ArticlePDF Available

Strain distribution and optical phonons in InAs/InP self-assembled quantum dots

Authors:

Abstract and Figures

The strain distribution in self-assembled InAs/InP (001) quantum dots is calculated, using an atomistic valence force-field description. Two typical dot shapes are considered. Strain relaxation is found to depend much on the dot shape. From these modeling results we deduce the strain-induced phonon frequency shifts. Unlike confinement, strain induces large frequency shifts. The calculations agree well with experimental results obtained by Raman scattering. It is shown that alloying effects are small. Finally, we show that average strain values can be obtained experimentally if one combines longitudinal and transverse optical-phonon Raman scattering.
Content may be subject to copyright.
Strain distribution and optical phonons in InAