 5.1; in particular,
5 7 √ 5 7
Nq − 1 aq Nq = n q and Nq + 1 aq† Nq = n q + 1. (8.7)
In (8.6) the upper signs correspond to phonon emission and the lower signs to
phonon absorption. The integral over the factor e−ik ·r+ik·r∓iq·r /V is, of course,
8.1 Fröhlich potential for LO phonons in bulk zincblende and würtzite structures 133
|C|2 1
= nq + 1
± 1
, (8.8)
4π h̄ 2 |k − k |2 2 2
where
2π 2 1 1
|C| =
2
e h̄ω − . (8.9)
V (∞) (0)
In evaluating the integral over q it has been assumed that the frequency of the
phonon mode, ωq , is independent of q and is equal to the zone-center phonon
frequency, ωq . This is a good approximation for many materials. For example,
in GaAs the LO phonon frequency varies by about 10% over the entire Brillouin
zone. Moreover, since the largest contributions to the integral occur near the zone
center – as a result of the q −1 dependence in the integrand – the approximation that
ωq = ωLO causes little error in the value of the integral in question.
Accordingly, the transition rate S {e,a} (k, k ) is given by
|C|2 1
S {e,a} (k, k ) = nq + 1
± 1
δ(E(k ) − E(k) ± h̄ω)
4π h̄ 2 |k − k |2 2 2
|C|2 1
= nq + 1
2 ± 1
2 δ(E(k ) − E(k) ± h̄ω), (8.10)
4π h̄ 2 q 2
where q 2 = |q|2 and the delta function expresses the conservation of energy, E(k )−
E(k) = ∓h̄ω, the upper sign representing phonon emission and the lower sign
representing phonon absorption. The |q|−2 dependence in S {e,a} (k, k ) is the same
as that appearing in the Coulomb interaction. Accordingly, it is anticipated that finite
134 8 Carrier–LO-phonon scattering
|C|2 q2
S {e,a} (k, k ) = nq + 1
2 ± 1
2 δ(E(k ) − E(k) ± h̄ω),
4π h̄ (q + q0 )
2 2 2 2
(8.11)
1 |C|2 V
{e,a}
= 2
n q + 12 ± 12
τ3D (k) 4π h̄
2π π
q2
× dϕ dθ sin θ dq q 2 δ(E(k ) − E(k) ± h̄ω).
0 0 (q 2 + q02 )2
(8.12)
h̄ 2
f (θ) = ∓ (2qk cos θ), (8.15)
2m
it follows by letting u = cos θ that
8.1 Fröhlich potential for LO phonons in bulk zincblende and würtzite structures 135
2π h̄ 2 2
dθ sin θ δ (q ∓ 2qk cos θ ) ± h̄ω
0 2m
1
h̄ 2 h̄ 2 2
= du δ (∓2qku) + q ± h̄ω
−1 2m 2m
m
=∓ 2 , (8.16)
h̄ qk
so that
{e,a}
|C|2 V
1 2π m qmax q3
{e,a}
= nq + 1
2 ± 1
2 dq ,
τ3D (k) 4π 2 h̄ h̄ 2 k qmin
{e,a} (q 2 + q02 )2
{e,a}
|C|2 V m qmax q3
= nq + 1
2 ± 1
2 dq . (8.17)
2π h̄ 3 k qmin
{e,a} (q 2 + q02 )2
{e} {e}
We consider the ranges of q associated with emission and absorption, {qmin , qmax }
{a} {a}
and {qmin , qmax } respectively. The range of q is determined by
2mω
q 2 − 2qk cos θ ± = 0, (8.18)
h̄
where the upper sign corresponds to emission and the lower sign to absorption. The
roots of this equation are
{e} 2mω
q = k cos θ ± k cos θ −
2 2 , (8.19)
h̄
and
{a} 2mω
q = −k cos θ ± k 2 cos2 θ + . (8.20)
h̄
Hence, by taking cos θ = ±1, the minimum and maximum values are found:
{e} {e} 2mω 2mω
{qmin , qmax } = k − k 2 − , k + k2 − , (8.21)
h̄ h̄
{a} {a} 2mω 2mω
{qmin , qmax } = k2 + − k, k2 + +k . (8.22)
h̄ h̄
with
1 e2 1 1
α= √ − (8.25)
2h̄ω h̄/2mω (∞) (0)
{e,a}
has been used to replace |C|2 by α. The scattering rate, 1/τ3D (k), may be evaluated
analytically through use of the identity
{e,a} {e,a}2 {e,a}2 {e,a}2
qmax q3 1 qmax + q02 q02 qmin − qmax
dq = ln {e,a}2 + $ %$ % .
{e,a}
qmin (q 2 + q02 )2 2 qmin + q02 2 q {e,a}2 + q 2 q {e,a}2 + q 2
max 0 min 0
(8.26)
$ %2
1 αω 1 2mω k + k 2 − 2mω/h̄ + q02
{e,a}
= (n q + 1) ln $ %2
τ3D (k) 2 k h̄
k − k 2 − 2mω/h̄ + q02
$ %2
k 2 + 2mω/h̄ + k + q02
αω 1 2mω
+ n q ln $ %2 . (8.27)
2 k h̄
k 2 − 2mω/h̄ − k + q02
†
UA
HFr = (−e)φ(q) eiq·r (aq + a−q ) (8.29)
q
2π e2 h̄ 1 iq·r †
=i e (aq + a−q )(ω⊥,TO
2
− ωq2 )(ωTO,
2
− ωq2 )
q V ωq q
× [(0)⊥ − (∞)⊥ ]ω⊥,TO
2
(ω,TO
2
− ωq2 )2 sin2 θ
−1/2
+ [(0) − (∞) ]ω,TO
2
(ω⊥,TO
2
− ωq2 )2 cos2 θ
1/2
4π e2 h̄V −1 1 iq·r †
=i e (aq + a−q ),
q (∂/∂ω)[(ω)⊥ sin θ + (ω) cos θ]
2 2 q
(8.30)
where
ω2 − ω⊥,LO
2 ω2 − ω,LO
2
⊥ (ω) = ⊥ (∞) , (ω) = (∞) , (8.31)
ω2 − ω⊥,TO
2 ω2 − ω,TO
2
and θ is the angle between the phonon wavevector q and the c-axis. Moreover, from
Section 3.2, equation (3.21),
ωLO,⊥
2 − ω2 ωLO,
2 − ω2
sin θ +
2
cos2 θ = 0 (8.33)
ωTO,⊥
2 − ω2 ωTO,
2 − ω2
or, equivalently,
where
ω12 = ωTO,
2
sin2 θ + ωTO,⊥
2
cos2 θ, ω22 = ωLO,
2
cos2 θ + ωLO,⊥
2
sin2 θ.
(8.35)
In Section 3.2 it was shown that when ωTO, − ωTO,⊥ ωLO, − ωTO, and
ωLO,⊥ − ωTO,⊥ this equation has roots
ω2 = 1
2 (ω12 + ω22 ) ± [(ω12 − ω22 ) + 2ω2 (θ)] , (8.36)
where
138 8 Carrier–LO-phonon scattering
(ωLO,
2 − ωLO,⊥
2 )(ωTO,
2 − ωTO,⊥
2 )
ω2 (θ) = 2 sin2 θ cos2 θ; (8.37)
ω22 − ω12
thus, one root is
2TO = ω2 = ωTO,
2
sin2 θ + ωTO,⊥
2
cos2 θ
(ωLO,
2 − ωLO,⊥
2 )(ωTO,
2 − ωTO,⊥
2 )
− sin2 θ cos2 θ
ω22 − ω12
≈ ωTO,
2
sin2 θ + ωTO,⊥
2
cos2 θ (8.38)
2LO = ω2 = ωLO,
2
cos2 θ + ωLO,⊥
2
sin2 θ
(ωLO,
2 − ωLO,⊥
2 )(ωTO,
2 − ωTO,⊥
2 )
+ sin2 θ cos2 θ
ω22 − ω12
≈ ωLO,
2
cos2 θ + ωLO,⊥
2
sin2 θ. (8.39)
The inequalities assumed for the derivation of 2TO and 2LO are satisfied for the
parameters of both GaN and AlN as may be verified from the numerical values of
the parameters given in Section 3.3. For the case of GaN, the dependence for the
phonon frequencies on θ is shown in Figure 8.1 for these infrared-active phonons. θ
is the angle between the phonon wavevector q and the c-axis.
As in subsection 8.1.1, the transition rate for electron–optical-phonon scattering
may be estimated by calculating the transition rate predicted by the Fermi golden
rule for the perturbation Hamiltonian, HFr UA . According to the Fermi golden rule, the
{e,a}
transition rate, SUA (k, k ), from an initial electron state |k and an initial phonon
4 4 5
state Nq + 1
±– denoted by k, Nq + 12 ± 12 – to a final electron state k and
1
2
6 2
6
a final phonon state Nq + 12 ± 12 – denoted by k , Nq + 12 ± 12 – per unit time
per unit volume is given by
{e,a} 2π {e,a} 2
SUA (k, k ) = MUA (q) δ(E(k ) − E(k) ± h̄ω), (8.40)
h̄ q
where
6 4
{e,a} UA
MUA (q) = k , Nq + 1
2 ± 12 HFr k, Nq + 1
2 ± 1
2 . (8.41)
2π e2 h̄ (ω⊥,TO
2 − ω,TO
2 )2 sin2 θ cos2 θ
=
V q 2 TO [⊥ (0) − ⊥ (∞)] ω⊥,TO
2 cos2 θ + (0) − (∞) ω,TO
2 sin2 θ
× n q + 12 ± 12 . (8.43)
This matrix element does not vanish, in general, since the TO-like mode is in reality
not a pure TO mode in a uniaxial material. In the limit of an isotropic material,
ω⊥,TO = ω,TO and the transverse matrix element vanishes. Likewise, for ω2 =
{e,a},LO 2
2LO , the matrix element MUA (q) is given by
/
2π e2 h̄ sin2 θ
{e,a},LO 2
MUA (q) =
V q 2 LO [1/⊥ (∞) − 1/⊥ (0)] ω⊥,LO
2
0−1
cos2 θ
+
1/ (∞) − 1/ (0) ω,LO
2
× n q + 12 ± 12 . (8.44)
In the isotropic limit, ω⊥,LO = ω,LO and the longitudinal matrix element reduces
to
2π e2 h̄ωLO
{e,a},LO 2
MUA (q) = [1/⊥ (∞) − 1/⊥ (0)] n q + 12 ± 12 . (8.45)
V q2
{a} 2
Numerical evaluation of Mq = MUA (q) indicates that the LO-like and TO-
like contributions to the absorption matrix element for bulk GaN are as shown in
Figure 8.2.
140 8 Carrier–LO-phonon scattering
The scattering rates for LO-like and TO-like phonons have been evaluated
numerically by Lee et al. (1997) for two cases: as a function of the incident angle of
the electron with respect to the c-axis, θk , for an initial electron energy of 0.3 eV; and
for θk = 0 and θk = π/2, for a range of electron energies. These results are depicted
in Figures 8.3 and 8.4. In Figure 8.3 the scattering rates for emission and absorption
of LO-like and TO-like phonons are presented as a function of the incident angle of
the electron with respect to the c-axis, θk , for an initial electron energy of 0.3 eV. In
Figure 8.4 the scattering rates for LO-like and TO-like absorption are presented for
a range of incident electron energies for θk = 0 and θk = π/2.
Figure 8.3. Scattering rates for emission and absorption of LO-like and TO-like
phonons as a function of the incident angle of the electron with respect to the c-axis,
θk , for an initial electron energy of 0.3 eV. From Lee et al. (1997), American
Physical Society, with permission.
142 8 Carrier–LO-phonon scattering
where
6 4
{e,a}
M2D (q) = k2D , Nq + 1
2 ± 12 HFr k2D , Nq + 1
2 ± 1
2 . (8.47)
In these expressions, the positive and negative signs are chosen to select the
desired initial and final phonon states. In the case of phonon emission, the final
phonon state contains one phonon more than the initial phonon6state; accordingly,
7
the initial and final states are selected to be k2D , Nq and k2D , Nq + 12 ± 12
respectively. For phonon absorption processes, the initial and final states are chosen
in a comparable manner. In the energy-conserving delta function in S {e,a} (k, k ), the
upper sign corresponds to phonon emission and the lower sign to phonon absorption.
As in subsection 8.1.1, the
electron and phonon states represent different spaces, so
7 7
we have |k2D Nq ≡ k2D , Nq . The phonon states are as defined in Section 5.1
and the Fröhlich Hamiltonian is given by equation (5.34),
C
HFr = aq eiq·r − aq† e−iq·r , (8.48)
q q
Figure 8.4. Scattering rates for LO-like and TO-like absorption are presented for a
range of incident electron energies for θk = 0 (solid line) and θk = π/2 (broken
line). From Lee et al. (1997), American Physical Society, with permission.
8.2 Fröhlich potential in quantum wells 143
where
2π e2 h̄ωLO 1 1
C = −i − . (8.49)
V (∞) (0)
The electron states are taken as plane waves in the directions parallel to the
heterointerfaces and as the ground state of an electron in an infinitely deep quantum
well in the z-direction (Leburton, 1984),
eik ·r 2 πz 5 e−ik ·r 2 πz
|k2D = √ sin and k2D = √ sin ,
A Lz Lz A Lz Lz
(8.50)
where A is the area of the heterointerface over which the electron wavefunction is
normalized.
{e,a}
Following the procedures of subsection 8.1.1, the matrix element M2D (q) is
then given by
{e,a} e−ik ·r+ik ·r 2 2 πz C
M2D (q) = d r $√ %2 3
sin
Lz Lz q
A
6 4
× Nq + 12 ± 12 (aq eiq·r − aq† e−iq·r ) Nq + 1
2 ± 1
2
C
= (∓)(n q + 1
2 ± 12 )1/2 δk −k ±q F(qz L z ), (8.51)
q
{e,a} L z |C|2
S2D (k , k ) = (n q + 12 ± 12 )I2D (q , L z )
4π h̄ 2
× δk −k ±q δ(E(k ∓ q ) − E(k ) ± h̄ω), (8.53)
where
2π 2 1 1
|C| =
2
e h̄ω − , (8.54)
V (∞) (0)
and
144 8 Carrier–LO-phonon scattering
+∞ |F(qz L z )|2
I2D (q , L z ) = dqz
−∞ q2 + qz2
/
0
π 1 1 S2 1
= Lz [1 − G(S)] + + 2− 2 G(S)
2 S2 2 S + π2 S2 + π 2
(8.55)
with
As in subsection 8.1.1, it has been assumed that the frequency of the phonon mode,
ωq , is independent of q and is equal to the zone-center phonon frequency, ωq .
{e,a} {e,a}
The total scattering rate, 1/τ2D (k ), associated with S2D (k , k ) results from
{e,a}
integrating S2D (k , k ) over q and multiplying by the area of the heterostructure,
V2D :
1 |C|2 L z V2D
{e,a}
= n q + 12 ± 12
τ2D (k ) 4π 2 h̄
× dφ dq I2D (q , L z )δ(E(k ∓ q ) − E(k ) ± h̄ω).
(8.57)
{e} 2mω
q± = k cos φ ± k2 cos2 φ − , (8.60)
h̄
and
{a} 2mω
q± = −k cos φ ± k2 cos2 φ + . (8.61)
h̄
{e}
Both roots of q± are allowed for φ when it is in the range 0 < φ < φmax , where
√ {a}
φmax = arccos h̄ω/E with E = h̄ 2 k2 /2m but with q− forbidden. For the case of
emission,
h̄ 2 {e,} {e,}
E(k ∓ q ) − E(k ) ± h̄ω = q − q+ q − q−
2m
h̄ 2 2mω
= q − 2q k cos φ +
2
. (8.62)
2m h̄
The integral over q is now performed by use of the relation
g(y)
dy g(y)δ( f (y) − a) = , (8.63)
d f /dy y=y0
where y0 is determined by f (y0 ) = a. Then, taking
h̄ 2 $ 2 %
f (q ) = q − 2q k cos φ , (8.64)
2m
it follows that
d f (q ) h̄ 2
= 2q − 2k cos φ
dq 2m
h̄ 2 2mω
=± k cos φ −
2 2 . (8.65)
m h̄
With these results, it follows that
1
{e}
τ2D (E)
φmax {e} {e} {e} {e}
I2D (q+ (E, φ))q+ (E, φ) + I2D (q− (E, φ))q− (E, φ)
αω
= (n q + 1) dφ ,
π 0 (E/h̄ω) cos2 φ − 1
(8.66)
and by an analogous derivation (Leburton, 1984),
π {a} {a}
I2D (q+ (E, φ))q+ (E, φ)
1 αω
{a}
= nq dφ , (8.67)
τ2D (E) π 0 (E/h̄ω) cos2 φ + 1
where
1 e2 1 1
α= √ − . (8.68)
2h̄ω h̄/2mω (∞) (0)
146 8 Carrier–LO-phonon scattering
carrier states are now those of an infinitely deep quantum well. According to the
{e,a}
Fermi golden rule, the transition rate, S1D (k, 4 k ), from an initial electron state
4
|k1D and an initial phonon state Nq + 12 ± 12 – denoted by k1D , Nq + 12 ± 12 –
5 6
to a final electron state k1D and a final phonon state Nq + 1 + 1 – denoted by
6 2 2
, N + 1 + 1 – per unit time per unit heterostructure area A is given by
k1D q 2 2
where
6 4
{e,a}
M1D (q) = k1D , Nq + 1
2 + 12 HFr k1D , Nq + 1
2 ± 1
2 . (8.70)
Figure 8.5. Main figure: total electron–optical-phonon scattering rates in bulk GaN
(thin lines) and in a free-standing GaN quantum well (thick solid line) as functions
of the electron energy E k over a range of 0 to about 500 meV. Scattering rates at
three particular angles for bulk GaN crystals are shown as follows: thin solid lines,
θk = π/2; thin dotted lines, θk = π/4; thin broken lines, θk = 0. The broken and
dotted line indicates the scattering rate calculated for confined electrons and bulk
phonons.
Inset: the LO-like and TO-like emission and absorption rates in bulk GaN for
three different angles: solid lines, θk = π/2; dotted lines, θk = π/4; broken lines,
θk = 0. From Komirenko et al. (2000a), American Physical Society, with
permission.
148 8 Carrier–LO-phonon scattering
In these expressions, the positive and negative signs are chosen to select the desired
initial and final phonon states. In the case of phonon emission, the final phonon
state contains one phonon more than the initial phonon4 state;6 accordingly, the
, N + 1 ± 1
initial and final states are selected to be k1D , Nq + 12 ± 12 and k1D q 2 2
respectively. For phonon absorption processes the initial and final states are chosen
in similar manner. In the energy-conserving delta function in S {e,a} (k, k ), the upper
sign corresponds to phonon emission and the lower sign corresponds to phonon
absorption. As in subsection 8.1.1, the electron
and phonon states represent different
7 7
spaces, so that we have |k1D Nq ≡ k1D , Nq . Again, the phonon states are as
defined in Section 5.1 and the Fröhlich Hamiltonian is given by equation (5.34) of
Section 5.2,
C
HFr = aq eiq·r − aq† e−iq·r , (8.71)
q q
where
2π e2 h̄ωLO 1 1
C = −i − . (8.72)
V (∞) (0)
The electron states are taken as plane-wave states along the quantum-wire axis,
taken here as the x-axis, and as the ground states of an electron in an infinitely deep
quantum wire in the y- and z-directions (Leburton, 1984),
eik x x 2 πy 2 πz
|k1D = √ sin sin , (8.73)
Lx Ly Lz Lz Lz
and
−i
A− (qz )± = √ [a± (qx , q y ) − a± (qx , −q y )], (8.74)
2
where L x is the length in the x-direction over which the free-electron wavefunction
is normalized and L y and L z are the height and the width of the quantum wire; it is
assumed that L x L y , L z .
{e,a}
Following the procedures of subsection 8.1.1, the matrix element, M1D (q), is
then given by
{e,a} e−ik x x+ik x x 2 πy 2 πz
M1D (q) = d 3 r √ 2 sin2 sin2
Lx Ly L y Lz Lz
6 7
C
× Nq + 12 + 12 (aq eiq·r − aq† e−iq·r ) Nq
q
C
= (∓)(n q + 12 ± 12 )1/2 δk x −k x ±qx F(q y L y )F(qz L z ), (8.75)
q
where F(q y L y ) and F(qz L z ) are defined by
Ly
2 πy
F(q y L y ) = dy e−iq y y sin2
Ly 0 Ly
π2 sin q y L y /2 −iq y L y /2
= e ,
π − (q y L y /2) (q y L y /2)
2 2
Lz (8.76)
2 πz
F(qz L z ) = dz e−iqz z sin2
Lz 0 Lz
π 2 sin qz L z /2 −iqz L z /2
= 2 e .
π − (qz L z /2)2 (qz L z /2)
Then, following the procedure of subsection 8.1.1, it follows that
{e,a} L z |C|2
S1D (k x , k x ) = nq + 1
2 ± 1
2 I1D (qx , L y , L z )
4π h̄ 2
× δk x −k x ±qx δ(E(k x ∓ qx ) − E(k x ) ± h̄ω), (8.77)
where
150 8 Carrier–LO-phonon scattering
2π 2 1 1
|C|2 = e h̄ω − (8.78)
V (∞) (0)
and
+∞ +∞ F(q y L y )2 |F(qz L z )|2
I1D (qx , L y , L z ) = dq y dqz , (8.79)
−∞ −∞ qx2 + q y2 + qz2
the Fröhlich interaction Hamiltonian for a quantum well, HFr 2D , was obtained
from HFr by imposing the boundary condition that HFr (x, y, z)z=L /2 = 0. In
3D 3D
z
particular, the analysis of subsection 7.3.2 may be extended by applying not just
the boundary condition HFr 3D (x, y, z) = 0 but also the boundary condition
z=L z /2
H (x, y, z)
3D = 0. Starting with the canonical form of the bulk Fröhlich
Fr y=L y /2
interaction Hamiltonian,
2πe2 h̄ωLO 1 1 1
3D
HFr =− − †
(aq + a−q ) e−iq·r , (8.80)
V (∞) (0) q q
we write the sum over q as a sum over positive values of qz in equation (8.84):
HFr = Vq e−iq ·r e−iqz z aq ,qz + a−q
†
,−qz
q ,qz >0
†
+ eiqz z aq ,−qz + a−q ,qz
, (8.81)
where
2πe2 h̄ωLO 1 1
Vq = − − . (8.82)
V (∞) (0)
Then, defining
1 −i
a+ (q ) = √ (aq ,qz + aq ,−qz ) a− (q ) = √ (aq ,qz − a−q ,qz ),
2 2
(8.83)
with
8.3 Scattering of carriers by LO phonons in quantum wires 151
1
A+ (qz )± = √ [a± (qx , q y ) + a± (qx , −q y )],
2
(8.84)
−i
A− (qz )± = √ [a± (qx , q y ) − a± (qx , −q y )],
2
expanding e±iq y y and e±iqz z , and taking q y = ±mπ/L y and qz = ±mπ/L z to
enforce the electromagnetic boundary conditions that the confined phonon modes
vanish at y = ±L y /2 and z = ±L z /2, leads to the result (Stroscio, 1989)
1D
HFr
mπ y nπ z
= 2α e−iqx x λmn cos cos [A+ (qx )+ + A†+ (−qx )+ ]
qx m=1,3,5,... n=1,3,5,...
Ly Lz
mπ y nπ z
+ λmn cos sin [A+ (qx )− + A†+ (−qx )− ]
m=1,3,5,... n=2,4,6,...
Ly Lz
mπ y nπ z
+ λmn sin cos [A− (qx )+ + A†− (−qx )+ ]
m=2,4,6,... n=1,3,5,...
Ly Lz
mπ y nπ z †
+ λmn sin sin [A− (qx )− + A− (−qx )− ] ,
m=2,4,6,... n=2,4,6,...
Ly Lz
(8.85)
where
/ 2 2 0−1/2
mπ nπ
λmn = qx2 + + (8.86)
Ly Lz
and
! "1/2
2π e2 1 1
α = h̄ω − . (8.87)
V (∞) (0)
Equation (8.85) may be written more compactly as (Campos et al., 1992),
∞
∞
1D
HFr = 2α e−iqx x ξm (y)ξn (y)λmn βmn , (8.88)
qx m=1 n=1
where
nπ t π
ξn (t) = sin + δn (8.91)
Lt 2
with δn = 1 for odd n and δn = 0 for even n.
The Hamiltonian (8.88) includes contributions only from the confined modes of
the dielectric continuum model with electrostatic boundary conditions – also known
as the slab modes. The interface and half-space modes are not included and must be
considered separately. For calculations based on the guided modes, the contributions
to the Hamiltonian in the interior of the quantum wire may be written in terms of
the Hamiltonian for the slab modes by taking
nπ t π
ξn (t) = cos + δn . (8.92)
Lt 2
Of course, it is necessary to augment this expression with terms containing
the evanescent contributions of the guided modes in the regions surrounding the
quantum wire. In the present account, scattering rates will be calculated within the
dielectric continuum model with electrostatic boundary conditions. As discussed in
Appendix C, intrasubband and intersubband transitions rates are the same in either
the slab or guided models as long as complete sets of orthogonal phonon modes are
used in performing the calculation.
{e,a}
As in the previous subsection, the transition rate, S1D (k, k ), from
4 an ini-
tial electron state |k1D and an initial phonon state Nq + 2 ± 2 – denoted
1 1
4 5
and a final phonon state
by k1D , Nq + 12 ± 12 – to a final electron state k1D
6 6
, N + 1 + 1 – per unit time per unit length, L,
Nq + 12 + 12 – denoted by k1D q 2 2
is given by
{e,a} 2π {e,a} 2
S1D (k, k ) = M1D (q) δ(E(k ) − E(k) ± h̄ω), (8.93)
h̄ q
where
6 4
{e,a} 1D
M1D (q) = k1D , Nq + 1
2 + 12 HFr k1D , Nq + 1
2 ± 1
2 . (8.94)
within the finitely long quantum wire with cross-sectional area defined by −L y <
y < L y /2, −L z < z < L z /2 and is taken to vanish outside this region. In this
so-called extreme quantum limit, the electron energy for the ground state is
h̄ 2 k x2 h̄ 2 π 2 1 1
E 1D = + + 2 . (8.96)
2m ∗ 2m ∗ L 2y Lz
8.3 Scattering of carriers by LO phonons in quantum wires 153
The four expressions for Pmn correspond to the four terms of (8.85). The numerical
values for the dominant values of Pmn are: P11 = (8/3π)2 , P13 = P31 =
(1/5)P11 , P15 = P51 = −(1/35)(8/3π )2 , P33 = (1/25)(8/3π)2 , P35 = P53 =
−(1/175)(8/3π)2 , and P55 = (1/1225)(8/3π )2 . The transition rate is then given
by
1
{e,a}
τ1D (k x )
+∞
|α |2 V
= nq + 1
2 ± 1
2 dqx I1D (qx , L y , L z )δ(E(k x ± qx ) − E(k x ) ± h̄ω),
4π 2 h̄ −∞
(8.99)
where
(2π )2
I1D (qx , L y , L z ) = 16 |Pmn |2 λmn . (8.100)
L y Lz m=1,3,5,... n=1,3,5,...
Figure 8.7 provides numerical values for I1D (qx , L y , L z ) for wires with selected
cross-sectional dimensions.
Rewriting the argument of the delta function δ(E(k x ±qx )− E(k x )± h̄ω) through
the use of
h̄ 2 {e,a} {e,a}
E(k x ± qx ) − E(k x ) ± h̄ω = ∗
q x − q+ q x − q− , (8.101)
2m
where
1/2
{e,a} 2m ∗ ω
q± = k x ± k x2 − (8.102)
h̄
154 8 Carrier–LO-phonon scattering
with
1 e2 1 1
α= √ − . (8.104)
2h̄ω h̄/2mω (∞) (0)
Then
E(r) = −∇(r) = E(qx , y, z) e−iqx x ,
qx
(8.106)
P(r) = (ω)E(r) = P(qx , y, z) e−iqx x .
qx
where the subscript i refers to a general material region i. From the results of
subsection 7.3.4 for a quantum well, it is clear that
where the sum over i is over the two material regions in the problem at hand: i = 1
inside and i = 2 outside the quantum wire. The Fröhlich interaction Hamiltonian,
1D , is then given by
HIF
1D
HIF = −e (qx , y, z) e−iqx x [AIF (qx ) + A†IF (−qx )], (8.110)
qx
where AIF (qx ) and A†IF (−qx ) are the annihilation and creation operators for the
appropriate IF-phonon modes of the quantum wire and where in the separation-of-
156 8 Carrier–LO-phonon scattering
for the antisymmetric IF modes; in these potentials, α and β satisfy the relation
α 2 + β 2 − qx2 = 0, (8.113)
and in (8.112) the plus sign is taken when yz > 0 and the minus sign when
yz < 0. As usual, the dispersion relations are obtained by imposing the boundary
conditions. For the symmetric solution this dispersion relation is
where
αL y = β L z (8.116)
must also be satisfied for both modes. These dispersion relations are of the same
forms as those for a quantum well, as is expected since the separation-of-variables
approximation has been assumed thus far in this treatment of the optical phonons
in quantum wires. The numerical solutions for these dispersion relations are
8.3 Scattering of carriers by LO phonons in quantum wires 157
depicted in Figure 8.8 for the case where L y = L z = d. In this figure the
dispersion relations are given for a GaAs quantum wire in two cases: a free-
standing quantum wire having frequencies ω S, f and ω A, f for the symmetric and
antisymmetric interface modes respectively, and a GaAs quantum wire embedded
in AlAs having frequencies for the symmetric and antisymmetric interface modes
of ω S,± and ω A,± respectively. The subscripts ± designate the high-frequency
(+) and low-frequency (−) modes for both the symmetric and antisymmetric
cases.
{e,a} 2π {e,a} 2
S1D,IF (k, k ) = M1D,IF (qx ) δ(E(k ) − E(k) ± h̄ω), (8.118)
h̄ q
6 4
{e,a} 1D
M1D,IF (qx ) = k1D , Nq + 12 + 12 HIF k1D , Nq + 12 ± 12 . (8.119)
The electron wavefunction and the electron energy for the ground state are as
given in the last subsection. Then the procedures of the last subsection lead
straightforwardly to the equation
{e,a} {e,a}
1 α ω 1 + I1D,IF q+ + I1D,IF q−
{e,a}
= nq + √ , (8.120)
τ1D,IF (k x ) 2π 2 E/h̄ω −
1/2
{e,a} 2m ∗ ω
q± = k x ± k x2 − (8.121)
h̄
and
1 e2
α = √ . (8.122)
2h̄ω h̄/2mω
Furthermore,
(2π )2 L
I1D,IF = C 2 Ps2 , (8.123)
ω2 4π
where
1
Ps =
cosh αL y /2 cosh β L z /2
L y /2 L z /2
dy dz πy πz
× cos2 cos2
−L y /2 L y /2 −L z /2 L z /2 Ly Lz
× cosh αy cosh βz
1
=
cosh αL y /2 cosh β L z /2
4 sinh(αL y /2) sinh (β L z /2)
× . (8.124)
αβ L y L z (αL y /2π )2 + 1 (β L z /2π)2 + 1
Only the symmetric mode contributes to the scattering rate for the ground state
electronic wavefunction in the extreme quantum limit.
{e,a} {e,a}
The scattering rate 1/τ1D,IF (k x ) has the same form as 1/τ1D (k x ), (8.103), of
the last subsection. Indeed, α and α differ only by the factor −1 (∞) − −1 (0) ,
and I1D is replaced by I1D,IF . Otherwise the expressions are identical. Therefore, it
follows that the scattering rate due to both confined optical phonons and interface
optical phonons is given by
{e,a} {e,a}
1 α ω 1 + Iwire q+ + Iwire q−
{e,a}
= nq + √ , (8.125)
τwire (k x ) 2π 2 E/h̄ω −
where
1 1
Iwire = − I1D + I1D,IF . (8.126)
(∞) (0)
160 8 Carrier–LO-phonon scattering
The electron–optical-phonon scattering rates for emission and absorption for se-
lected phonon modes are displayed in Figure 8.9 for a 40 Å × 40 Å GaAs quantum
wire embedded in AlAs. The interface phonons are labeled as SO (surface-optical)
modes and the confined phonons are labeled simply as LO modes.
The one-dimensional density-of-states peaks are prominent features of the curves
representing the phonon emission rates. The electrons are in the ground states in the
y- and z-directions. The maximum scattering rates for SO-phonon emission occur
at the energies of the symmetric high-frequency interface mode (about 50 meV) and
the symmetric low-frequency interface mode (about 33 meV).
Knipp and Reinecke (1992) examined the accuracy of using the separation-
of-variables approximation to the dielectric continuum model for calculating the
Fröhlich potential of the interface optical phonons in quantum wires. They consid-
Figure 8.10. Cross sections considered by Knipp and Reinecke (1992). The
long-broken line represents an elliptical cross section. The shorter broken line
designates a cross section that is basically elliptical but has flattened sides. The solid
and dotted lines (visible at the corners) represent ‘rectangles’ with rounded corners.
The geometrical parameters R and r are as indicated in this figure. From Knipp and
Reinecke (1992), American Physical Society, with permission.
162 8 Carrier–LO-phonon scattering
Figure 8.11. Interface modes in a nearly rectangular quantum wire. The Fröhlich
potential φ in the plane of the quantum wire is depicted for the mode of highest
azimuthal symmetry (m = 0) for a quantum wire with R/r = 2 and for a corner
curvature a = r/10; qr = 2. From Knipp and Reinecke (1992), American Physical
Society, with permission.
8.3 Scattering of carriers by LO phonons in quantum wires 163
energy from phonon absorption. This latter rate can be significant in quantum wires
and quantum wells, since the phonons emitted by energetic carriers can accumulate
in these structures. Indeed, the phonon densities in many dimensionally confined
semiconductor devices are typically well above those of the equilibrium phonon
population and there is an appreciable probability that these non-equilibrium – or
‘hot’ – phonons will be reabsorbed. The net loss of energy by a carrier in such a
situation depends on the rates for both phonon absorption and emission.
The lifetimes of the optical phonons are also important in determining the total
energy loss rate for such carriers. As discussed in Chapter 6, the longitudinal optical
(LO) phonons in GaAs and in many other polar materials decay into acoustic
phonons through the Klemens’ channel. Over a wide range of temperatures and
phonon wavevectors, the lifetimes of these longitudinal optical phonons in GaAs
vary from a few picoseconds to about 10 ps (Bhatt et al., 1994). Typical lifetimes for
other polar semiconductors are also of this magnitude. The LO phonons undergoing
decay into acoustic phonons are not available for absorption by the carriers and,
Figure 8.12. The Fröhlich potential in the plane of an elliptical quantum wire with
R/r = 2. This potential was derived analytically by Knipp and Reinecke (1992) for
the mode of highest azimuthal symmetry (m = 0), with qr = 2. From Knipp and
Reinecke (1992), American Physical Society, with permission.
164 8 Carrier–LO-phonon scattering
as a result of the Klemens’ channel, the carrier energy loss is more rapid than it
would be otherwise; this phenomenon is referred to as the ‘hot-phonon-bottleneck’
effect. In fact, the numerical calculations of Campos et al. and Das Sarma et al. for
a 50 Å × 50 Å GaAs quantum wire indicate, over a temperature range from 50 K to
300 K, for wire dimensions from 50 Å × 50 Å to 500 Å × 500 Å, and for a linear
density range of 104 to 106 cm−1 , that the hot-phonon-bottleneck effect changes the
net power loss by an order of magnitude or more while the effect of carrier screening
results in a change of only, roughly, a factor of two.
Figure 8.13 illustrates these findings for the case of a linear charge density of
106 cm−1 (Campos et al., 1992). The power loss P per carrier is given as a function
of the inverse electron temperature for both the slab modes and the guided modes.
The curve of highest power loss (long-broken line) is calculated for the slab modes
with no screening effects. The thinner solid line represents the power loss with
screening effects included for a phonon decay time τphonon of 0 ps; the thicker solid
line represents the power loss with screening effects included for a phonon decay
time τphonon of 7 ps. For the case of the guided modes, the unscreened power loss
function is represented by the broken-and-dotted line, and the screened power loss
functions for τphonon = 0 ps and τphonon = 7 ps are represented by the thinner and
thicker short-broken lines respectively.
Figure 8.13. Results of Campos et al. for a 50 Å × 50 Å GaAs quantum wire for the
case of a linear charge density of 106 cm−1 . The power loss per carrier is given as a
function of the inverse electron temperature. The curve of highest power loss
(long-broken line) is calculated for the slab modes with no screening effects. For the
guided modes, the unscreened power loss function is represented by the
broken-and-dotted line and the screened power loss functions for τphonon = 0 ps and
τphonon = 7 ps are represented by the thinner and thicker short-broken lines
respectively. From Campos et al. (1992), American Physical Society, with
permission.
8.3 Scattering of carriers by LO phonons in quantum wires 165
Thus while dynamical screening effects can be noticeable the effects of the
hot-photon bottleneck can lead to a change of an order of magnitude or more in the
carrier energy loss rates associated with carrier–LO-photon scattering in quantum
wires. The screening calculations of Campos et al. (1992) and Das Sarma et al.
(1992) involve non-trivial numerical calculations. For additional information on
dynamical screening and hot-phonon effects the reader is referred to these papers.
where the constants A, B, and C as well as the dispersion relation for the mode
are determined by the conditions that φi and i ∂φi /∂z are continuous at the
two heterointerfaces. It is straightforward to show by imposing these boundary
conditions that
−1
1 m
A = eqd 1+ ,
1 − 2 qδ1
1 −qd 2
B= e 1− , (8.128)
2 1
1 qd 2
C= e 1+
2 1
and that the dispersion relation is
1 − 2 −2qd qδ1 − m
e − 1 = 0. (8.129)
1 + 2 qδ1 + m
In these results, the dielectric function of the metal is given by
ω2p τ 2
m = 1 − , (8.130)
1 − ω2 τ 2
where ω p is the plasma frequency for the carriers in the metal and τ is the dielectric
relaxation time for the carriers in the metal.
Since the ratio |A/B| is a measure of the amplitude of the IF optical phonon
mode in the metal and since
A 2
=
B 1 + /qδ 1, (8.131)
m 1
where
2π e2 h̄ωLO 1 1
C = −i − . (8.133)
V (∞) (0)
Consider a quantum dot with the geometry of a free-standing cube each of whose
sides is of length L (de la Cruz et al., 1993). Then by writing the sum over q as a
where
$ 1/2
mπ %2 $ nπ %2 $ pπ %2
β(m, n, p) = + + , (8.135)
L L L
8.4 Scattering of carriers and LO phonons in quantum dots 169
and where the phonon annihilation operators for the zero-dimensional confined
modes are
1
A++ (q)± = √ A+ (qx )± + A+ (−qx )± ,
2
−i
A+− (q)± = √ A+ (qx )± − A+ (−qx )± ,
2
(8.136)
1
A−+ (q)± = √ A− (qx )± + A− (−qx )± ,
2
−i
A−− (q)−− = √ A− (qx )± − A− (−qx )± .
2
The operators A± (qx )± can be written in terms of the annihilation operator for a
three-dimensional semiconductor, in a manner analogous to that employed for the
quantum wire (Stroscio, 1989), through use of the relations
1
A+ (qx )± = √ a± (qx , q y ) + a± (qx , −q y ) ,
2
(8.137)
−i
A− (qx )± = √ a± (qx , q y ) − a± (qx , −q y ) ,
2
where a± (qx , q y ) are defined by
1
a+ (qx , q y ) = √ a(qx , q y , qz ) + a(qx , q y , −qz ) ,
2
(8.138)
−i
a− (qx , q y ) = √ a(qx , q y , qz ) − a(qx , q y , −qz ) .
2
For the case of a GaAs quantum dot in vacuum, the electron energy levels are
generally separated by more than the typical longitudinal optical phonon energy,
h̄ωLO , unless the dimensions of the dot are such that L is greater than several
hundred ångstroms. Phonon confinement effects are generally not significant for
such large confinement lengths. Specifically, for such a GaAs quantum ‘box’ the
confined phonons and the interface phonons have energies within a few meV of the
bulk phonon energy, 36 meV. In the case where the quantum dot is embedded, in, say,
AlAs, there would be interface phonons with energies close to 50 meV but even these
energies are small compared to typical separations between electron energy states
unless the length L is several hundred ångstroms or more. For holes, the de Broglie
wavelength is considerably shorter and energy level separations are smaller than
for the case of electrons. Accordingly, for transitions involving holes the quantum
‘box’ may be small enough for phonon confinement effects to be significant. It will
be instructive to summarize key features of the calculation of de la Cruz (1993)
because it illustrates the general procedure within the effective-mass approximation
and the Fermi golden rule approximation (Appendix E).
170 8 Carrier–LO-phonon scattering
For a ‘box’ in vacuum with infinitely high potential barriers at its boundaries,
defined by −L/2 ≤ x, y, z ≤ L/2, the carrier wavefunctions are represented by
3 2 iπ x jπ y lπ z
|i, j, l = cos cos cos , for odd i, j, l,
L L L L
(8.139)
3 2 iπ x jπ y lπ z
|i, j, l = sin sin sin , for even i, j, l.
L L L L
The carrier eigenenergies for this system in the effective-mass approximation, with
effective mass m ∗ , are given by
/ 2 2 2 0
h̄ 2 iπ jπ lπ
E i, j,l = + + . (8.140)
2m ∗ L L L
2π {e,a} 2
W {e,a} = M δ(E i , j ,l − E i, j,l ± h̄ωLO ), (8.141)
h̄
with
5 7
(0D)
M {e,a} = i , j , l ; Nq + 1
2 ± 1
2 HFr i, j, l; Nq + 1
2 ± 1
2 (8.142)
where, as discussed several times previously, the signs are chosen to describe
the desired phonon process. The energy-conservation constraint required by the
δ-function may be expressed as
$ % $ % $ % 2m ∗ L 2
i 2 − i 2 + j 2 − j 2 + l 2 − l 2 ± ωLO = 0. (8.143)
h̄π 2
For the system under consideration, non-zero matrix elements correspond to tran-
sitions between states of the same parity. Accordingly, the transitions corresponding
to the smallest quantum boxes are those between the |1, 1, 1 states and states where
the carrier quantum numbers, i , j , l and i, j, l, are selected from 1, 3, and 5.
For these latter states, all the degenerate states for a given set of carrier quantum
numbers are denoted by {|i, j, l}. The transition probabilities for a selection of
such transitions were calculated by de la Cruz (1993) for GaAs at room temperature.
These probabilities are summarized in Table 8.1, where, in the case of emission, the
carrier transition occurs between the indicated set of degenerate states {|i, j, l} and
the |1, 1, 1 state. Likewise, for the case of absorption, the initial state is the |1, 1, 1
state and the final state is indicated by the specific set {|i, j, l} corresponding to the
transition.
8.4 Scattering of carriers and LO phonons in quantum dots 171
unpert α
Hα,def = E αstrained (k) − E α (k) = E µν1 Sµν , (9.1)
where Sµν is the strain tensor defined in Section 7.2, (7.27), and the deformation
potential is the difference in the energy of band α when strain is present, E αstrained (k),
unpert
and the energy of band α when there is no strain, E α (k). Since the phonon field is
always present, the case where there is no strain is unphysical; it corresponds to the
situation where there is no deformation potential. For cubic crystals and for carriers
experiencing a spherically symmetric energy surface, symmetry considerations
make it evident that the crystalline potential should be proportional to ∇V /V , as
in Section 5.3. This conclusion holds for carriers at a non-degenerate point but
172
9.1 Carrier–acoustic-phonon scattering in bulk zincblende structures 173
not for the cases of other symmetry points such as the X or L points. Accordingly,
for the case of a cubic crystal and for carriers at the point,
Hα,def = E 1α ∇V /V = E 1α ∇ · u, (9.2)
Hα,def = E 1α , (9.3)
where the minimum is on an axis along the unit vector î. For this more general case,
α term contributes to the deformation potential and, in general, the TA and
the E 1t
optical phonons produce deformation potentials leading to intravalley processes.
where Sµν is the strain tensor and eλ,µν is the generalized tensor describing the
piezo
piezoelectric coupling of the polarization, Pλ , to the strain field Sµν . In practice,
the components of eλ,µν may be determined through knowledge of the strain and
electromagnetic fields and the static dielectric constant, λµ
0 , via the relation
1 0
eλ,µν Sµν = Pλ − ( − δλµ )E µ . (9.6)
4π λµ
As described previously (Vogl, 1980), eλ,µν is manifestly a bulk quantity. In
general, Pλ and E λ depend on the geometry of the crystal structure but the
174 9 Carrier–acoustic-phonon scattering
3 × 3 matrix with all non-diagonal elements equal to zero. For cubic structures,
qλ λµ
0 q reduces to 0 q · q since the three diagonal elements of 0 are all equal to
µ λµ
0.
For non-cubic crystals it is convenient to use the tensor notation of Section 7.2 to
piezo piezo
determine Pλ from Pλ = eλ,µν Sµν . As in Section 7.2, Sµν may be represented
piezo
as a six-element vector, Pλ is a three-element vector, and eλ, µν is a 3 × 6 matrix;
for zincblende structures (Auld, 1973), the 3 × 6 matrix is
0 0 0 ex4 0 0
0 0 0 0 ex4 0 (9.10)
0 0 0 0 0 ex4
and for würtzite crystals this matrix takes the form
0 0 0 0 ex5 0
0 0 0 ex5 0 0 . (9.11)
ex1 ex2 ex3 0 0 0
Hdef = E 1 ∇V /V = E 1 ∇ · u = E 1 , (9.12)
and with the expression for the quantized displacement derived in (7.145), Section
7.6 (Bannov, 1995), it follows that
h̄ 2 3
Hdef = E 1 †
(an,q + an,−q )∇ · wn (q , z) eiq ·r
q ,n 2L ρωn (q )
2
= eiq ·r (q , n, z)(an,q + an,−q
†
) (9.13)
q ,n
where
E 12 h̄ ∂wnz (q , z)
(q , n, z) = iq · wn (q , z) + . (9.14)
2L 2 ρωn (q ) ∂z
and by normalizing the acoustic phonon Fourier amplitude, u(q, r, ϕ, z), according
to
2π a
1 h̄
dϕ dr r u(q, r, ϕ, z) · u∗ (q, r, ϕ, z) = . (9.16)
πa 2 0 0 2Mω q
Here a is the radius of the quantum wire, N is the number of unit cells in the
normalization volume V , aq is the phonon annihilation operator, q is the phonon
wavevector, ωq is the angular frequency of the phonon mode, M is the mass of the
ions in the unit cell, and r, ϕ, z are the usual cylindrical coordinates. As discussed
in Chapter 7, the elastic continuum model provides an adequate description of such
176 9 Carrier–acoustic-phonon scattering
a quantum wire as long as the diameter of the quantum wire is several times larger
than the linear dimension of the unit cell. Within the elastic continuum model, the
equation describing the acoustic phonon modes is (Yu et al., 1996)
∂ 2u
ρ = c44 ∇ 2 u + (c12 + c44 )∇(∇ · u)
∂t 2
where c∗ = c11 − c12 − 2c44 . For the isotropic case of Section 7.6, c∗ = c11 − c12 −
2c44 = 0, and
∂ 2u
ρ = c44 ∇ 2 u + (c44 − c11 )∇(∇ · u)
∂t 2
= ρct2 ∇ 2 u + ρ(cl2 − ct2 )∇(∇ · u), (9.18)
or, equivalently,
∂ 2u
= cl2 ∇(∇ · u) − ct2 ∇ × ∇ × u, (9.19)
∂t 2
√
where cl = c11 /ρ is the sound speed for longitudinal acoustic waves and ct =
√
c44 /ρ is the sound speed for transverse acoustic waves. In considering the acoustic
phonon modes in a cylinder it is convenient to introduce two equivalent quantities,
the Young’s modulus velocity v0 and the Poisson ratio σ , through
2 (cl /ct ) − 1
1
(3cl2 − ct2 ) 2
c0 = cl and σ = . (9.20)
cl2 − ct2 (cl /ct )2 − 1
ω2
2
ql,t = − q 2. (9.22)
vl,t
2
9.3 Carrier–acoustic-phonon scattering in quantum wires 177
here β = B/A.
In classical acoustics (Auld, 1973), there are two particularly simple types
of boundary condition. They are referred to frequently as the cardinal boundary
conditions and are the free-surface boundary condition (FSBC) and the clamped-
surface boundary condition (CSBC). In the free-surface boundary condition, at the
boundary between the cylindrical wire and the vacuum the normal components of
the stress tensor are zero and the displacement is unrestricted. In the clamped-surface
boundary condition, at the boundary between the cylindrical wire and the vacuum
the normal components of the stress tensor are unrestricted and the displacement is
zero. The dispersion relation for the free-surface boundary condition is then given
by
(q̃ 2 − q̃t2 )T1 (q̃l π ) − 2q̃l2 (q̃ 2 + q̃t2 ) + 4q̃ 2 q̃l2 T1 (q̃t π) = 0, (9.26)
and the dispersion relation for the clamped-surface boundary condition is given by
where
q̃t J1 (q̃t π )
β=− . (9.30)
q̃l J1 (q̃l π )
eikz
|k = √ ψ(r, θ), (9.31)
Lz
where
1 X 10
ψ(r, θ ) = √ J0 r , (9.32)
π Y10 a a
X 10 being the position of the first zero of J0 (x) and Y10 being equal to J1 (X 10 ); L z ,
the length of the cylinder, is assumed to be much larger than a. The energy of the
ground state electron energy is [(h̄k)2 + (X 10 /a)2 ]/2m ∗ , where m ∗ is the electron’s
effective mass. Then it follows that the matrix element of the deformation potential,
Hdef , takes the form
6
Mn{e,a} = k , Nn + 12 + 12 Hdef |k, Nn
h̄ a
= E1 √ βn (q 2 + ql2 )
2N Mωn 2sn
#
× F(ql a) Nn + 12 + 12 δ−k +k−q , (9.33)
1 2π {e,a} 2
= Mn δ(E(k ) − E(k) + h̄ω)
τ {e,a} (k) h̄
π m ∗ |E 1 |2
= 0 4 2
n Y1 h̄ ρ
1 1 {e,a} {e,a}
× 4 2
I1cn q̃+ + I1cn q̃− , (9.35)
c0 c̃l a
with
|βn (q̃)|2 $ %
I1cn (q̃) = |ω̃n |3 |Fn (q̃l π )|2 Nn + 12 + 12
|sn (q̃)|
1
× . (9.36)
∗
(q̃ + k̃) − (m c0 a/π h̄)(d ω̃/d q̃)
{e,a}
From energy- and momentum-conservation conditions it follows that q̃± is the
solution of (h̄ 2 /2m ∗ )(q̃ 2 ± 2q̃ k̃) − h̄ω = 0; accordingly, the plus sign is taken for
forward scattering and the negative sign for backward scattering. As usual, n labels
the quantized acoustic phonon modes.
For the piezoelectric scattering of carriers in a cylindrical quantum wire, the
normalization procedure is of course the same as that for scattering by a deformation
potential but the interaction Hamiltonian is based on the interaction potential of
subsection 9.1.2. For cubic materials this piezoelectric interaction potential takes
the form
qλ eλ,µν qµ
V piezo (q) = 4π uν . (9.37)
0q · q
In cylindrical coordinates the piezoelectric tensor of subsection 9.1.2 takes the form
(Auld, 1973)
0 0 0 ex4 A(ϕ) 2ex4 B(ϕ) 0
e= 0 0 0 −2ex4 B(ϕ) ex4 A(ϕ) 0
2ex4 B(ϕ) −2ex4 B(ϕ) 0 0 0 ex4 A(ϕ)
(9.38)
where A(ϕ) = cos2 ϕ − sin2 ϕ and B(ϕ) = cos ϕ sin ϕ. Moreover, the polarization
vector P and strain components Sν are given by
P1
P = P2 (9.39)
P3
and
180 9 Carrier–acoustic-phonon scattering
∂u r ur 1 ∂u ϕ
S1 = Srr = , S2 = Sϕϕ = + ,
∂r r r ∂ϕ
∂u z ∂u ϕ 1 ∂u z
S3 = Szz = , S4 = 2Szϕ = 2Sϕz = + ,
∂z ∂z r ∂ϕ
(9.40)
∂u r ∂u z
S5 = 2Sr z = 2Szr = + ,
∂z ∂r
1 ∂u r ∂u ϕ uϕ
S6 = 2Sr ϕ = 2Sϕr = + − .
r ∂ϕ ∂r r
Accordingly, the relation P = e S implies that (Stroscio and Kim, 1993)
P1 = 2ex4 (cos2 ϕ − sin2 ϕ)Szϕ + 4ex4 cos ϕ sin ϕ Sr z ,
u wire
ϕ = γ r e−iωt/vt , (9.42)
Strictly speaking, this case of a quantum wire of finite length also corresponds to
a quantum dot since dimensional confinement is considered in all three dimensions.
For a quantum wire of infinite length the normalization introduced at the beginning
of this section requires the condition
1/2
1 h̄
γ = , (9.46)
a Mωq
so that
h̄ 1 −iωt/vt
u finite
ϕ
wire
= re . (9.47)
Mωq a
α Eα ωγ
Hdef = √1 Aγ 2 cos q1 x cos hy [an,m (γ ) + an,m (−γ )] eiγ z ,
L n,m,γ cl q1
(9.48)
with
2h̄
A2γ = ,
Mωγ Bγ (9.49)
hd = (n + 12 )π, n = 0, 1, 2, . . . .
In the extreme quantum limit where the carriers are assumed to be confined within
the rectangular boundaries of the quantum wire by an infinitely high potential, the
carrier wavefunction is
1 πx π y ikz
ψk (x, y, z) = |k = √ cos cos e , (9.50)
ad 2a 2d
182 9 Carrier–acoustic-phonon scattering
Eα ωγ2
= √1 A 2 k | cos q1 x cos hy eiγ z |k
L n,m,γ cl q1
6 4
× N + 12 + 12 [an,m (γ ) + an,m (−γ )]N
Eα ωγ2 1
= √1 A 2 δk−k +γ
L n,m,γ cl q1 ad
a d
πx πy
× d x cos2 cos q1 x dy cos2 cos (hy)
−a 2a −d 2d
6 4
× N + 12 + 12 [an,m (γ ) + an,m (−γ )]N
Eα ωγ2 π 2 sin q1 a 1
= √1 A 2 2a2) 2 2 3
L n,m,γ c q
l 1 1 q a(π 2 − q 1 π n + 2 1 − n + 12
1
6 4
× N + 12 + 12 [an,m (γ ) + an,m (−γ )]N
We have used
1 a πx π 2 sin q1 a
d x cos2 cos q1 x = (9.53)
a −a 2a q1 a(π 2 − q12 a 2 )
and
1 d πy π 2 sin hd 1
d
dy cos2
2d
cos hy =
hd(π 2 − h 2 d 2 )
= 2 2 3 .
−d π n + 2 1 − n + 12
1
(9.54)
As usual the plus signs in the phonon-occupancy and momentum-conservation terms
correspond to phonon absorption and the negative signs to phonon emission. Then
the scattering rate predicted by the Fermi golden rule is given by
9.3 Carrier–acoustic-phonon scattering in quantum wires 183
1 L +∞2π {e,a} 2
= dγMn δ(E(k ) − E(k) ± h̄ωγ )
τ 2π −∞ h̄
2 / 02
L +∞ 2π E 1α ωγ2 π 2 sin q1 a
= dγ √ Aγ 2
n,m 2π −∞ h̄ L cl q1 q1 a(π 2 − q12 a 2 )
2
1
× 2 3 n + 12 ± 12
π n + 1 1 − n + 1 2
2 2
h̄ 2 2
×δ (γ ∓ 2kγ ) ± h̄ωγ , (9.55)
2m
where the sum over γ has been converted to an integral, n is the Bose–Einstein
occupation number for the acoustic phonons, and L is the normalization length.
Yu et al. (1994) evaluated this deformation-potential scattering rate numerically for
GaAs quantum wires at 77 K for three different cross sections, 28.3 Å × 56.6 Å,
100 Å × 200 Å, and 50 Å × 200 Å. The results of these numerical calculations
Figure 9.1. Deformation-potential emission and absorption scattering rates for bulk
and confined acoustic phonons in a 28.3 Å × 56.6 Å GaAs quantum wire at 77 K.
Confined modes: –•—•—•–, emission; – – – – – –, absorption. Bulk modes:
–◦—◦—◦–, emission; – – – – – –, absorption. Energy thresholds for the width
modes are at 0.03, 2.36, 2.55, 4.90, 7.30, and 7.40 meV. Energy thresholds for
thickness modes are at 2.06, 4.44, 5.90, 9.87, 14.5, and 15.1 meV. One-dimensional
density-of-states peaks are evident in the emission rate for the quantum wire. From
Yu et al. (1994), American Physical Society, with permission.
184 9 Carrier–acoustic-phonon scattering
are shown in Figures 9.1, 9.2, and 9.3 respectively. In each of these cases the
acoustic phonon emission and absorption rates are shown for the GaAs quantum
wire and for bulk GaAs. The distinct one-dimensional density-of-states peaks are
evident for the case of phonon emission in each of the quantum-wire structures.
These peaks occur at the threshold electron energy for the emission of one of the
various width or thickness modes. The peak magnitudes of the scattering rates
in these quantum wires suggest that carrier–acoustic-phonon scattering via the
deformation potential is significant in quantum-wire elements of nanoscale devices.
The calculations of Yu et al. (1994) are based on the assumption that the wires
have perfectly uniform cross sections. In envisioned practical situations it is unlikely
that the dimensional tolerances associated with the cross-sectional areas will be
small enough to realize subatomic dimensional control. It is thus unlikely that the
measured one-dimensional density-of-states peaks will be as pronounced as they are
in Figures 9.1, 9.2, and 9.3.
Figure 9.2. Deformation-potential emission and absorption scattering rates for bulk
and confined acoustic phonons in a 100 Å × 200 Å GaAs quantum wire at 77 K.
Confined modes: –•—•—•–, emission; – – – – – –, absorption. Bulk modes:
–◦—◦—◦–, emission; – – – – – –, absorption. Energy thresholds for the confined
modes are at 0.03, 0.65, 0.75, 1.39, 2.06, and 2.12 meV. Energy thresholds for the
bulk modes are at 0.59, 1.26, 1.68, 2.80, 4.11, and 4.28 meV. One-dimensional
density-of-states peaks are evident in the emission rate for the quantum wire. From
Yu et al. (1994), American Physical Society, with permission.
9.3 Carrier–acoustic-phonon scattering in quantum wires 185
Figure 9.3. Deformation-potential emission and absorption scattering rates for bulk
and confined acoustic phonons in a 50 Å × 200 Å GaAs quantum wire at 77 K.
Confined modes: –•—•—•–, emission; – – – – – –, absorption. Bulk modes:
–◦—◦—◦–, emission; – – – – – –, absorption. Energy thresholds for the confined
modes are at 0.03, 0.65, 0.75, 1.39, 2.06, and 2.12 meV. Energy thresholds for the
bulk modes are at 0.61, 2.58, 3.05, 5.56, 8.24, and 8.43 meV. One-dimensional
density-of-states peaks are evident in the emission rate for the quantum wire. From
Yu et al. (1994), American Physical Society, with permission.
Chapter 10
Recent developments
Nothing is too wonderful to be true, if it is consistent with the
laws of nature, and in such things as these, experiment is the
best test of such consistency.
Michael Faraday, 1849
186
10.1 Phonon effects in intersubband lasers 187
the dipole pairs BC. The condition that c3+ (ω) = 0 is met for 10 frequencies
and, accordingly, there are 10 interface modes: six of these correspond to GaAs-like
modes with energies in the range of 32 to 37 meV and four are AlAs-like modes with
energies close to 46 meV. The dispersion relations for these 10 modes are shown in
Figure 10.3.
As in Teng et al. (1998), the dielectric constant is taken as
where x represents the content of the Al in Alx Ga1−x As. The phonon energies used
in determining these dispersion relations are given in Table 10.1.
For the structure of Figure 10.2(b), the boundary condition determines that
c4+ /c0+ = 0 and the dispersion relations for the interface modes – as determined
by the previously described iteration procedure – are given by
Figure 10.3. Dispersion relations for the interface LO phonon modes of the
structure depicted in Figure 10.2(a). From Teng et al. (1998), American Institute of
Physics, with permission.
10.1 Phonon effects in intersubband lasers 191
1 2
0= −(3 − 2 )2 (1 + 2 )2 e2qb − 2(32 − 22 )(22 − 12 ) eq(a+b)
1 22 3
+ (32 − 22 )2 (2 − 1 )2 e2q(b−a) + 2(32 − 22 )(22 − 12 ) eq(b−a)
3
− (3 + 2 )2 (2 − 1 )2 e2qa +(3 + 2 )2 (1 + 2 )2 .
(10.10)
The 14 solutions of this equation correspond to six GaAs-like interface modes and
eight AlAs-like interface modes. The potentials for the six GaAs-like modes are
plotted in Figure 10.4 for qa = 0.5.
These LO phonon modes are either symmetric or antisymmetric, as is to be
expected from the symmetry of the heterostructure. They must still be normalized.
The normalization condition of subsection 7.3.4, equation (7.116),
2
1 1 ∂i (ω) ∂ (q, z)
dz q 2 |i (q, z)|2 +
i = h̄ ,
4π 2ω ∂ω ∂z 2ωL 2
i Ri
(10.11)
provides the necessary condition to determine the normalization constant for each
mode. Applying this condition leads to the following normalized LO phonon
potentials: in polar material 1, where i = 2, the confined LO modes are given by
1/2 1/2
4π h̄ 1 1
φ(z) =
L 2 ∂i (ω)/∂ω q 2 + (mπ /a)2
mπ
1/2
cos z, m = 1, 3, 5, . . .
2 a a
× |z| < . (10.12)
a
mπ 2
sin z, m = 2, 4, 6, . . .
a
1/2 ! "1/2
4π h̄ 1 1
φ(z) =
L 2 ∂i (ω)/∂ω q 2 + [2mπ/(b − a)]2
2mπ b+a
1/2 cos z + m = 1, 3, 5, . . .
4 b−a 4
×
b−a
2mπ b+a
sin z+ m = 2, 4, 6, . . .
b−a 4
b a
− <z<− . (10.13)
2 2
For the half-space LO modes in polar material 3, where i = 0, 3,
1/2 1/2
4π h̄ 1 1
φ(z) =
L 2 ∂i (ω)/∂ω q 2 + (2mπ /L)2
2mπ b b
1/2 sin z+ m = 1, 2, 3, . . . z≤−
4 L 2 2
×
L
$ %
sin 2mπ z − a m = 1, 2, 3, . . . z≥ .
a
L 2 2
(10.14)
Figure 10.4. Potential profiles for the six GaAs-like interface LO phonon modes for
the structure of Figure 10.2(b). The mode frequencies are (1) 35.86 meV, (2) 34.91
meV, (3) 34.26 meV, (4) 34.01 meV, (5) 33.17 meV, and (6) 33.07 meV. From Teng
et al. (1998), American Institute of Physics, with permission.
10.1 Phonon effects in intersubband lasers 193
As in subsection 7.3.4, the interface LO phonon mode potentials are of the form,
equations (7.117), (7.118),
i (q, z) = A(ci− e−qz + ci+
e+qz ) = Ai (q, z), (10.15)
with
6 4
{e,a}
Mn,n (q) = n , k , Nq + 1
2 ± 12 Hj n, k, Nq + 1
2 ± 1
2 , (10.20)
where the initial and final energies each have been written as a sum of the in-plane
and subband energies, E(k) + E n and E(k ) + E n respectively; E(k) = h̄ 2 k 2 /2m.
Since the in-plane energy associated with the two-dimensional wavevector k has
been separated from the subband energy, the subscript 2D of Section 8.1 has been
omitted. Phonon-assisted transitions between the initial and final carrier states are
depicted in Figure 10.5.
Teng et al. (1998) solved the Schrödinger equation numerically to determine
the electron wavefunctions, for selected phonon modes, needed to evaluate the
matrix elements for phonon-assisted transitions. These have focused on the emission
process to gain insights into phonon-assisted processes of importance in narrow-well
semiconductor lasers. As expected (Stroscio, 1996), their results show that the
half-space modes and confined LO modes yield phonon-assisted rates that are small
194 10 Recent developments
50 meV. As is apparent from Figure 10.6, the rate is substantially larger than the rate
associated with the emission of a bulk phonon. Clearly, in the design of narrow-well
semiconductor lasers it is essential that the interface modes be considered. The usual
approximation of treating the phonons as bulk phonons is inadequate.
Figure 10.6. For the symmetric structure of Figure 10.2(b), the maximum transition
rates for the antisymmetric interface mode and the bulk mode are depicted as
functions of the difference E 2 − E 1 between the final and initial energies.
196 10 Recent developments
are modeled (Kisin et al., 1997) with a four-band Kane model. The electronic
wavefunctions (Gorfinkel et al., 1996) for the states in subband n are taken to be
(n) 1
k (ρρ , z) = √ eik·ρρ ψk(n) (z), (10.21)
Sa
where, as defined previously, ρ = (x, y) and S is the area of the heterostructure. For
intersubband lasers – such as the quantum-cascade laser – involving only transitions
(n)
in the conduction band, the four-component Kane wavefunction envelopes, ψk ,
may be described simply in terms of scalar wavefunctions. For the lowest subband,
n = 1,
√ cos k z1 z |z| < a/2,
ψ1 (z) = 2C1 (10.22)
−λ
cos(k z1 a/2) e 1 (|z|−a/2) |z| > a/2,
with
with
Here E wn and E bn are the energies for the nth subband of the well and the barrier
measured with respect to the valence band in each material. For these approximate
wavefunctions, it is clear that phonon-assisted transitions from the second to the low-
est subband will involve only the antisymmetric interface phonons, the odd-parity
confined phonon modes, and barrier modes. However, intrasubband phonon-assisted
transitions will include contributions from the symmetric interface phonons, the
even-parity confined phonon modes, and barrier modes. As will become apparent,
such phonon-assisted processes have line-broadening effects that exert a major
influence on the gain of the laser. Moreover, it will become apparent that the
energy dependence of the line broadening resulting from the energy spectrum of
the phonons also plays a significant role in determining the properties of the gain of
the laser.
Subband energy dispersion curves for 60-ångstrom-wide and 100-ångstrom-wide
quantum wells are depicted in Figure 10.7 along with examples of intersubband and
intrasubband phonon-assisted processes; in Figure 10.7, x = 0.3 and the conduction
band and valence band off-sets are c = 300 meV and v = 150 meV respectively.
10.2 Effect of confined phonons on gain of intersubband lasers 197
The phonon spectrum used in the calculation of the intersubband optical gain
in the quantum-well system under consideration includes the symmetric interface
modes, the antisymmetric interface modes, the two lowest-order confined modes,
and the half-space modes, which are referred to as the barrier modes in this
section. The Fröhlich interaction Hamiltonians used here are those of the dielectric
continuum model with electrostatic boundary conditions, as discussed in Chapter 7
and in Appendix C.
These modes may be obtained straightforwardly from the modes for the double-
heterojunction, uniaxial structure of Appendix D by taking the limit where both
(ω)⊥ and (ω) are equal to (ω). Figures 10.8(a), (b) and 10.9(a), (b) depict
the contributing scattering rates (Kisin et al., 1997) for the scalar wavefunctions
presented previously and for the phonon modes of the dielectric continuum model
with electrostatic boundary conditions.
In calculating these results, the parameters have been selected as follows: x =
0.4, c = 300 meV, v = 200 meV, and a = 60 Å. The energy spectrum of Figure
10.8(a) illustrates clearly that there are several distinct thresholds.
Figure 10.7. Energy dispersion curves for 60-ångstrom-wide (solid lines) and
100-ångstrom-wide (broken lines) Al0.3 Ga0.7 As/GaAs/Al0.3 Ga0.7 As quantum
wells. Typical intersubband and intrasubband transitions are shown for both quantum
wells. The energy gap for the GaAs well, E g (GaAs), is taken as 1.4 eV and the ratio
of the effective mass to the electron mass is taken as 0.067 for GaAs. From Kisin
et al. (1997), American Institute of Physics, with permission.
198 10 Recent developments
Figure 10.8. (a) Intrasubband scattering rates in the second subband as a function of
the electron energy in the second subband. The high- and low-frequency symmetric
interface LO phonon modes, the confined modes, and the barrier modes are labeled
by is +, is −, c and b respectively. (b) Intersubband scattering rates from the second
to the lowest subband as a function of the electron energy in the second subband.
The high- and low-frequency antisymmetric interface LO phonon modes, the
confined modes, and the barrier modes are labeled by ia+, ia−, c and b respectively.
From Kisin et al. (1997), American Institute of Physics, with permission.
10.2 Effect of confined phonons on gain of intersubband lasers 199
Figure 10.9. (a) Intrasubband scattering rates in the lowest subband as a function of
well thickness. The initial electron energy in the first subband is 60 meV. (b)
Intersubband scattering rates for transitions from the second to the first subband as a
function of well thickness. The initial electron energy in the second subband is
10 meV. From Kisin et al. (1997), American Institute of Physics, with permission.
200 10 Recent developments
The optical gain spectrum g() for intersubband transitions between the second
and the lowest subband (Gelmont et al., 1996; Gorfinkel et al., 1996) is given by
4πn 2 e2 |z 12 |2 ∞ −ε/k B Te f1
g() = √ dε τε () e 1− , (10.26)
h̄ac (∞)k B Te 0 f2
where the line-shape function is defined by
W (ε)
τε () = , (10.27)
[ − ε ]2 + [W (ε)]2
the transition energy is related to the energy difference h̄0 between the subbands
at k = 0 by
Figure 10.10. Optical gain in cm−1 for different rates of electron escape, τout , from
the lowest subband: (a) τout is 0.4 ps and 0.6 ps; (b) τout is 0.55 ps. The broken lines
are calculated assuming that only bulk GaAs LO phonons contribute; the
broken-and-dotted lines assuming that only bulk GaAlAs phonons contribute. The
solid lines are calculated from the full spectrum of dimensionally confined phonon
modes. From Kisin et al. (1997), American Institute of Physics, with permission.
202 10 Recent developments
Figure 10.11. Optical gain in cm−1 for two different sets of phonon modes: broken
lines, all phonon modes are included; solid lines, all modes but the barrier mode are
included. Results are given for three different values of the electron escape rate, τout :
(a) 0.4 ps; (b) 0.5 ps; (c) 0.6 ps. From Kisin et al. (1997), American Institute of
Physics, with permission.
10.3 Phonon contribution to valley current in double-barrier structures 203
Figure 10.13. Magnetic field versus applied voltage: (a) full-field range; (b)
low-field range. The heterostructure is a double-barrier structure with an
80-ångstrom-wide GaAs quantum well having a 33-ångstrom-wide AlAs emitter
barrier and a 45-ångstrom-wide AlAs collector barrier. From Turley et al. (1993),
American Physical Society, with permission.
206 10 Recent developments
(Faist et al., 1996a, b; Gmachl et al., 1998; Stroscio, 1996; Stroscio et al., 1999).
This fast depopulation promotes the lasing in intersubband lasers. Phonon-assisted
depopulation rates can be enhanced in asymmetric double-quantum-well het-
erostructures if the subband being depopulated is degenerate with a second subband
in the wider quantum well, as shown in Figure 10.19 (Stroscio et al., 1999).
The double-well active region of the AlAs/GaAs heterostructure depicted in this
figure is designed so that the lower carrier state A1 of well A is depopulated
by phonon-assisted transitions to state B1 of well B. As will be demonstrated,
these phonon-assisted transitions are enhanced greatly when levels A1 and B1 are
degenerate.
The interwell transition rate between states A1 and B1 is enhanced when A1
and B2 are nearly degenerate, as is illustrated by Figures 10.20 and 10.21. This
effect is associated primarily with an enhancement in the overlap between the
final and initial state wavefunctions involved in the phonon-assisted transition. In
Figure 10.20 the interwell transition rate is shown as a function of a1 for three
different models of the optical phonon modes participating in the phonon-assisted
transitions. The broken curve (a) is calculated with the Fermi golden rule, assuming
that only the bulk GaAs phonon with an energy of 36 meV participates in the
phonon-assisted transitions between states A1 and B1. Likewise, the broken curve
(b) is calculated with the Fermi golden rule, assuming that only the bulk AlAs
Figure 10.14. Magnetic field versus applied voltage: reverse bias diagram. The
heterostructure is a double-barrier structure with an 80-ångstrom-wide GaAs
quantum well having a 33-ångstrom-wide AlAs emitter barrier and a
45-ångstrom-wide AlAs collector barrier. From Turley et al. (1993), American
Physical Society, with permission.
10.4 Population inversion in asymmetric double-barrier quantum-well lasers 207
Figure 10.15. Band structure of the device studied by Choi et al. (1990) under a
bias V E . The energy distribution of the current, the so-called hot-electron
distribution ρ(E), is shown in the barrier region B. The current emerging from the
emitter E is I E and the current at the collector C is indicated by IC . The states E n
(n = 1, 2, 3, 4) are resonant states created in the barrier. E F is the Fermi energy and
E P is the energy of the emitted phonon. U and L are the effective ‘upper’ and
‘lower’ contributions to the half-width of the energy distribution after phonon
emission. From Choi et al. (1990), American Physical Society, with permission.
208 10 Recent developments
It has been recognized for a number of years that the dimensional confinement of
carriers in thin film supeconductors leads to modification of the superconducting
energy gap and the critical temperature Tc (Blatt and Thompson, 1963; Thompson
and Blatt, 1963; Yu et al., 1976). In particular, it has been understood that size-
Figure 10.16. Transfer ratio α (smooth curve) and its derivative (peaked curve) as
functions of V E : (a) experimental values and (b) theoretical values. The subsidiary
peaks at V P1 and V P2 are due to phonon emission. The arrows show the current
directions. From Choi et al. (1990), American Physical Society, with permission.
10.5 Confined-phonon effects in thin film superconductors 209
quantization effects are manifest, as a function of the film thickness d, each time
one of the two-dimensional subband energy levels E n (d) passes through the Fermi
level as the film thickness is varied. Moreover, the Bardeen–Cooper–Schrieffer
(BCS) theory of superconductivity is based on the formation of Cooper pairs whose
binding energy is determined by the phonon-mediated pairing of electrons. Thus,
it is reasonable to expect that the effects of phonon confinement play a role in
determining the characteristics of thin film superconductors. Hwang et al. (2000)
have recently investigated this possibility within the context of the BCS theory for
thin films, by replacing the bulk phonon by the spectrum of confined phonons for the
thin film. These results of Hwang et al. are expected to reveal selective qualitative
features associated with phonon confinement in thin film superconductors but
are not complete enough to give rigorous qualitative predictions. Only selected
qualitative features are expected, since the calculations assume infinitely high
electronic barriers at the interfaces of the thin film and since the interface phonon
modes are not included in the spectrum of phonon modes. Figure 10.22 shows the
Figure 10.17. (a) Theoretical and (b) experimental electron energy distribution
ρ(E) for the heterostructure described previously and shown in the inset. The arrow
in the inset shows the current direction. The theoretical electron energy distributions
ρ(E) are shown at 0 Å, 600 Å, and 1500 Å from the point of injection. The
experimental curve for ρ(E) was obtained at Ve = −0.3 V. From Choi et al. (1996),
American Institute of Physics, with permission.
210 10 Recent developments
Figure 10.18. Experimental results for the photocurrent transfer ratio α are shown
as a solid curve. The theoretical results for plasmon emission, phonon emission, and
both phonon and plasmon emission are designated by triangles, crosses, and circles
respectively. The difference between the theoretical and experimental contributions
at high Ve is due to the non-linear relation between the injection energy and Ve .
From Choi et al. (1996), American Institute of Physics, with permission.
10.5 Confined-phonon effects in thin film superconductors 211
Figure 10.21. Peak values of the interwell phonon-assisted transition rates for
conditions of near-degeneracy of energy levels A1 and B2 for a series of values of
the width a3 of the wider well. The curves corresponding to a3 = 10 nm show the
relative contributions of the confined phonons (broken curve) and interface phonons
(dotted curve). The bold broken line with a minimum at about 8.5 nm is the rate of
non-radiative intrawell intersubband transitions, 1/τ12 , for the case where
a3 = 18 nm. From Stroscio et al. (1999), American Institute of Physics, with
permission.
c,v
Hdef = E c,v (a) = E 1c,v ∇ · u, (10.33)
These modes are normalized according to the procedures of Section 5.1 but in terms
of wn (q , z) instead of un (q , z), since the considerations of Appendix A make it
√
clear that it is convenient to use wn = ρun ; un (q , z) ≡ un,q is obtained from
the mode amplitudes of subsection 7.6.2 by dividing by exp i(q · r − ωq t) ,
since this factor is included separately in the above equation for u(r) and in
the energy-conserving delta function in the Fermi golden rule. Clearly, transverse
modes do not contribute to the deformation potential. Two of the types of mode
in Section 7.2 contain longitudinal components: the dilatational modes and the
flexural modes. The dilatational modes are irrotational and they are associated
with compressional distortions of the medium; the compressional character of
these modes leads to local changes in the volume of the medium. As discussed
in subsection 7.6.2, Wendler and Grigoryan (1988) derived the localized acoustic
modes for an embedded quantum well. For a symmetric quantum well, the electrons
couple via the deformation potential to the symmetric shear vertical (SSV) confined
acoustic modes. The quantum-well heterostructure and the lowest-order SSV mode
are depicted in Figures 10.23(a), (b).
The electronic wavefunction for the lowest two-dimensional subband of the layer
is given by
1
k (ρ, z) = √ exp(ik · ρ )χ(z), (10.35)
S
where k is the two-dimensional wavevector and, as defined previously, ρ ≡ (x, y).
Only this lowest electronic subband is taken to be occupied. Then, the probability
of transitions between electronic states k and k due to the emission or absorption
of a confined phonon of band n and wavevector q is given by
2π $ %
P (±) (k , k n, q = |M(q)|2 n n,q + 12 ± 12 δk x ∓q,k x δk y ,k y
h̄
$ %
× δ E(k ) − E(k ) ∓ h̄ωn,q
2 3
× F(k ) 1 − F(k ) , (10.36)
where
∞
Ec
M(q) = el 1 ∇ · un,q χ 2 (z)dz, (10.37)
κ (q) −∞
n n,q is the phonon occupation number of mode n, q , κ el (q) is the electron
permittivity described in Appendix F (Bastard, 1988; Komirenko et al., 2000b),
F(k ) = F(k x , k y ) is the electron distribution function, and q = q . Appendix
F provides a derivation of κ el (q) based on the Lindhart method as applied to a
two-dimensional electron gas (Bastard, 1988).
To determine the net rate of generation of localized acoustic phonons, Komirenko
et al. (2000b) considered
dn n,q (+) (−)
= γn,q
(1 + n n,q ) − γn,q n
n,q
− βn,q n n,q , (10.38)
dt
(±)
where the γn,q are determined by calculating the total rates of absorption and
emission of phonons for mode n, q and βn,q represents phonon losses such
as non-electronic phonon absorption or anharmonic phonon decay. Defining the
phonon increment by
(+) (−)
γn,q ≡ γn,q
− γn,q
, (10.39)
it follows that
m∗ 2 3
γn,q = |M(q)|2 I (+) (q) − I (−) (q) , (10.40)
π h̄ 3 q
where m ∗ is the effective mass and
∞
(±)
m ∗ ωn,q 1
I (q) = dk y F (sign q) ± q, k y . (10.41)
−∞ h̄|q| 2
216 10 Recent developments
For the case where an external electric field causes the electrons to drift with
a velocity Vdr , taking the electron distribution function to be a drifted Fermi
distribution FF , we have
m∗
F(k x , k y ) = FF k x − Vdr , k y . (10.42)
h̄
It then follows that αn,q = γn,q /Vg > 0 if the electron drift velocity exceeds the
confined phonon phase velocity, Vdr > ωn,q /|q|, where the phonon group velocity
Vg = dωn,q /dq. This condition is the same as the criterion for Cerenkov emission.
Komirenko et al. (2000b) evaluated αn,q for a p-doped, Si/SiGe/Si quantum well 10
nm thick, for the two lowest SSV phonon branches. The amplification coefficients
for these phonon branches are shown in Figure 10.24 for temperatures T of 50, 100,
150, and 200 K. In these numerical results, the hole density is taken as 1012 cm−2 ,
the drift velocity is given by Vdr = 2.5 ct1 and ct1 = 3.4 × 105 cm s−1 for the SiGe
layer.
The amplification coefficient for this p-doped Si/SiGe/Si quantum well is seen
to be of the order of tens to hundreds of cm−1 for confined modes in the sub-THz
frequency range.
Chapter 11
Concluding considerations
Now there is one outstandingly important fact regarding
Spaceship Earth, and that is that no instruction book came with
it.
R. Buckminster Fuller, Operating Manual for Spaceship Earth, 1969
218
11.2 Future trends: phonon effects in nanostructures and phonon engineering 219
and in many cases dominate over all other scattering mechanisms. It is well
known that these relatively large carrier–phonon interactions play a major role
in determining carrier mobilities (Ferry, 1991; Mitin et al., 1999). In Chapter
10, phonons in dimensionally confined structures were seen to make a dominant
contribution to the valley current in specific double-barrier quantum-well structures
(Turley et al., 1993) and the properties of thin film superconductors were shown
to depend on the spectrum of confined phonons for sufficiently thin films (Hwang
et al., 2000).
Moreover, it was shown in Chapter 10 that dimensional confinement of phonons
in intersubband semiconductor lasers changes the laser gain (Kisin et al., 1997,
1998a, b) and leads to enhanced population inversions in some asymmetric double-
barrier quantum-well lasers (Stroscio et al., 1999). Indeed, Educato et al. (1993),
Julien et al. (1995), Wang et al. (1996a, b), and Gauthier-Lafaye et al. (1997)
have examined optically pumped intersubband scattering in coupled quantum-well
lasers. These same authors have shown that interface–phonon-assisted transitions
are important in such structures. Such results illustrate the importance of phonon
confinement effects in intersubband lasers. As discussed in Chapter 10, the proper
treatment of optical phonon confinement in optical systems such as intersubband
semiconductor lasers depends critically on the detailed energy spectrum of the
phonons. This is also true for a number of novel current-injection semiconductor
intersubband lasers (Sun et al., 1993; Zhang et al., 1996; Sung et al., 1996; Faist
et al., 1996a, b), as is illustrated by the calculations of Stroscio (1996) and Kisin
et al. (1997, 1998a, b).
effects are likely to represent one of the future trends in in nanostructures as well
as in phonon engineering. Pötz and Shroeder (1999) pointed to the potential for
achieving coherent control in atoms, molecules, and semiconductors.
Phonon engineering in nanostructures is likely to be a future avenue for
device-related research and development. Indeed, the ability to model the phonon
modes in dimensionally confined structures has been the basis for efforts to design
nanostructures such that the resulting carrier and phonon states are tailored to yield
dissipative and scattering mechanisms different from those of the corresponding
bulk structures. Sakaki (1989) analyzed theoretically the electronic structure of
quantum-wire superlattices and coupled quantum-box arrays with the purpose of
engineering the electron energy subbands in such a way that carrier–optical-phonon
scattering is suppressed. As another example, the use of metal–semiconductor
heterointerfaces to reduce carrier–interface–phonon scattering rates (Stroscio et al.,
1992) represented one of the first approaches to phonon engineering in nanos-
tructures. Indeed, the first calculations of carrier–LO-phonon scattering rates in
polar-semiconductor quantum wires (Stroscio et al., 1990; Kim et al., 1991) revealed
that the scattering of carriers with interface optical phonon modes dominates over
the scattering associated with confined and half-space phonons. As a means of
reducing the unwanted inelastic scattering caused by the interface LO phonons
in semiconductor quantum wires, it was suggested (Stroscio et al., 1992) that the
large carrier–LO-phonon scattering caused by interface phonons could be eliminated
by the encapsulation of the quantum wire in a metal. In a related example of
phonon engineering, Leburton (1997) modeled dissipation- and scattering-time
engineering in heterostructure-based quantum devices. Indeed, Leburton and his
collaborators – Educato et al. (1993), Julien et al. (1995), Wang et al. (1996a, b),
and Gauthier-Lafaye et al. (1997) – have exploited such effects in optically pumped
intersubband scattering in coupled quantum-well lasers. Clearly, efforts in phonon
engineering will be one of the future research trends based on the theories of
confined phonons presented in this book.
Appendices
Huang (1951) and Born and Huang (1954) took the most general form of the
microscopic theory of diatomic polar crystals to be described by a pair of equations
relating w, E, and P,
Born and Huang showed that d equals b as a result of energy conservation. In (A.1),
√
w = µN /V u; u, µ, N , and V are the relative displacement of the two ions, the
reduced mass of the ion pair, the number of unit cells in the crystal, and the volume
of the crystal respectively. Assuming a time dependence e−iωt and eliminating w
from these equations yields the relation
b2
P= c+ E. (A.2)
−a − ω2
4π b2
(ω) = 1 + 4π c + . (A.3)
−a − ω2
221
222 Appendices
ωLO
2 − ω2 2 − ω2 ) + (ω2 − ω2 )]
[(ωTO LO TO
(ω) = (∞) = (∞)
ωTO
2 − ω2 ωTO
2 − ω2
ωLO
2 /ω2 − 1
TO
= (∞) + (∞)
1 − ω2 /ωTO
2
(0) − (∞)
= (∞) + (∞) . (A.4)
1 − ω2 /ωTO
2
Comparing the relation (A.3) of Born and Huang with that based in the Lyddane–
Sachs–Teller equation, it is evident that
1/2
(∞) − 1 (0) − (∞)
c= , a = −ωTO
2
, and b= ωTO .
4π 4π
(A.5)
Thus, the pair of equations in (A.1) and (A.2) put forth by Huang and Born may
be written as
1/2
(0) − (∞)
(ωTO
2
− ω )w =
2
ωTO E (A.6)
4π
and
1/2
(0) − (∞) (∞) − 1
P= ωTO w + E. (A.7)
4π 4π
and
1/2
µN (∞) − 1
P= (0) − (∞)ωTO u + E. (A.9)
4π V 4π
equation and through the effective charge, en∗ : for a binary medium labeled by
subscript n,
where (total (r, t) includes both bulk and surface polarization charge density; using
the Green’s function approach Wendler took a solution to Poisson’s equation of the
form
∂ 1
(r, t) = − d 3 r P (r , t).
| β
(B.5)
β
∂r β |r − r
Then, with E α (r, t) = −∂(r, t)/∂rα and with the Lorentz equation (B.3), it
follows that
4π
E αlocal (r, t) = (λn − λ0n )Pα (r, t)
3
+ 4π d 3 r αβ (r − r )Pβ (r , t), (B.6)
β
1 ∂2 1
αβ (r − r ) = . (B.7)
4π ∂rα ∂rβ |r − r |
where
λn = 4πω2 /ωplasma,n
2
, λ0n = 4π ω0n
2
/ωplasma,n
2
, (B.9)
∗2
ωplasma,n,a(b)
2
= 4π nn en,a(b) /µn,a(b) . (B.10)
By defining a two-dimensional Fourier transform for the electrostatic fields of the
form
∞
1
P(r) = d 2 q eiq ·r P(q , r3 ), (B.11)
2π −∞
with r in the plane of the heterointerface of the bilayer system, Wendler showed
that
Elocal (q , r3 ) = n P(q , r3 ), (B.12)
where
λn − λ0n
n = (B.13)
1 − nn αn (λn − λ0n )
and
1/2
θn
u(q , r3 ) = P(q , r3 ), (B.14)
nn en∗
with
1/2 1
θn = . (B.15)
1 − nn αn (λn − λ0n )
Moreover, the Green’s function approach of Wendler yields several useful
conditions:
nn αn
n (∞) = 1 + 4π ,
1 − 43 π nn αn
1
ωLO,n
2
= ω0,n
2
+ 23 ωplasma,n
2
,
1+ 3 π nn αn
8
1
ωTO,n
2
= ω0,n
2
+ 13 ωplasma,n
2
,
1− 3 π nn αn
4
1 1
ωLO,n
2
− ωTO,n
2
= 23 ωplasma,n
2
+ 13 ωplasma,n
2
1+ 3 π nn αn
8
1− 3 π nn αn
4
ωplasma,n
2
= .
(1 + 83 π nn αn )(1 − 43 π nn αn )
(B.16)
where n represents either material 1 or material 2. As may be verified alge-
braically, the Lyddane–Sachs–Teller relations of subsection 2.3.3 are satisfied by
the frequencies in (B.16) (Wendler, 1985). Wendler’s treatment of the effect of
dielectric polarizability provides the basis for many works on optical phonons in
dimensionally confined systems. The utility of Wendler’s solution is illustrated in
Chapter 7.
Appendices 225
where
1/2
1
γ = (∞)(ωLO − ωTO )
2 2
4π
! "1/2
1
= [(0) − (∞)] ωTO
4π
! "1/2
(∞) 1 1
= ωLO 4π − (C.2)
4π (∞) (0)
√
and where the displacement is described in terms of w = ρu for the same reasons
as in Appendix A. As will become apparent, it is convenient to define
! "1/2
1 4π 1 1
= γ = ωLO 4π − . (C.3)
β (∞) (∞) (0)
In the Lagrangian density (C.1), the first term is the kinetic energy, the second
term is the potential energy of the lattice due to short-range forces, the third term
226 Appendices
is due to the potential energy of the macroscopic electric field in the absence of
ionic motion, the fourth term is the potential energy associated with the coupling
of the lattice to the macroscopic electric field, and the fifth term represents the
quadratic dispersion associated with the short-range forces between ions, for the
case of isotropic dispersion, Z i jkl = Aδi j δkl + Bδik δ jl + Cδil δ jk .
The Euler–Lagrange equations derived by Nash from the variations of w and
to describe the optical phonon modes in a slab situated between z = 0 and z = d
are
∂ 2w
+ ωTO
2
w + γ ∇ + A∇ 2 w + (B + C)∇(∇ · w) = 0 (C.4)
∂t 2
and
respectively. In these equations, γ is as given in (C.2) and f (z) = )(z) − )(d − z),
) being the Heaviside step function. Following Nash, w is written as the gradient
of a mechanical potential χ̄, w = −∇ χ̄ . The second of these equations is Poisson’s
equation,
or
1
− ∇ · [(∞)∇]
(∞)
4π γ ∂ χ̄ ∂ χ̄
= ∇ χ̄ −
2
δ(z − d + η) + δ(z − η) , (C.7)
(∞) ∂z z=d ∂z z=0
where the last two terms on the right-hand side of (C.7) arise from the surface
polarization charge due to the discontinuity of the z-component of w. In these
results, w is taken to be normalized as described previously. That is, these modes
are normalized (according to the procedures of Section 5.1) in terms of wm (q , z)
instead of um (q , z), since the considerations of Appendix A make it clear that it is
√
convenient to use wm = ρum :
h̄
um (r) = (a
2 ρω (q ) m,q
†
+ am,−q )wm (q , z) eiq ·r . (C.8)
q ,m 2L m
Here the subscript m is retained to remind us that this result holds for a general
medium m. This last result is, of course, consistent with the normalization condition
of subsection 7.3.1, (7.44),
√ √ h̄
L 2 [ µn um (q, z)]∗ · [ µn um (q, z)] dz = ; (C.9)
2ωm (q)
Appendices 227
(ωLO
2
− ω2 )∇ 2 χ̄ + µ∇ 2 ∇ 2 χ̄ = 0 (C.11)
and
1 2
= (ω − ω2 )χ̄ + µ∇ 2 χ̄ − B0 , (C.12)
γ TO
for 0 < z < d, with µ = A + B + C and B0 a constant. For solutions outside the
slab, satisfies Poisson’s equation. As discussed previously, the fields for a slab
exhibit translational invariance in the direction normal to the surface of the slab;
accordingly, χ̄(r) = χ(z) eiq·ρρ and (r) = φ(z) eiq·ρρ , where q and ρ are as defined
previously. Thus, for q = 0 it follows that B0 = 0. Then the last two equations may
be written as
d2 2 − ω2
ωLO d2
µ + − q 2
− q 2
χ(z) = 0 (C.13)
dz 2 µ dz 2
and
1 2 d 2 χ(z)
φ(z) = (ωTO − ω2 − µq 2 )χ(z) + µ , (C.14)
γ dz 2
respectively. Nash has applied the methods used in analyzing the Sturm–Liouville
equation to show that the normal-mode solutions for the slab must satisfy the
following boundary conditions:
= 0 at z = ±∞ (C.15)
and
χ = χ = 0 and χ = 0 (C.17)
respectively.
228 Appendices
(ωLO − ω )/(µ − q ) in the fourth-order equation for χ(z) to define the fourth-order
2 2 2
eigenvalue equation
2 2
d d
+ k 2
nq − q 2
χnq (z) = 0, (C.18)
dz 2 dz 2
where knq2 is the eigenvalue of χ (z) and where the subscripts nq identify different
nq
members of the basis set. As usual, q is the wavevector in the x y-plane and n is
here an integer used to distinguish different modes having the same value of q. It
is this equation that Nash uses to generate complete sets of orthonormal functions
representing the phonon modes. As discussed in Chapter 7, the orthogonality
relation for the normal modes is
d 3 r wα (r) · wβ (r) = 0 for α = β, (C.19)
where α and β each denote the set of quantum numbers nq. Recalling that
w = −∇ χ̄ and that χ̄(r) = χ(z) eiq·ρρ this orthogonality relation may be written
as
d ∗
∗ 1 ∗
dχnq dχn q
χnq · χn q = dz q 2 χnq (z) · χn q (z) + . (C.20)
2d 0 dz dz
The various sets of confined and interface optical phonon modes, which have
been discussed in the literature include the so-called slab modes that satisfy the
electrostatic boundary conditions, the Huang–Zhu modes, which are based on a
reformulation of the slab modes, and the so-called guided modes, which satisfy
mechanical boundary conditions. For the slab modes,
(n + 1)π z
χnq (z) = sin ; (C.21)
d
the boundary conditions on χnq (z) at z = 0 and d are χnq = χnq = 0 and it is
necessary to include the interface modes to obtain a complete set of modes. The
electrostatic potentials for these modes, nq , are given by
4π γ
nq = − χnq (z). (C.22)
(∞)
The normal modes satisfying = 0 at z = ±∞ and χ = χ = 0 at z = 0 and d
are the slab modes.
For the modified Huang–Zhu modes given by Nash,
cos µnq π z /d + Dnq cosh qz n even,
χnq (z) = (C.23)
sin µnq π z /d + Dnq sinh qz n odd;
Appendices 229
the boundary conditions on χnq (z) at z = 0 and d are χnq = χnq = 0 and it is
necessary to include the interface modes to obtain a complete set of modes. The
Huang–Zhu (H–Z) modes have been defined to satisfy these boundary conditions
at the interfaces since the available microscopic calculations exhibit this behavior.
In practice, the boundary conditions χnq = χnq = 0 are satisfied by selecting the
necessary values of Dnq and µnq . The H–Z modes are referred to frequently as
the reformulated slab modes and also as the reformulated modes of the dielectric
continuum model (with electromagnetic boundary conditions). Huang and Zhu
(1988) showed that the modes of the dispersionless microscopic theory at small q
are approximated well by these reformulated slab modes. The electrostatic potentials
for these modes, nq , are given by
4πγ
nq = − χnq (z). (C.24)
(∞)
For the guided modes,
nπ z
χnq (z) = cos ; (C.25)
d
the boundary conditions on χnq (z) at z = 0 and d are again χnq = χ = 0 and, as
nq
pointed out by Nash, it is not necessary to include the interface modes to obtain a
complete set of modes. The electrostatic potentials for these modes, nq , are given
by
4πγ cos nπ z/d − e−qd/2 cosh q(z − d/2) n even,
nq = −
(∞) cos nπ z/d + e−qd/2 sinh q(z − d/2) n odd.
(C.26)
While these guided modes satisfy χnq = 0 at the interfaces, in agreement with the
Clearly, of the three commonly used sets of phonon modes, only the slab modes
satisfy the conditions that = 0 at z = ±∞ and χnq = χnq = 0 at z = 0 and
d. The alternative set of boundary conditions, = 0 at z = ±∞ and χnq = 0 at
z = 0 and d, is not satisfied by any of the commonly used sets of phonon modes. As
argued by Nash (1992), the Euler–Lagrange equations, (C.4) and (C.5) are satisfied
in each material layer and provide the basis for determining the so-called connection
rules at each heterointerface. In the dispersionless limit the Lagrangian density does
not contain terms with spatial derivatives of w and, accordingly, the significance of
mechanical boundary conditions is unclear in this case.
For this reason, Nash examined the Euler–Lagrange equations at the heteroin-
terfaces for the case where the modes have non-zero dispersion, as described by
the Lagrangian density of (C.1). In this case, the connection rules may be seen
to be that w/ρ 1/2 , , and the z-components of the electric displacement vectors,
Dz = (1/4π)(∞) − γ wz , are continuous at the heterointerfaces. Thus in
the dispersive continuum model there is no contradiction in applying mechanical
boundary conditions to w and w and at the same time applying electromagnetic
boundary conditions to and D. Indeed, both types of boundary condition must be
invoked to derive the full set of normal modes.
As an illustration leading to an appreciation of the importance of Nash’s con-
tribution to the understanding of the boundary conditions for the system at hand,
it is instructive to oversimplify the current analysis artificially by considering the
equation
(∞)∇ − 4π γ w = 0. (C.29)
The left-hand side of this equation appears in the Euler–Lagrange equation obtained
from variation of in the Lagrangian density: ∇ · [(∞)∇ − 4πγ w] = 0. Clearly,
this simplified equation is not sufficient to describe the system fully but it is in
fact the equation assumed in the mechanical models to derive longitudinal modes
for the construction of normal modes. Let us consider (C.29) and show that it in
fact gives the correct Fröhlich interaction Hamiltonian for electron–polar-optical-
phonon interactions, as derived in Section 5.1. Substituting the expression for γ in
terms of ωLO , (0), and (∞), it follows immediately that
! "1/2
1 1
∇ = ωLO 4π − w. (C.30)
(∞) (0)
√
Then using the expression of Appendix A, w = µN /V u, and the normalized
displacement of Section 5.1 for the case where ωq = ωLO and only the longitudinal
polarization contributes, it follows immediately that
Appendices 231
! "1/2
1 1
∇ = ωLO 4π −
(∞) (0)
N 1 h̄ iq·r
× aq e + aq† e−iq·r
V N q 2ωLO
! "1/2
2π h̄ωLO 1 1
= ωLO 4π −
V (∞) (0)
† −iq·r
× aq e iq·r
+ aq e . (C.31)
q
Then, using the result that ∇ = −iq and multiplying by −e to obtain the
Hamiltonian, it follows that
2πe2 h̄ωLO 1 1 1 iq·r
HFr = −i − aq e − aq† e−iq·r ,
V (∞) (0) q q
(C.32)
which is precisely the result obtained in Section 5.2. Thus, the simplified result,
(∞)∇ − 4πγ w = 0, gives the correct three-dimensional Fröhlich interac-
tion Hamiltonian. However, this successful derivation does not justify the use of
(∞)∇ − 4πγ w = 0 instead of ∇ · [(∞)∇ − 4πγ w] = 0 in analyzing the
boundary conditions for the phonon modes in a slab. Indeed, as argued by Nash
(1992) the longitudinal waves obtained from this simplified equation do not provide
sufficient degrees of freedom for the mechanical and electromagnetic boundary
conditions to be satisfied simultaneously.
The understanding gained through Nash’s paper (1992) has been critical in
sorting out the correct approaches for calculating carrier–optical-phonon scattering
rates in dimensionally confined structures. Of course, the key feature of such cal-
culations is selecting a complete, orthogonal set of phonon modes. Notwithstanding
the situation that only slab modes satisfy the desired boundary conditions, Nash
demonstrated by explicit calculations that any of the three sets of complete and
orthogonal modes – slab modes plus IF modes, H–Z modes plus IF modes, and
guided modes – may be used as a basis set for determining the intrasubband and
intersubband electron–phonon scattering rates. Nash showed for a quantum well of
width d that
5
72 1
e 2
f nq (z) i ∝ ωnq d 2 q n 1 (q) n ∗1 (q) f n (q),
2
(C.33)
q 2q
and where
232 Appendices
Thus, for a given in-plane momentum transfer q the form factor f n (q) describes
the probability that an electron will scatter from the initial subband |i to the
final subband | f as a result of interaction with phonons of branch, n. The
explicit numerical calculations of Nash (1992) showed that the form factors f n (q)
corresponding to the three sets of complete and orthogonal modes – slab modes plus
IF modes, H–Z modes plus IF modes, and guided modes – are identical. This result
is evident from Figure C.1 for the case of intrasubband transitions and Figure C.2
for the case of intersubband transitions.
These results indicate that any of the three sets of complete and orthogonal
modes for a semiconductor layer mentioned above may be used to calculate
electron–phonon scattering for the nearly degenerate longitudinal optical phonon
modes in the layer. The key point is that the modes of each of the three sets must be
complete and orthonormal. This result explains why the relatively simple modes of
the so-called slab model may be used to perform such a scattering-rate calculation
even though these modes are not the normal modes of a heterolayer.
As further proof of the validity of this approach, Nash showed that it is possible
to perform unitary transformations between the slab modes plus IF modes, the H–Z
modes plus IF modes, and the guided modes. These calculations demonstrate that
it is essential to include the IF optical phonon modes in the set of modes for the
slab model – also known as the dielectric continuum model with electromagnetic
boundary conditions – and the Huang–Zhu models. Moreover, it is essential that the
IF modes are not included in the set of guided modes; this set of guided modes
is complete without the IF modes. Indeed, the guided modes explicitly include
IF-like exponentially decaying components outside the slab, unlike the slab and H–Z
modes. The dielectric continuum model with electromagnetic boundary conditions
leads to a convenient set of modes – the slab modes plus the interface modes –
for making relatively simple calculations of electron–phonon scattering rates in the
case where it is not necessary to account for the dispersion of the confined modes,
also referred to as the bulk-like modes in a slab. For GaAs and AlAs the total
dispersion over the entire Brillouin zone is only a few meV and it follows that
in typical carrier-transport calculations – including those of solid-state electronic
devices with dimensionally confined structures – it is not necessary to consider this
dispersion. As is evident in Chapter 10, it is necessary to consider energy differences
(ω),1 q,1
2
+ (ω)⊥,1 q 2 = 0 (ω),1 (ω)⊥,1 > 0,
(D.2)
(ω),2 κ22 − (ω)⊥,2 q 2 = 0 (ω),2 (ω)⊥,2 > 0.
From the continuity of the tangential component of the electric field at z = 0 it
then follows that A = C. From the continuity of the normal component of the
electric displacement at z = 0, ,1 q,1 B = ,2 κ2 C; thus, C = 1 B with 1 =
,1 k1, /(,2 κ2 ). Thus
1 cos q,1 + sin q,1 z > 0,
φ(r ) = φ0 eiq·ρρ × κ
(D.3)
1 e 2 z z < 0.
Now E⊥ and E are given by the appropriate gradients of φ0 . The normalization
condition of subsection 7.3.3,
! "
1 h̄ 1 1 ∂(ω)⊥,n 2
1 1 ∂(ω),n 2
= E ⊥,n + E,n dz,
L 2 2ω 4π 2ω ∂ω 4π 2ω ∂ω
(D.4)
2 2
may be calculated in terms of E⊥,n and E,n through
L/2 L/2
E⊥,1 2 dz = φ 2 q 2 sin q,1 z + 1 cos q,1 z 2 dz
0
−L/2 0
1 L
= φ02 q 2 (1 + 21 ) , (D.5)
2 2
L/2 L/2
E,1 2 dz = φ 2 q 2 sin q,1 z + 1 cos q,1 z 2 dz
0 ,1
−L/2 0
2 1 L
= φ02 q,1 (1 + 21 ) , (D.6)
2 2
where the normalization length L in the z-direction is defined over the region
(−L/2, +L/2) and where the limit L/2 → ∞ will be taken at a later stage of
238 Appendices
the calculation. Evidently, in the limit L/2 → ∞ the exponentially decaying mode
in region 2 makes a negligible contribution to the normalization integrals. Thus, it
follows that
4π h̄ 4 ∂⊥,1 2 ∂,1 2 −1
φ0 = 3
2
q + q , (D.7)
L 1 + 21 ∂ω ∂ω 1,
and that
4π e2 h̄ L −3
HHS =
q q1, (∂/∂ω)(⊥,1 sin2 θ 1 + ,1 cos2 θ1 )
1 2
× # # eiq·ρρ (aq + a−q
†
)
q + q,1 ,1 q,1 + ,2 κ2
2 2 2 2 2 2
,1 q,1 cos q,1 z + ,2 κ2 sin q,1 z z > 0,
× (D.8)
κ
,1 q,1 e 2 z z < 0,
where the summation is taken only over values of q,1 consistent with the condition
(ω),2 (ω)⊥,2 > 0. The case in which an optical phonon mode propagates into
region 1 from region 2 may be treated in the same manner.
The angular dependence of the phonon frequencies in würtzite heterostructures
leads to the situation – in contrast to the typical case for zincblende heterostructures
– where there is an overlap in the allowed phonon frequencies in adjacent material
regions. This situation leads to the occurence of modes that propagate across the
heterointerface. Consider an optical phonon incident on region 2 from region 1. In
this case, the phonon will propagate in region 2 if the phonon frequency in region 1,
ω, is such that (ω),2 (ω)⊥,2 < 0. As before, the allowed frequencies in region 1
2 + (ω)
are the solutions of (ω),1 q,1 ⊥,1 q = 0 or q,1 =
2 −(ω)⊥,1 /(ω),1 q.
If (ω),2 (ω)⊥,2 < 0, there exists a real solution for q,2 which satisfies
(ω),2 q,1
2
+ (ω)⊥,2 q 2 = 0, (D.9)
or
−(ω)⊥,2 (ω),1 (ω)⊥,2
q,2 = q= q,1 . (D.10)
(ω),2 (ω)⊥,1 (ω),2
it follows that continuity of the tangential component of the electric field at the
heterointerface leads to A + B = C and continuity of the normal component of the
electric displacement at the heterointerface results in the condition
Appendices 239
,1 k1,
(A − B)2 = C where 2 = .
,2 k2,
2 2
may be calculated in terms of E⊥,n and E,n through
L/2 2
E⊥,1 2 dz = φ 2 q 2 1 + 2 − 1 L
,
0
0 1 + 2 2
L/2 2
E,1 2 dz = φ 2 q 2 1 + 2 − 1 L
,
0 1,
0 1 + 2 2
0 2 (D.14)
E⊥,2 2 dz = φ 2 q 2 22 L
,
0
−L/2 1 + 2 2
0
22 2 L
E,2 2 dz = φ 2 q 2 .
0 1,
−L/2 1 + 2 2
These normalization integrals determine the propagating (PR) optical phonon modes
in a single-heterointerface würtzite structure. The corresponding electron–optical-
phonon interaction Hamiltonians are:
$ %1/2
S
HPR = 4π e2 h̄ L −3
q q1,
%−1/2
∂ $
× ⊥,1 (ω)q 2 + ,1 (ω)q1,
2
+ {1 ↔ 2}
∂ω
ρ †
2 cos q1, z z > 0,
× e (aq + a−q )
iq·ρ
(D.15)
2 cos(q2, z) z < 0,
$ %1/2
A
HPR = 4π e2 h̄ L −3
q q1,
−1/2
∂
× {[⊥,1 (ω)q 2 + ,1 (ω)q1,
2 2
],2 2
q2, } + {1 ↔ 2}
∂ω
ρ †
,2 q2, sin q1, z z > 0,
× e (aq + a−q )
iq·ρ
(D.16)
,1 q1, sin q2, z z < 0,
and
and
and
q2
= φ02 ,
2κ2
d/2 1 d/2
E⊥,1 2 dz = φ 2 q 2 cosh2 κ1 z dz
0
−d/2 cosh2 κ1 d/2 −d/2
2φ02 q 2 1 κ1 d κ1 d d
= sinh cosh + ,
cosh2 κ1 d/2 2κ1 2 2 4
−d/2 ∞ ∞
E,2 2 dz = E,2 2 dz = φ 2 κ 2 e−2κ2 (z−d/2) dz
0 2
−∞ d/2 d/2
κ2
= φ02 ,
2
d/2 1 d/2
E,1 2 dz = φ 2 κ 2 sinh 2 (κ1 z)dz
0 1
−d/2 cosh (κ1 d/2) −d/2
2
2φ02 κ12 1 κ1 d κ1 d d
= sinh cosh − .
cosh2 κ1 d/2 2κ1 2 2 4
(D.33)
−d/2
∞ ∞
E,2 2 dz = E,2 2 dz = φ 2 κ 2 e−2κ2 (z−d/2) dz
0 2
−∞ d/2 d/2
κ2
= φ02 ,
2
d/2
d/2 1
E,1 2 dz = φ 2 κ 2 cosh 2 κ1 z dz
0 1
−d/2 sinh 2 (κ1 d/2) −d/2
2φ02 κ12 1 κ1 d κ1 d d
= sinh cosh + .
sinh 2 κ1 d/2 2κ1 2 2 4
(D.34)
with the range of frequencies determined by ,1 ,2 < 0 and, as discussed
previously, ⊥,1 ,1 > 0 and ⊥,2 ,2 > 0. For the antisymmetric mode,
! "−1/2
−2 ∂ √ √
HIF =
A 2
4π e h̄ L [ ⊥,1 ,1 coth( ⊥,1 /,1 qd/2) − ⊥,2 ,2 ]
q ∂ω
1
× √ eiq·ρρ (aq + a−q†
)
2q
sinh( ⊥,1 /,1 qz)
|z| < d/2,
× sinh( ⊥,1 /,1 qd/2) (D.37)
√
sgn(z) e− ⊥,2 /,2 q(|z|−d/2) |z| > d/2,
where the frequency is determined by
√ √
⊥,1 ,1 coth( ⊥,1 /,1 qd/2) − ⊥,2 ,2 = 0, with ,1 ,2 < 0.
The dispersion relations for these interface modes are displayed in Figure D.2 for a
würtzite AlN/GaN/AlN heterostructure having a quantum well of thickness d.
244 Appendices
(ω),1 q,1
2
− (ω)⊥,1 q 2 = 0 (ω),1 (ω)⊥,1 < 0 for |z| < d/2,
(ω),2 q,2
2
− (ω)⊥,2 q 2 = 0 (ω),2 (ω)⊥,2 > 0 for |z| > d/2.
(D.39)
Clearly, the continuity of the tangential component of the electric field at the
heterointerfaces is ensured by the choice of φ(r ). The continuity of the normal com-
ponent of the electric displacement at the heterointerfaces leads to the requirement
that
(ω),1 q,1 sin q1, d/2 − (ω),2 κ2 cos q,1 d/2 = 0. (D.40)
1 h̄ 1 1 ∂(ω)⊥,n 2 1 1 ∂(ω),n 2
= E⊥,n + E,n dz,
L 2 2ω 4π 2ω ∂ω 4π 2ω ∂ω
(D.41)
2
may be calculated in terms of the appropriate integrals of E⊥,n = q 2 |φ(z)|2 and
2
E,n = |∂φ(z)/∂z|2 . For the symmetric case, these integrals are
2
+∞ ∂(ω)
⊥,n E⊥,n 2
n=1 −∞
∂ω
∂(ω)⊥,1 d 1 q,1 d q,1 d
= q 2 φ02 + sin cos
∂ω 2 q,1 2 2
1 ∂(ω)⊥,2 q,1 d
+ q 2 φ02 cos2 (D.42)
κ2 ∂ω 2
and
2
+∞ ∂(ω)
,n E⊥,n 2
n=1 −∞
∂ω
∂(ω),1 d 1 q,1 d q,1 d
= φ02 − sin cos
∂ω 2 q,1 2 2
∂(ω),2 q ,1 d
+ φ02 κ2 cos2 ; (D.43)
∂ω 2
thus
!
1 h̄ 1 1 2 ∂(ω)⊥,1 2 ∂(ω),1 2 d
= φ q + q,1
L 2 2ω 4π 2ω 0 ∂ω ∂ω 2
/ 0
∂(ω)⊥,1 q 2 ∂(ω),1 q,1 d q,1 d
+ − q,1 sin cos
∂ω q,1 ∂ω 2 2
/ 0
∂(ω)⊥,2 q 2 ∂(ω),2 q1, d
+ + κ2 cos2
∂ω κ2 ∂ω 2
! "
1 1 2 ∂(ω)⊥,1 2 ∂(ω),1 2 d ∂ fs q,1 d
= φ q + q,1 − 2q cos
4π 2ω 0 ∂ω ∂ω 2 ∂ω 2
! 2 3 "
1 1 2 ∂ d ∂ fs q,1 d
= φ (ω)⊥,1 q + (ω),1 q,1
2 2
− 2q cos ,
4π 2ω 0 ∂ω 2 ∂ω 2
(D.44)
Thus,
! "
∂ 2 3d ∂ fs q,1 d −1
φ02 = 4π h̄ L −2 (ω)⊥,1 q 2 + (ω),1 q,1
2
− 2q cos
∂ω 2 ∂ω 2
(D.46)
and
!
∂ 2 3d
HCS = eiq·ρρ (aq + a−q
†
)4π h̄ L −2 (ω)⊥,1 q 2 + (ω),1 qm,1
2
q m ∂ω 2
"
∂ qm,1 d −1/2
× −2q f s (ω) cos
∂ω 2
cos qm,1 z |z| < d/2,
× (D.47)
−κ
cos qm,1 d/2 e 2 (z−d/2) z > d/2,
where
q,1 d
f a (ω) = sgn[(ω),1 ] −(ω)⊥,1 (ω),1 cos
2
q,1 d
+ sgn[(ω),2 ] (ω)⊥,2 (ω),2 sin (D.50)
2
and where the discrete confined wavevectors q1,m are determined from
(ω),1 qm,1 cos qm,1 d/2 + (ω),2 κ2 sin qm,1 d/2 = 0 (D.51)
with 2(m − 1)π/d < qm,1 < 2(m + 1)π/d and κ2 = (ω)⊥,2 /(ω),2 q.
Appendices 247
Defining 3 = (ω),2 q,2 /(ω),1 κ1 it follows that A = B3 coth κ1 d/2. Then,
with B ≡ φ0 ,
φ(r ) = φ0 eiq·ρρ
− sin[q,1 (z + d/2)] + 3 coth κ1 d/2 cos[q,1 (z + d/2)] z < d/2,
× 3 (cosh κ1 z)/(sinh κ1 d/2) |z| < d/2,
sin[q (z − d/2)] + coth κ d/2 cos[q (z − d/2)]
,1 3 1 ,1 z > d/2,
(D.53)
or
sin[q,1 (|z| − d/2)]
φ(r ) = φ0 eiq·ρρ × + 3 coth κ1 d/2 cos[q,1 (|z| − d/2)] |z| > −d/2,
3 (cosh κ1 z)/(sinh κ1 d/2) |z| < d/2.
(D.54)
For the antisymmetric modes take
−A cos[q,1 (z + d/2)] + B sin[q,1 (z + d/2)] z < −d/2,
φ(r ) = eiq·ρρ × C sinh κ1 z |z| < d/2,
A cos[q,1 (z − d/2)] + B sin[q,1 (z − d/2)] z > d/2.
(D.55)
Then, continuity of the tangential component of the electric field at the heteroin-
terfaces yields A = C sinh κ1 d/2 and continuity of the normal component of
the electric displacement at the heterointerfaces requires that (ω),2 q,2 B =
(ω),1 κ1 C cosh κ1 d/2. Thus, it follows that A = B3 tanh κ1 d/2. Then, with
B ≡ φ0 ,
φ(r ) = φ0 eiq·ρρ
sin[q,2 (z + d/2)] − 3 tanh κ1 d/2 cos[q,2 (z + d/2)] z < −d/2,
× 3 sinh κ1 z/(cosh κ1 d/2) |z| < d/2,
sin[q (z − d/2)] + tanh κ d/2 cos[q (z − d/2)]
,2 3 1 ,2 z > d/2,
(D.56)
248 Appendices
or
φ(r ) = φ0 eiq·ρρ
sgn(z) sin[q,2 (|z| − d/2)]
× + 3 tanh κ1 d/2 cos[q,2 (|z| − d/2)] |z| > −d/2, (D.57)
(sinh κ z)/(cosh κ d/2)
3 1 1 |z| < d/2.
Once again, the normalization condition of subsection 7.3.1,
1 h̄ 1 1 ∂(ω)⊥,n 2
+ 1 1 ∂(ω),n E,n 2 dz,
= E ⊥,n
L 2 2ω 4π 2ω ∂ω 4π 2ω ∂ω
(D.58)
2
may be calculated in terms of the appropriate integrals of E⊥,n = q 2 |φ(z)|2 and
E,n 2 = |∂φ(z)/∂z|2 . For the symmetric case, the integrals that make contributions
in the limit L −→ ∞ are
L/2 $ %
E⊥,2 2 dz = φ 2 q 2 1 1 + 2 coth2 κ1 d/2) L ,
0 3
d/2 2 2
L/2 (D.59)
2 3
E,2 2 dz = φ 2 q 2 1 1 + 2 coth2 (κ1 d/2) L .
0 ,2 3
d/2 2 2
For the antisymmetric case, the integrals that make contributions in the limit L →
∞ are
−d/2 L/2 $ %
E⊥,2 2 dz = E⊥,2 2 dz = φ 2 q 2 1 1 + 2 tanh 2 κ1 d/2 L ,
0 3
−L/2 d/2 2 2
−d/2 L/2 $ %
E,2 2 dz = E,2 2 dz = φ 2 q 2 1 1 + 2 tanh 2 κ1 d/2 L .
0 ,2 3
−L/2 d/2 2 2
(D.60)
Thus
4π h̄ ∂(ω)⊥,2 2 ∂(ω)⊥,2 2 −1
φ02 = 3 ×2 q + q,2
L ∂ω ∂ω
(1 + 23 coth2 κ1 d/2)−1 symmetric case,
× (D.61)
(1 + 23 tanh2 κ1 d/2)−1 antisymmetric case.
The interaction Hamiltonian for the symmetric HS modes is then
−1/2
−3 ∂
HHS =
S 2
4π e h̄ L (⊥,2 sin θ2 + ,2 cos θ2 )
2 2
eiq·ρρ (aq + a−q
†
)
q q,2 ∂ω
√
2 −1/2 2
−1/2
× 2 q 2 + q,2 ,1 κ12 sinh2 κ1 d/2 + ,2
2 2
q,2 cosh2 κ1 d/2
{,1 κ1 sinh κ1 d/2 sin[q,2 (|z| − d/2)]
× + ,2 q,2 cosh κ1 d/2 cos[q,2 (|z| − d/2)]} |z| > d/2, (D.62)
q cosh κ z
,2 ,2 1 |z| < d/2,
Appendices 249
where the sum is over q,2 and is taken only over those values (D.62) such that
(ω),1 (ω)⊥,1 > 0. Likewise, the interaction Hamiltonian for the antisymmetric
HS modes is
−1/2
∂
HHS =
A 2
4π e h̄ L −3
(⊥,2 sin θ 2 + ,2 cos θ2 )
2 2
eiq·ρρ (aq + a−q
†
)
q q,2 ∂ω
√ 2 −1/2 2
× 2(q 2 + q,2 ) (,1 κ12 cosh2 κ1 d/2 + ,2
2 2
q,2 sinh2 κ1 d/2)−1/2
sgn(z) ,1 κ1 cosh κ1 d/2 sin[q,2 (|z| − d/2)]
× + ,2 q,2 sinh κ1 d/2 cos[q,2 (|z| − d/2)] |z| > d/2,
(D.63)
,2 q,2 sinh κ1 z |z| < d/2.
The interaction Hamiltonians for the propagating (PR) modes may be obtained
straightforwardly from those of the HS modes by making the substitution κ1 →
iq,1 . Indeed, this substitution results in 3 = −i3 with
(ω),2 q,2
3 = ,
(ω),1 q,1
cosh κ1 z → cos q,1 z,
sinh κ1 z → i sin q,1 z, (D.64)
coth κ1 d/2 → −i cot q,1 d/2,
tanh κ1 d/2 → −i tan q,1 d/2.
The interaction Hamiltonian for the symmetric PR modes is then
−1/2
∂
S
HPR = 4π e2 h̄ L −3 (⊥,2 q 2 + ,2 q,2
2
) eiq·ρρ (aq + a−q
†
)
q q,2 ∂ω
√ 2 2 −1/2
× 2 ,1 q,1 sin2 q,1 d/2 + ,2
2 2
q,2 cos2 q,1 d/2
,2 q,2 cos q,1 d/2 cos[q,2 (|z| − d/2)]
× − ,1 q,1 sin q,1 d/2 sin[q,2 (|z| − d/2)] |z| > d/2,
(D.65)
,2 q,2 cos q,1 z |z| < d/2,
where the sum is over q,2 and is taken only over these values such that
(ω),1 (ω)⊥,1 < 0. Likewise, the interaction Hamiltonian for the antisymmetric
PR modes is
−1/2
∂
A
HPR = 4π e2 h̄ L −3 (⊥,2 q 2 + ,2 q,2
2
) eiq·ρρ (aq + a−q
†
)
q q,2 ∂ω
√ 2 2 −1/2
× 2 ,1 q,1 cos2 q,1 d/2 + ,22 2
q,2 sin2 q,1 d/2
sgn(z) ,2 q,2 sin q,1 d/2 cos[q,2 (|z| − d/2)]
× + ,1 q,1 cos q,1 d/2 sin[q,2 (|z| − d/2)] |z| > d/2,
q sin q z (D.66)
,2 ,2 ,1 |z| < d/2.
250 Appendices
h̄ ∂ψ(x, t)
[H + Hc−p (t)]ψ(x, t) = − , (E.1)
i ∂t
where a one-dimensional system is considered for the sake of simplicity. The final
result for the Fermi golden rule will be independent of the assumption that the
system is one dimensional. Taking the wavefunction to be of the form
ψ(x, t) = cn (t)φn (x) e−i E n t/h̄ , (E.2)
n
where φn (x) is the eigenstate of the unperturbed system with quantum number n, it
follows that,
[H + Hc−p (t)]ψ(x, t) = cn (t)E n φn (x) e−i E n t/h̄
n
+ cn (t) e−i E n t/h̄ Hc−p (t)φn (x) (E.3)
n
and
h̄ ∂ψ(x, t) h̄ dcn (t)
− =− φn (x) e−i E n t/h̄
i ∂t i n dt
+ cn (t)E n φn (x) e−i E n t/h̄ . (E.4)
n
Now if it is assumed that the system is initially in state n, so that cn (0) = 1 and all
other cm (0) are zero, the time evolution of state m is approximated by
Appendices 251
i t 5 7
cm (t) − dt φm Hc−p (t) φn e−i(E n −E m )t/h̄ . (E.7)
h̄ 0
5 7 e−i(E n −E m )t/h̄ − 1
cm (t) = φm Hc−p φn (E.8)
En − Em
or
5 72 sin2 [(E n − E m )t/(2h̄)]
|cm (t)|2 = φm Hc−p φn . (E.9)
[(E n − E m )/2]2
In the case of nearly degenerate states E n E m , and where the transition from
state n is to a dense group of final states centered around E m , the probability of a
transition, |cm (t)|2 , takes the form
E n +E/2
5 72 sin2 [(E n − E)t/(2h̄)]
P= d Eρ(E) φm Hc−p φn
E n −E/2 [(E n − E)/2]2
E n +E/2
5 72 sin2 [(E n − E)t/(2h̄)]
4ρ(E) φm Hc−p φn dE , (E.10)
E n −E/2 (E n − E)2
5 72
where the last result follows since ρ(E) φm Hc−p φn varies slowly with energy
relative to [sin2 (E n − E)t/(2h̄)]/(E n − E)2 . Taking x = (E − E n )t/(2h̄),
2t 5 72 tE/4h̄ sin2 x
P = ρ(E) φm Hc−p φn dx . (E.11)
h̄ −tE/4h̄ x2
Thus, the total interaction potential energy, V̄ (r), is the sum of that associated
with the phonons in the absence of the Coulomb screening, −ephonon (r), and that
induced by the Coulomb screening effects, −einduced (r):
where
V̄ = V e−iωt+γ t + V ∗ eiωt+γ t ,
φk |V |φk ≡ d 2ρ e−ik ·ρρ V eik·ρρ , (F.9)
5 ∗ 7∗
φk V φk ≡ φk |V |φk .
254 Appendices
The induced charge density is then proportional to the difference of the probability
densities of the perturbed and unperturbed wavefunctions, multiplied by the Fermi–
Dirac distribution function Fk :
! 1
"
ρ induced
(r) = −2e |φk (r, t)| − |χ1 (z)| Fk ,
2 2
(F.10)
k
S
where −e is the charge of a single carrier and the factor of two is introduced since
there are two spin states for every electronic quantum state. Thus,
5 7
χ 2 (z) k |V |k ei(k −k)·ρρ e−iωt
ρ induced
(r) = −2e 1
k
S k
E k − E k + h̄ω + iγ
5 7
k |V |k ∗ e−i(k −k)·ρρ eiωt
+
k
E k − E k + h̄ω − iγ
5 ∗ 7 i(k −k)·ρρ iωt
k |V | k e e
+
k
E k − E k − h̄ω + iγ
5 ∗ 7∗ −i(k −k)·ρρ −iωt
k |V | k e e
+ Fk
k
E k − E k − h̄ω − iγ
5 7
χ12 (z) k |V |k ei(k −k)·ρρ
= −2e
S k,k E k − E k + h̄ω + iγ
5 ∗ 7∗ −i(k −k)·ρρ
k |V | k e
+ e−iωt Fk + c.c.
E k − E k − h̄ω − iγ
χ12 (z) 5 7 Fk − Fk
= −2e k |V |k ei(k −k)·ρρ . (F.11)
S k,k E k − E k + h̄ω + iγ
5 7
Defining q = k −k, noting that the matrix element k |V |k depends on the modulus
5 7
of k − k, and defining V (|k − k|) ≡ k |V |k , it follows that
or, alternatively,
where
Appendices 255
2 Fk+q − Fk
A(q, ω) ≡ − . (F.14)
S k E k − E k+q + h̄ωq + iγ
Then
6 4 6 4
induced (q) = −A(q, ω) total (q) C(q), (F.16)
and
dρρ eiq·ρρ
C(q) = 4πe dzdz χ12 (z)χ12 (z ) . (F.18)
ρ 2 + (z − z )2
The identity
1 2π 1 iq·(ρ −ρ ) −q |z−z |
= e e , (F.19)
|r − r | S q q
so that
6 4 6 4
total (q) = phonon (q) /[1 − A(q, ω)C(q)]
6 4
= phonon (q) /κel (q), (F.23)
256 Appendices
In Section 10.6, this expression for the electron permittivity, κel (q), is used to take
into account the effect of electron screening.
References
Alexson, D., Bergman, L., Dutta, M., Kim, K.W., Komirenko, S., Nemanich, R.J., Lee, B.C.,
Stroscio, M.A., and Yu, S. (1999), Confined phonons and phonon-mode properties of
III-V nitrides with würtzite crystal structure, Physics, B263 and B264, 510–513.
Alexson, D., Bergman, L., Nemanich, R.J., Dutta, M., Stroscio, M.A., Parker, C.A., Bedair,
S.M., El-Masry, N.A., and Adar, F. (2000), UV Raman study of A1 (LO) and E2 phonons
in Inx Ga1−x N alloys, Applied Physics Letters, 89, 798–800.
Auld, B.A. (1973), Acoustic Fields and Waves in Solids. Wiley-Interscience, John Wiley &
Sons, New York.
Azuhata, T., Sota, T., Suzuki, K., and Nakamura, S. (1995), Polarized Raman spectra in GaN,
Journal of the Physics of Condensed Matter, 7, L129–L135.
Bannov, N., Mitin, V., and Stroscio, M. (1994a), Confined acoustic phonons in semiconductor
slabs and their interactions with electrons, Physica Status Solidi B, 183, 131–138.
Bannov, N., Mitin, V., Stroscio, M.A., and Aristov, V. (1994b), Confined acoustic phonons:
density of states, in Proceedings of the 185th Meeting of the Electrochemical Society, Vol.
94-1, 521; San Francisco, California, May 22–27, 1994.
Bannov, N., Aristov, V., Mitin, V., and Stroscio, M.A. (1995), Electron relaxation times due to
deformation-potential interaction of electrons with confined acoustic phonons in a
free-standing quantum well, Physical Review, B51, 9930–9938.
Bastard, G. (1988), Wave Mechanics Applied to Semiconductor Heterostructures. Halsted
Press, New York.
Behr, D., Niebuhr, R., Wagnar, J., Bachem, K.H., and Kaufmann, U. (1997), Resonant Raman
scattering in GaN/(AlGa)N single quantum wells, Applied Physics Letters, 70, 363–365.
Belenky, G., Dutta, M., Gorfinkel, V.B., Haddad, G.I., Iafrate, G.J., Kim, K.W., Kisin, M.,
Luryi, S., Stroscio, M.A., Sun, J.P., Teng, H.B., and Yu, S. (1999), Tailoring of optical
modes in nanoscale semiconductor structures: role of interface-optical phonons in
quantum-well lasers, Physica B, 263–264, 462–465.
257
258 References
Beltzer, A.I. (1988), Acoustics of Solids. Springer-Verlag, Berlin, Heidelberg, New York,
London, Paris, and Tokyo.
Bergman, L., Alexson, D., Murphy, P.L., Nemanich, R.J., Dutta, M., Stroscio, M.A., Balkas,
C., Shih, H., and Davis, R.F. (1999a), Raman analysis of phonon lifetimes in AlN and
GaN of würtzite structure, Physical Review, B59, 12 977–12 982.
Bergman, L., Dutta, M., Balkas, C., Davis, R.F., Christman, J.A., Alexon, D., and Nemanich,
R.J. (1999b), Raman analysis of the E1 and A1 quasi-longitudinal optical and
quasi-transverse optical modes in würtzite AlN, Journal of Applied Physics, 85,
3535–3539.
Bergman, L., Dutta, M., Kim, K.W., Klemens, P.G., Komirenko, S., and Stroscio, M.A. (2000),
Phonons, electron–phonon interactions, and phonon–phonon interactions in III-V nitrides,
Proceedings of the SPIE, Vol. 3914, 13–22.
Bhatt, A.R., Kim, K.W., Stroscio, M.A., Iafrate, G.J., Dutta, M., Grubin, H.L., Haque, R., and
Zhu, X.T. (1993a), Reduction of LO-phonon interface modes using metal–semiconductor
heterostructures, Journal of Applied Physics, 73, 2338–2342.
Bhatt, A.R., Kim, K.W., Stroscio, M.A., and Higman, J.M. (1993b), Simplified microscopic
model for electron–optical-phonon interactions in quantum wells, Physical Review, B48,
14 671–14 674.
Bhatt, A.R., Kim, K.W., and Stroscio, M.A. (1994), Theoretical calculation of the
longitudinal-optical lifetime in GaAs, Journal of Applied Physics, 76, 3905–3907.
Blakemore, J.S. (1985), Solid State Physics, 2nd edition. Cambridge University Press,
Cambridge.
Blatt, J.M. and Thompson, C.J. (1963), Shape resonances in superconducting films, Physical
Review Letters, 10, 332–334.
Borer, W.J., Mitra, S.S., and Namjoshi, K.V. (1971), Line shape and temperature dependence of
the first order Raman spectrum of diamond, Solid State Communications, 9, 1377–1381.
Born, M. and Huang, K. (1954), Dynamical Theory of Crystal Lattices. Oxford University
Press, Oxford.
Burns, G., Dacol, F., Marinace, J.C., and Scott, B.A. (1973), Raman scattering in thin-film
waveguides, Applied Physics Letters, 22, 356–357.
Campos, V.B., Das Sarma, S., and Stroscio, M.A. (1992), Hot carrier relaxation in
polar-semiconductor quantum wires: confined LO-phonon emission, Physical Review,
B46, 3849–3853.
Cardona, M. (1975), ed. Light Scattering in Solids. Springer-Verlag, New York.
Cardona, M. and Güntherodt, G., eds. (1982a), Light Scattering in Solids II. Springer-Verlag,
Berlin.
Cardona, M. and Güntherodt, G., eds. (1982b), Light Scattering in Solids III. Springer-Verlag,
Berlin.
Cardona, M. and Güntherodt, G., eds. (1984), Light Scattering in Solids IV. Springer-Verlag,
Berlin.
Cardona, M. and Güntherodt, G., eds. (1989), Light Scattering in Solids V. Springer-Verlag,
Berlin.
Cardona, M. and Güntherodt, G., eds. (1991), Light Scattering in Solids VI. Springer-Verlag,
New York.
References 259
Castro, G.R. and Cardona, M., eds. (1987), Lectures on Surface Science, Proceedings of the
Fourth Latin-American Symposium. Springer-Verlag, Berlin.
Chang, I.F. and Mitra, S.S. (1968), Application of modified random-element-isodisplacement
model to long-wavelength optic phonons in mixed crystals, Physical Review, 172,
924–933.
Cheng, T.K., Vidal, J., Zeiger, H.J., Dresselhaus, G., Dresselhaus, M.S., and Ippen, E.P. (1991),
Mechanism for displacive excitation of coherent phonons in Sb, Bi, Te, and Ti2 O3 ,
Applied Physics Letters, 59, 1923–1925.
Cho, G.C., Kutt, W., and Kurz, H. (1990), Subpicosecond time-resolved coherent-phonon
oscillations in GaAs, Physical Review Letters, 65, 764–766.
Choi, K.-K., Newman, P.G., and Iafrate, G.J. (1990), Quantum transport and phonon emission
of nonequilibrium hot electrons, Physical Review, B41, 10 250–10 253.
Choi, K.-K., Tidrow, M.Z., Chang, W.H. (1996), Electron energy relaxation in infrared
hot-electron transistors, Applied Physics Letters, 68, 358–360.
Cingolani, A., Ferrara, M., Lugara, M., and Scamarcio, G. (1986), First order Raman scattering
in GaN, Solid State Communications, 58, 823–824.
Cleland, A.N. and Roukes, M.L. (1996), Fabrication of high frequency nanometer scale
mechanical resonators from bulk Si crystals, Applied Physics Letters, 69, 2653–2655.
Colvard, C., Merlin, R., Klein, M.V., and Gossard, A.C. (1980), Observation of folded acoustic
phonons in semiconductor superlattices, Physical Review Letters, 45, 298–301.
Comas, F., Trallero-Giner, C., and Cardona, M. (1997), Continuum treatment of phonon
polaritons in semiconductor heterogeneous structures, Physical Review, B56, 4115–4127.
Constantinou, N.C. (1993), Interface optical phonons near perfectly conducting boundaries and
their coupling to electrons, Physical Review, B48, 11 931–11 935.
Cowley, R.A. (1963), Advances in Physics, 12, 421–480.
Cros, A., Angerer, H., Ambacher, O., Stutzmann, M., Hopler, R., and Metzger, T. (1997),
Raman study of optical phonons in Alx Ga1−x N alloys, Solid State Communications, 104,
35–39.
Das Sarma, S., Stroscio, M.A., and Kim, K.W. (1992), Confined phonon modes and
hot-electron relaxation in semiconductor microstructures, Semiconductor Science and
Technology, 7, B60–B66.
de la Cruz, R.M., Teitsworth, S.W., and Stroscio, M.A. (1993), Phonon bottleneck effects for
confined longitudinal optical phonons in quantum boxes, Superlattices and
Microstructures, 13, 481–486.
de la Cruz, R.M., Teitsworth, S.W., and Stroscio, M.A. (1995), Interface phonons in spherical
GaAs/Alx Ga1−x As quantum dots, Physical Review, B52, 1489–1492.
Debernardi, A. (1998), Phonon linewidth in III-V semiconductors from density-functional
perturbation theory, Physical Review, B57, 12 847–12 858.
Demangeot, F., Frandon, J., Renucci, M.A., Meny, C., Briot, O., and Aulombard, R.L. (1997),
Interplay of electrons and phonons in heavily doped GaN epilayers, Journal of Applied
Physics, 82, 1305–1309.
Demangeot, F., Groenen, J., Frandon, J., Renucci, M.A., Briot, O., Clur, S., and Aulombard,
R.L. (1998), Coupling of GaN- and AlN-like longitudinal optic phonons in Ga1−x Alx N
solid solutions, Applied Physics Letters, 72, 2674–2676.
260 References
Di Bartolo, B. (1969), Optical Interactions in Solids. John Wiley and Sons, New York.
Dutta, M. and Stroscio, M.A., eds. (1998), Quantum-based Electronic Devices and Systems.
World Scientific, Singapore, New Jersey, London, Hong Kong.
Dutta, M. and Stroscio, M.A., eds. (2000), Advances in Semiconductor Lasers and
Applications to Optoelectronics. World Scientific, Singapore, New Jersey, London, Hong
Kong.
Dutta, M., Grubin, H.L., Iafrate, G.J., Kim, K.W., and Stroscio, M.A. (1993),
Metal-encapsulated quantum wire for enhanced charge transport, US Patent Number
5 264 711 issued November 23, 1993.
Dutta, M., Stroscio, M.A., and Kim, K.W. (1998), Recent developments on electron–phonon
interactions in structures for electronic and optoelectronic devices, in Quantum-Based
Electronic Devices and Systems, Selected Topics in Electronics and Systems 14. World
Scientific, Singapore, New Jersey, London, Hong Kong.
Dutta, M., Stroscio, M.A., Bergman, L., Alexson, D., Nemanich, R.L., Dupuis, R. (2000),
Observation of interface optical phonons in würtzite GaN/AlN superlattices, to be
published.
Edgar, J.H., ed. (1994), Properties of Group III Nitrides. INSPEC, London.
Educato, J.L., Leburton, J.-P., Boucaud, P., Vagos, P., and Julien, F.H. (1993), Influence of
interface phonons on intersubband scattering in asymmetric coupled quantum wells,
Physical Review, B47, 12 949–12 952.
Empedocles, S.A., Norris, D.J., and Bawendi, M.G. (1996), Photoluminescence spectroscopy
of single CdSe nanocrystallite quantum dots, Physical Review Letters, 77, 3873–3876.
Engleman, R. and Ruppin, R. (1968a), Optical lattice vibrations in finite ionic crystals: I,
Journal of Physics C, Series 2, 1, 614–629.
Engleman, R. and Ruppin, R. (1968b), Optical Lattice vibrations in finite ionic crystals: II,
Journal of Physics C, Series 2, 1, 630–643.
Engleman, R. and Ruppin, R. (1968c), Optical lattice vibrations in finite ionic crystals: III,
Journal of Physics C, Series 2, 1, 1515–1531.
Esaki, L. and Tsu, R. (1970), Superlattice and negative differential conductivity in
semiconductors, IBM Journal of Research and Development, 14, 61–65.
Faist, J., Capasso, F., Sirtori, C., Sivco, D.L., Baillargeon, J.N., Hutchinson, L.A., Chu,
Sung-Nee G., and Cho, A. (1996a), High power mid-infrared quantum cascade lasers
operating above room temperature, Applied Physics Letters, 68, 3680–3682; also see Hu,
Q., and Feng, S. (1991), Feasibility of far-infrared lasers using multiple semiconductor
quantum wells, Applied Physics Letters, 59, 2923–2925.
Faist, J., Capasso, F., Sirtori, C., Sivco, D.L., Hutchnison, A.L., Hybertson, M.S., and Cho,
A.Y. (1996b), Quantum cascade lasers without intersubband population inversion,
Physical Review Letters, 76, 411–414.
Fasol, G., Tanaka, M., Sakaki, H., and Horikosh, Y. (1988), Interface roughness and dispersion
of confined LO phonons in GaAs/AlAs quantum wells, Physical Review, B38,
6056–6065.
Ferry, D.K. (1991), Semiconductors. Macmillan Co., New York.
Fuchs, R. and Kliewer, K.L. (1965), Optical modes of vibration in an ionic crystal slab,
Physical Review, 140, A2076–A2088.
References 261
Fuchs, R., Kliewer, K.L., and Pardee, W.J. (1966), Optical properties of an ionic crystal slab,
Physical Review, 150, 589–596.
Gauthier-Lafaye, O., Sauvage, S., Boucaud, P., Julien, F.H., Prazeres, R., Glotin, F., Ortega,
J.-M., Thierry-Mieg, V., Planel, R., Leburton, J.-P., and Berger, V. (1997), Intersubband
stimulated emission in GaAs/AlGaAs quantum wells: pump-probe experiments using a
two-color free electron laser, Physical Review Letters, 70, 3197–3200.
Gelmont, B., Gorfinkel, S., and Luryi, S. (1996), Theory of the spectral line shape and gain in
quantum wells with intersubband transitions and ionic crystals, Applied Physics Letters,
68, 2171–2173.
Giehler, M., Ramsteiner, M., Brandt, O., Yarg, H., and Ploog, K.H. (1995), Optical phonons of
hexagonal and cubic GaN studied by infrared transmission and Raman spectroscopy,
Applied Physics Letters, 67, 733–735.
Gleize, J., Renucci, M.A., Frandon, J., and Demangeot, F. (1999), Anisotropy effects on polar
optical phonons in würtzite GaN/AlN superlattices, Physical Review, B60, 15 985–15 992.
Gleize J., Demangeot, F., Frandon J., Renucci, M.A., Kuball M., Grandjean N., and Massies, J.
(2000), Resonant Raman scattering in (Al,Ga)N/GaN quantum well structures, Thin Solid
Films, 364, 156–160.
Gmachl, C., Capasso, F., Tredicucci, D.L., Sivco, D.L., Hutchinson, A.L., Chu, S.N. G., and
Cho, A.Y. (1998), Noncascaded intersubband injection lasers at λ = 7.7 µm, Applied
Physics Letters, 73, 3830–3832.
Göppert M., Hetterich, M., Dinger A., Klingshorn C., and O’Donnell, K.P. (1998), Infrared
spectroscopy of confined optical and folded acoustical phonons in strained CdSe/CdS
superlattices, Physical Review, B57, 13 068–13 071.
Gorczyca, I., Christensen, N.E., Peltzer y Blanca, E.L., and Rodriguez, C.O. (1995), Optical
phonon modes in GaN and AlN, Physical Review, B51, 11 936–11 939.
Gorfinkel, V., Luryi, S., and Gelmont, B. (1996), Theory of gain spectra for quantum cascade
lasers and temperature dependence of their characteristics at low and moderate carrier
concentrations, IEEE Journal of Quantum Electronics, 32, 1995–2003.
Harima, H., Inoue, T., Nakashima, S., Okumura, H., Ishida, Y., Koiguari, T., Grille, H., and
Bechstedt, F. (1999), Raman studies on phonon modes in cubic AlGaN alloys, Applied
Physics Letters, 74, 191–193.
Hayashi, K., Itoh, K., Sawaki, N., and Akasaki, I. (1991), Raman scattering in Alx Ga1−x N
alloys, Solid State Communications, 77, 115–118.
Hayes, W. and Loudon, R. (1978), Scattering of Light by Crystals. John Wiley and Sons, New
York.
Hess, Karl (1999), Advanced Theory of Semiconductor Devices. IEEE Press, New Jersey.
Hewat, A.W. (1970), Lattice dynamics of ZnO and BeO, Solid State Communications, 8,
187–189.
Hoben, M.V. and Russell, J.P. (1964), The Raman spectrum of gallium phosphide, Physics
Letters, 13, 39–41.
Hon, D.T. and Faust, W.L. (1973), Dielectric parameterization of Raman lineshapes for GaP
with a plasma of charge carriers, Applied Physics Letters, 1, 241–256.
Hu, Q. and Feng, S. (1991), Feasibility of far-infrared lasers using multiple semiconductor
quantum wells, Applied Physics Letters, 59, 2923–2925.
262 References
Huang, K. (1951), On the interaction between the radiation field and ionic crystals,
Proceedings of the Royal Society A, 208, 352–365.
Huang, K. and Zhu, B. (1988), Dielectric continuum model and Fröhlich interaction in
superlattices, Physical Review, B38, 13 377–13 386.
Hwang, E.H., Das Sarma, S., and Stroscio, M.A. (2000), Role of confined phonons in thin-film
superconductivity, Physical Review, B61, 8659–8662.
Iafrate, Gerald J., and Stroscio, M.A. (1996), Application of quantum-based devices: trends
and challenges, IEEE Transactions on Electron Devices, 43, 1621–1625.
Iijima, S. (1991), Helical microtubules of graphitic carbon, Nature, 354, 56–57.
Irmer, G., Toporov, V.V., Bairamov, B.H., and Monecke, J. (1983), Determination of the charge
carrier concentration and mobility in n-GaP by Raman spectroscopy, Physics Status
Solidi, B119, 595–603.
Julien, F.H., Sa’ar, A., Wang, J., and Leburton, J.-P. (1995), Optically pumped intersubband
emission in quantum wells, Electronics Letters, 31, 838–839.
Jusserand, B. and Cardona M. (1991), Raman spectroscopy of vibrations in supertlattices, in
Light Scattering in Solids V, ed. M. Cardona, and G. Güntherodt. Springer-Verlag, Berlin.
Jusserand, B. and Paguet, D. (1986), Comment on ‘Resonance Raman Scattering by confined
LO and TO phonons in GaAs–AlAs superlattices’, Physical Review Letters, 56, 1752.
Kash, J.A. and Tsang, J.C. (1991), Light scattering and other secondary emission studies of
dynamical processes in semiconductors, in Light Scattering in Solids VI, M. Cardona, ed.
Springer-Verlag, Berlin.
Kash, J.A., Jha, S.S., and Tsang, J.C. (1987), Picosecond Raman studies of the Fröhlich
interaction in semiconductor alloys, Physical Review Letters, 58, 1869–1872; see also
Nash, K.J., and Skolnick, M.S. (1988), Comment on ‘Picosecond Raman studies of the
Fröhlich interaction in semiconductor alloys’, Physical Review Letters, 60, 863.
Kash, J.A., Jha, S.S., and Tsang, J.C. (1988), Kash, Jha, and Tsang reply, Physical Review
Letters, 60, 864.
Kash, J.A., Tsang, J.C., and Hvam, J.M. (1985), Subpicosecond time-resolved Raman
spectroscopy of LO phonons in GaAs, Physical Review Letters, 54, 2151–2154.
Keating, P.N. (1966), Theory of the third-order elastic constants of diamond-like crystals,
Physical Review, 149, 674–678.
Kim, K.W. and Stroscio, M.A. (1990), Electron–optical-phonon interaction in binary/ternary
heterostructures, Journal of Applied Physics, 68, 6289–6292.
Kim, K.W., Stroscio, M.A., Bhatt, A., Mickevicius, R., and Mitin, V.V. (1991),
Electron–optical-phonon scattering rates in a rectangular semiconductor quantum wire,
Journal of Applied Physics, 70, 319–327.
Kim, K.W., Bhatt, A.R., Stroscio, M.A., Turley, P.J., and Teitsworth, S.W. (1992), Effects of
interface phonon scattering in multi-heterointerface structures, Journal of Applied
Physics, 72, 2282–2287.
Kirillov, D., Lee, H., and Harris, J.S. (1996), Raman study of GaN films, Journal of Applied
Physics, 80, 4058–4062.
Kisin, M.V., Gorfinkel, V.B., Stroscio, M.A., Belenky, G., and Luryi, S. (1997), Influence of
complex phonon spectra on intersubband optical gain, Journal of Applied Physics, 82,
2031–2038.
References 263
Kisin, M.V., Stroscio, M.A., Belenky, G., and Luryi, S. (1998a), Electron–plasmon relaxation
in quantum wells with inverse subband occupation, Applied Physics Letters, 73 ,
2075–2077.
Kisin, M.V., Stroscio, M.A., Belenky, G., Gorfinkel, V.B., and Luryi, S. (1998b), Effects of
interface phonon scattering in three-interface heterostructures, Journal of Applied
Physics, 83, 4816–4822.
Kitaev, Yu. E., Limonov, M.F., Tronc, P., and Yushkin, G.N. (1998), Raman-active modes in
würtzite (GaN)m (AlN)n superlattices, Physical Review, B57, 14 209–14 212.
Kittel, C. (1976), Introduction to Solid State Physics, 5th edition. John Wiley & Sons, New
York.
Klein, M.V. (1975), Electronic Raman scattering, pp. 147–204, in Light Scattering in Solids,
M. Cardona, ed. Springer-Verlag, Heidelberg.
Klein, M.V. (1986), Phonons in semiconductor superlattices, IEEE Journal of Quantum
Electronics, QE-22, 1760–1770.
Klein, M.V., Ganguly, B.N., and Colwell, P.J. (1972), Theoretical and experimental study of
Raman scattering from coupled LO-phonon–plasmon modes in silicon carbide, Physical
Review, B6, 2380–2388.
Klemens, P.G. (1958), Thermal conductivity and lattice vibrational modes, in Solid State
Physics: Advances in Research and Applications, Vol. 7, eds. F. Seitz and D. Turnbull, pp.
1–98. Academic Press, New York.
Klemens, P.G. (1966), Anharmonic decay of optical phonons, Physical Review, 148, 845–848.
Klemens, P.G. (1975), Anharmonic decay of optical phonons in diamond, Physical Review,
B11, 3206–3207.
Kliewer, K.L. and Fuchs, R. (1966a), Optical modes of vibration in an ionic slab including
retardation. I.: Non-radiative region, Physical Review, 144, 495–503.
Kliewer, K.L. and Fuchs, R. (1966b), Optical modes of vibration in an ionic slab including
retardation. II.: Radiative region, Physical Review, 150, 573–588.
Knipp, P.A. and Reinecke, T.L. (1992), Interface phonons of quantum wires, Physical Review,
B45, 9091–9102.
Komirenko, S.M., Kim, K.W., Stroscio, M.A., and Dutta, M. (1999), Dispersion of polar
optical phonons in würtzite heterostructures, Physical Review, B59, 5013–5020.
Komirenko, S.M., Kim, K.W., Stroscio, M.A., and Dutta, M. (2000a), Energy-dependent
electron scattering via interaction with optical phonons in würtzite crystals and quantum
wells, Physical Review, B61, 2034–2040.
Komirenko, S.M., Kim, K.W., Demidenko, A.A., Kochelap, V.A., and Stroscio, M.A. (2000b),
Cerenkov generation of high-frequency confined acoustic phonons in quantum wells,
Applied Physics Letters, 76, 1869–1871.
Komirenko, S.M., Kim, K.W., Stroscio, M.A., and Dutta, M. (2000c), Applicability of the
Fermi golden rule, Fröhlich polaron in the Feynman model and new perspectives for
possibility of quasiballistic transport in nitrides, to be published.
Kozawa, T., Kachi, T., Kano, H., Taga, Y., Hashimoto, M., Koide, N., and Manabe, K. (1998),
Raman scattering from LO-phonon–plasmon coupled modes in gallium nitride, Journal of
Applied Physics, 75, 1098–1101.
Krauss, Todd D. and Wise, Frank W. (1997), Coherent acoustic phonons in a quantum dot,
Physical Review Letters, 79, 5102–5105.
264 References
Kroto, H.W., Heath, J.R., O’Brien, S.C., Curl, R.F., and Smalley, R.E. (1985), C60 :
Buckminsterfullerine, Nature, 318, 162–163.
Kwon, H.J., Lee, Y.H., Miki, O., Yamano, H., and Yoshida, A. (1996), Raman scattering of
indium nitride thin films grown by microwave-excited metalorganic vapor phase epitaxy
on (0001) sapphire substrates, Applied Physics Letters, 69, 937–939.
Leburton, J.-P. (1984), Size effects on polar optical phonon scattering of 1D and 2D electron
gas in synthetic semiconductors, Journal of Applied Physics, 56, 2850–2855.
Leburton, J.-P. (1997), Dissipation and scattering time engineering in quantum devices, in
Proceedings of the International Conference on Quantum Devices, eds. K. Ismail, S.
Bandyopadhyay, and J.-P. Leburton, pp. 242–252. World Scientific, Singapore.
Lee, B.C., Kim, K.W., Dutta, M., and Stroscio, M.A. (1997), Electron–optical-phonon
scattering in würtzite crystals, Physical Review, B56, 997–1000.
Lee, B.C., Kim, K.W., Stroscio, M.A., and Dutta, M. (1998), Optical phonon confinement and
scattering in würtzite heterostructures, Physical Review, B58, 4860–4865.
Lee, I., Goodnick, S.M., Gulia, M., Molinari, E., and Lugli, P. (1995), Microscopic
calculations of the electron–optical-phonon interaction in ultrathin GaAs/Alx Ga1−x As
alloy quantum-well systems, Physical Review, B51, 7046–7057.
Lemos, V., Arguello, C.A., and Leite, R.C.C. (1972), Resonant Raman scattering of TO(A1 ),
TO(E1 ) and E2 optical phonons in GaN, Solid State Communications, 11, 1351–1353.
Licari, J.J. and Evrard R. (1977), Electron–phonon interaction in a dielectric slab: effect of the
electronic polarizability, Physical Review, B15, 2254–2264.
Loudon, R. (1964), The Raman effect in crystals Advances in Physics, 13, 423–482; erratum,
ibid. (1965), 14, 621.
Lucovsky, G. and Chen, M.F. (1970), Long wave optical phonons in the alloy systems:
Ga1−x Inx As, GaAs1−x Sbx , and InAs1−x Sbx , Solid State Communications, 8,
1397–1401.
Lyubomirsky, I. and Hu, Q. (1998), Energy level schemes for far-infrared quantum well lasers,
Applied Physics Letters, 73, 300–302; see also Dutta, M., and Stroscio, M.A. (1999),
Comment on ‘Energy level schemes for far-infrared quantum well lasers’, Applied
Physics Letters, 74, 2555.
McNeil, L.E., Grimsditch, M., and French, R.H. (1993), Vibrational spectroscopy of aluminum
nitride, Journal of the American Ceramics Society, 76, 1132–1136.
Manchon, D.D., Barker, A.S., Dean, P.J., and Zetterstrom, R.B. (1970), Optical studies of the
phonons and electrons in gallium nitride, Solid State Communications, 8, 1227–1231.
Marcatili, E.J. (1969), Dielectric rectangular waveguide and directional coupler for integrated
optics, Bell Systems Technical Journal, 48, 2071–2102.
M̌arkus, Štefan (1988), The Mechanics of Vibrations of Cylindrical Shells. Elsevier,
Amsterdam.
Menéndez, J. and Cardona, M. (1984), Temperature dependence of the first-order Raman
scattering by phonons in Si, Ge, and α-Sn: anharmonic effects, Physical Review, B29,
2051–2059.
Mitin, V.V., Kochelap, V.A., and Stroscio, M.A. (1999), Quantum Heterostructures:
Microelectronics and Optoelectronics. Cambridge University Press, Cambridge.
References 265
Pfeifer, T., Kutt, W., Kurz, H., and Scholz, H. (1992), Generation and detection of coherent
optical phonons in germanium, Physical Review Letters, 69, 3248–3251.
Platzman, P.M. and Wolff, P.A. (1973), Waves and interactions in solid state plasmas, Solid
State Physics Series, Supplement 13, eds. H. Ehrenreich, Seitz, F., and Turnbull, D.
Academic Press, New York.
Ponce, F.A., Steeds, J.W., Dyer, C.D., and Pitt, G.D. (1996), Direct imaging of
impurity-induced Raman scattering in GaN, Applied Physics Letters, 69, 2650–2652.
Pötz, W. and Schroeder, W.A., eds. (1999), Coherent Control in Atoms, Molecules, and
Semiconductors. Kluwer Academic Publishers, Dordrecht, The Netherlands.
Register, L.F., Stroscio, M.A., and Littlejohn, M.A. (1988), A highly efficient computer
algorithm for evaluating Feynman path-integrals, Superlattices and Microstructures, 6,
233–236.
Register, L.F., Stroscio, M.A., and Littlejohn, M.A. (1991), Conservation law for confined
polar-optical phonon influence functionals, Physical Review, B44, 3850–3857.
Richter, E. and Strauch, D. (1987), Lattice dynamics of GaAs/AlAs superlattices, Solid State
Communications, 64, 867–870.
Ridley, B.K. (1996), The LO phonon lifetime in GaN, Journal of Physics: Condensed Matter,
8, L511–L513.
Ridley, B.K. (1997), Electrons and Phonons in Semiconductor Multilayers. Cambridge
University Press, Cambridge.
Ridley, B.K. and Gupta, R. (1991), Nonelectronic scattering of longitudinal-optical phonons in
bulk polar semiconductors, Physical Review, B43, 4939–4944.
Rücker, H., Molinari, E., and Lugli, P. (1992), Microscopic calculation of the electron–phonon
interactions in quantum wells, Physical Review, B45, 6747–6756.
Ruppin, R., and Engleman, R. (1970), Optical phonons in small crystals, Reports on the
Progress of Physics, 33, 149–196.
Sakaki, H. (1989), Quantum wire superlattices and coupled quantum box arrays: a novel
method to suppress optical phonon scattering in semiconductors, Japanese Journal of
Applied Physics, 28, L314–L316.
Seyler, J. and Wybourne, M.N. (1992), Phonon subbands observed in electrically heated metal
wires, Physics of Condensed Matter, 4, L231–L236.
Scott, J.F., Leite, R.C.C., Damen, T.C., and Shah, J. (1969), Resonant Raman effects in
semiconductors, Physical Review, 188, 1285–1290.
Shah, J., Leite, R.C.C., and Scott, J.F. (1970), Photo-excited hot LO phonons in GaAs, Solid
State Communications, 9, 1089–1093.
Shapiro, S.M. and Axe, J.D. (1972), Raman scattering from polar phonons, Physical Review,
B6, 2420–2427.
Siegle, H., Kaczmarczyk, G., Filippidis, L., Litvinchuk, A.P., Hoffmann, A., and Thomsen, C.
(1997), Zone-boundary phonons in hexagonal and cubic GaN, Physical Review, B55,
7000–7004.
Singh, J. (1993), Physics of Semiconductors and Their Heterostructures. McGraw-Hill, New
York.
Sirenko, Y.M., Kim, K.W., and Stroscio, M.A. (1995), Elastic vibrations of biological and
artificial microtubules and filaments, Electrochemical Society Proceedings, 95-17,
260–271.
References 267
Sirenko, Y.M., Stroscio, M.A., and Kim, K.W. (1996a), Dynamics of cytoskeleton filaments,
Physical Review, E54, 1816–1823.
Sirenko, Y.M., Stroscio, M.A., and Kim, K.W. (1996b), Elastic vibrations of microtubules in a
fluid, Physical Review, E53, 1003–1011.
Sood, A.K., Menéndez, J., Cardona, M., and Ploog, K. (1985), Resonance Raman scattering by
confined LO and TO phonons, Physical Review Letters, 54, 2111–2114.
Sood, A.K., Menéndez, J., Cardona, M., and Ploog, K. (1986), Sood et al. respond, Physical
Review Letters, 56, 1753.
Stroscio, M.A. (1989), Interaction between longitudinal-optical phonon modes of a rectangular
quantum wire with charge carriers of a one-dimensional electron gas, Physical Review,
B40, 6428–6431.
Stroscio, M.A. (1996), Interface-phonon-assisted transitions in quantum well lasers, Journal of
Applied Physics, 80, 6864–6867.
Stroscio, M.A. and Kim, K.W. (1993), Piezoelectric scattering of carriers in confined acoustic
modes in cylindrical quantum wires, Physical Review, B48, 1936–1939.
Stroscio, M.A. and Dutta, M. (1999), Damping of nonequilibrium acoustic phonon modes in
semiconductor quantum dot, Physical Review, B60, 7722–7724.
Stroscio, M.A., Kim, K.W., and Littlejohn, M.A. (1990), Theory of optical-phonon interactions
in rectangular quantum wires, in Proceedings of the Society of Photo-optical
Instrumentation Engineers Conference on Physical Concepts of Materials for Novel
Optoelectronic Device Applications II: Device Physics and Applications, 1362 ed. M.
Razeghi, pp. 556–579.
Stroscio, M.A., Iafrate, G.J., Kim, K.W., Littlejohn, M.A., Goronkin, H., and Maracas, G.
(1991a), Transitions from LO-phonon to SO-phonon scattering in short-period
AlAs–GaAs superlattices, Applied Physics Letters, 59, 1093–1096.
Stroscio, M.A., Kim, K.W., and Rudin, S. (1991b), Boundary conditions for
electron–LO-phonon interaction in polar semiconductor quantum wires, Superlattices and
Microstructures, 10, 55–58.
Stroscio, M.A., Kim, K.W., Iafrate, G.J., Dutta, M., and Grubin, H.L. (1991c), Reduction and
control of inelastic longitudinal-optical phonon scattering in nanoscale and mesoscopic
device structures, in Proceedings of the 1991 International Device Research Symposium,
pp. 87–89, ISBN 1-880920-00-X, University of Virginia Engineering Academic Outreach
Publication, Charlottesville.
Stroscio, M.A., Kim, K.W., Iafrate, G.J., Dutta, M., and Grubin H. (1992), Dramatic reduction
of the longitudinal-optical phonon emission rate in polar-semiconductor quantum wires,
Philosophical Magazine Letters, 65, 173–176.
Stroscio, M.A., Iafrate, G.J., Kim, K.W., Yu, S., Mitin, V., and Bannov, N. (1993), Scattering
of carriers from confined acoustic modes in nanostructures, Proceedings of the 1993
International Device Research Symposium, pp. 873–875, ISBN 1-880920-02–6,
University of Virginia Engineering Academic Outreach Publication, Charlottesville.
Stroscio, M.A., Kim, K.W., Yu, S., and Ballato, A. (1994), Quantized acoustic phonon modes
in quantum wires and quantum dots, Journal of Applied Physics, 76, 4670–4673.
Stroscio, M.A., Sirenko, Yu. M., Yu, S., and Kim, K.W. (1996), Acoustic phonon quantization
in buried waveguides and resonators, Journal of Physics: Condensed Matter, 8,
2143–2151.
268 References
Stroscio, M.A., Kisin, M.V., Belenky, G., and Luryi, S. (1999), Phonon enhanced inverse
population in asymmetric double quantum wells, Applied Physics Letters, 75, 3258–3260.
Sun, C.-K., Liang, J.-C., Stanton, C.J., Abare, A., Coldren, L., and DenBaars, S. (1999), Large
coherent acoustic-phonon oscillations observed in InGaN/GaN multiple-quantum wells,
Applied Physics Letters, 75, 1249–1251.
Sun, H.C., Davis, L., Sethi, S., Singh, J., and Bhattacharya, P. (1993), Properties of a tunneling
injection quantum-well laser: recipe for a ’cold’ device with a large modulation
bandwidth, IEEE Photonics Technology Letters 5, 870–872.
Sun, J.-P., Teng, H.B., Haddad, G.I., Stroscio, M.A., and Iafrate, G.J. (1997), Intersubband
relaxation in step quantum well structures, International Conference on Computational
Electronics, University of Notre Dame, 28–30 May 1997, private communication.
Sung, C.Y., Norris, T.B., Afzali-Kushaa, A., and Haddad, G.I. (1996), Femtosecond
intersubband relaxation and population inversion in stepped quantum well, Applied
Physics Letters, 68 , 435–437; also see the references in this paper.
Teng, H.B., Sun, J.P., Haddad, G.I., Stroscio, M.A., Yu, S., and Kim, K.W. (1998), Phonon
assisted intersubband transitions in step quantum well structures, Journal of Applied
Physics, 84, 2155–2164.
Thompson, C.J. and Blatt, J.M. (1963), Shape resonances in superconductors II: simplified
theory, Physics Letters, 5, 6–9.
Tsen, K.T. (1992), Picosecond time-resolved Raman studies of electron–optical phonon
interactions in ultrathin GaAs–AlAs multiple quantum well structures, Semiconductor
Science and Technology, 7, B191–B194.
Tsen, K.T. and Morkoç, H. (1988a), Picosecond Raman studies of the optical phonons in the
AlGaAs layers of GaAs–AlGaAs multiple-quantum-well structures, Physical Review,
B37, 7137–7139.
Tsen, K.T. and Morkoç, H. (1988b), Subpicosecond time-resolved Raman spectroscopy of LO
phonons in GaAs–Alx Ga1−x As multiple-quantum-well structures, Physical Review, B38,
5615–5616.
Tsen, K.T., Joshi, R.P., Ferry, D.K., and Morkoc, H. (1989), Time-resolved Raman scattering of
nonequilibrium LO phonons in GaAs quantum wells, Physical Review, B39, 1446–1449.
Tsen, K.T., Joshi, R.P., Ferry, D.K., Botchkarev, A., Sverdlov, B., Salvador, A., and Morkoc, H.
(1996), Nonequilibrium electron distributions and phonon dynamics in würtzite GaN,
Applied Physics Letters, 68, 2990–2992.
Tsen, K.T., Ferry, D.K., Botchkarev, A., Sverdlov, B., Salvador, A., and Morkoc, H. (1997),
Direct measurement of electron-longitudinal optical phonon scattering rates in würtzite
GaN, Applied Physics Letters, 71, 1852–1853.
Tsen, K.T., Ferry, D.K., Botchkarev, A., Sverdlov, B., Salvador, A., and Morkoc, H. (1998),
Time-resolved Raman studies of the decay of the longitudinal optical phonons in würtzite
GaN, Applied Physics Letters, 72, 2132–2136.
Tua, P.F. (1981), Lifetime of high-frequency longitudinal-acoustic phonons in CaF2 at low
crystal temperatures, Ph.D. thesis, Indiana University.
Tua, P.F. and Mahan, G.D. (1982), Lifetime of high-frequency longitudinal-acoustic phonons
in CaF2 at low crystal temperatures, Physical Review, B26, 2208–2215.
Turley, P.J., Wallis, C.R., and Teitworth, S.W. (1991a), Phonon-assisted tunneling due to
localized modes in double-barrier structures, Physical Review, B44, 8181–8184.
References 269
Turley, P.J., Wallis, C.R., and Teitworth, S.W. (1991b), Electronic wave functions and
electron-confined-phonon matrix elements in GaAs/Alx Ga1−x As double-barrier resonant
tunneling structures, Physical Review, B44, 3199–3210.
Turley, P.J., Wallis, C.R., and Teitworth, S.W. (1991c), Effects of localized phonon modes on
magnetotunneling spectra in double-barrier structures, Physical Review, B44,
12 959–12 963.
Turley, P.J., Wallis, C.R., and Teitworth, S.W. (1992), Theory of localized phonon modes and
their effects on electron tunneling in double-barrier structures, Journal of Applied
Physics, 76, 2356–2366.
Turley, P.J., Wallis, C.R., Teitworth, S.W., Li, W., and Bhattacharya, P.K. (1993), Tunneling
measurements of symmetric-interface phonons in GaAs/AlAs double-barrier structures,
Physical Review, B47, 12 640–12 648.
Vogl, P. (1980), The electron–phonon interaction in semiconductors, in Physics of Nonlinear
Transport in Semiconductors, Proceedings of the NATO Advanced Study Institute
Seminar, eds. D.K. Ferry, J.R. Barker, and C. Jacoboni. Plenum, New York.
von der Linde, D., Kuhl, J., and Klingenberg, H. (1980), Raman scattering from
nonequilibrium LO phonons with picosecond resolution, Physical Review Letters, 44,
1505–1508.
Wang, Jin, Leburton, J.-P., Moussa, Z., Julien, F.H., and Sa’ar, A. (1996a), Simulation of
optically pumped mid-infrared intersubband semiconductor laser structures, Journal of
Applied Physics, 80, 1970–1978.
Wang, Jin, Leburton, J.-P., Julien, F.H., and Sa’ar, A. (1996b), Design and performance
optimization of optically-pumped mid-infrared intersubband semiconductor lasers, IEEE
Photonics Technology Letters, 8, 1001–1003.
Wang, X.F. and Lei, X.L. (1994), Polar-optic phonons and high-field electron transport in
cylindrical GaAs/AlAs quantum wires, Physical Review, B49, 4780–4789.
Waugh, J.L.T. and Dolling, G. (1963), Crystal dynamics of gallium arsenide, Physical Review,
132, 2410–2412.
Wendler, L. (1985), Electron–phonon interaction in dielectric bilayer systems: effects of the
electronic polarizability, Physica Status Solidi B, 129, 513–530.
Wendler, L. and Grigoryan, V.G. (1988), Acoustic interface waves in sandwich structures,
Surface Science, 206, 203–224.
Wetzel, C., Walukiewicz, W., Haller, E.E., Ager, III, J., Grzegory, I., Porowski, S., and Suski,
T. (1996), Carrier localization of as-grown n-type gallium nitride under large hydrostatic
pressure, Physical Review, B53, 1322–1326.
Wisniewski, P., Knap, W., Malzak, J.P., Camassel, J., Bremser, M.D., Davis, R.F., and Suski, T.
(1998), Investigation of optically active E1 transversal optic phonon modes in
Alx Ga1−x N layers deposited on 6H-SiC substrates using infrared reflectance, Applied
Physics Letters, 73, 1760–1762.
Worlock, J.M. (1985), Phonons in superlattices, in Proceedings of the Second International
Conference on Phonon Physics, J. Kollár, N. Kroó, N. Menyhárd, and T. Siklós., eds., pp.
506–520. World Scientific, Singapore.
Xu, B., Hu, Q., and Mellock, M.R. (1997), Electrically pumped tunable terahertz emitter based
on intersubband transition, Applied Physics Letters, 71, 440–442.
Yafet, Y. (1966), Raman scattering by carriers in Landau levels, Physical Review, 152,
858–862.
270 References
271
272 Index
dielectric continuum model, 52, 60, 93 gain in intersubband laser, 195, 219
dielectric function, 13 Grüneisen constant, 47
dielectric polarizability, 224 guided modes, 225
dilatation, of medium, 58
dilatational modes, 99 half-space optical phonon modes, 152
dilatational solutions, 60 Hamiltonian for harmonic oscillator, 35
dilatational waves, 98 harmonic interactions, 8
dimensional confinement, 39 harmonic modes, 45
dispersion curves, in würtzites 32, 77, 78 harmonic oscillator, 35
dispersion relations, 8, 26 hexagonal würtzite structures, 53
displacement, 8 high-temperature electronics, 54
displacement eigenmodes, 98 Hooke’s law, 8, 56
displacement field, quantized, 40 hot electron distribution, 204
dissipative mechanisms, 220 hot-phonon-bottleneck effect, 164
distortional solutions, 60 hot phonon decay, 3
double-barrier heterostructure, 88 hot phonons, in polar quantum wires, 163
drifted Fermi distribution, 216 Huang–Born equations, 20, 54, 221
driven-oscillator equation, 14, 63 Huang–Zhu modes, 66, 228
dynamical screening, in polar quantum wires, 162,
165 impurity, 28
in-gap, 27
E 1 mode, 17 inhomogeneous broadening effects, 31
effective charge, 53 infrared-active modes, 20
elastic continuum model, 56, 176 infrared-active phonons, 138
elastic continuum theory, 60 LO-like, 138
elasticity, theory of, 47 TO-like, 138
electron–acoustic-phonon scattering interface disorder, 28
in cylindrical quantum wire, 176 interface modes
in rectangular quantum wire, 112 for optical phonons, 66, 68, 80, 83, 152
in slab, 61
electron permittivity, 215
optical phonon interaction Hamiltonian, 65
electron–phonon interaction, 30
intersubband lasers, 196, 219
for slab modes, 66
intersubband scattering rates, 91
electronic polarizability, 68
intrasubband transition rates, 187
electrostatic boundary conditions, 66, 79, 152
interwell phonon-assisted transition, 207
energy-conserving delta function, 142
intrasubband scattering rates, 91
energy loss rate, 48
inversion symmetry, 43
energy–wavevector relationship, 97
ionic bonding, 7
equivoluminal solutions, 60
irrotational solutions, 60
Euler–Lagrange equations, 226
isotopic mass, 33
excitonic states, 27
isotropic medium, 43
extraordinary waves, 19
Kane wavefunction, 196
face-centered cubic lattices, 7 Klemens’ channel, 2, 46, 163
femtosecond lasers, 219
Fermi golden rule, 39, 47 Lagrangian density, 225, 230
flexural modes, 100, 123 Lamé’s constants, 58, 101, 118
flexural thickness acoustic modes, in quantum dot, Landau fans, 203
128 lifetimes, longitudinal optical phonons, 3, 49
flexural waves, 98 linear-chain model, 7
folded acoustic modes, in semiconductor localized acoustic modes, 214
superlattices, 60 longitudinal acoustic mode, 9, 17
force equations, 59, 118 longitudinal electromagnetic disturbance, 66
Fourier decomposition, 38 longitudinal electromagnetic wave, 14
free-standing cylindrical structure, 180 longitudinal optical mode, 9, 17
free-surface boundary condition, 177 longitudinal solutions, 60
form factor, 232 longitudinal sound speed, 57, 59
Fröhlich Hamiltonian, for two-dimensional slab, 69 in thin plate, 119
Fröhlich interaction, 31, 40 Loudon model, 14, 15, 23, 54
Fröhlich interaction Hamiltonian, 41 Lyddane–Sachs–Teller relation, 13, 15, 52, 80, 221
for polar uniaxial crystal, 136
for quantum box, 168 macroscopic theory of polar modes, 14
for two-dimensional slab, 140 magnetotunneling spectra, 202
Index 273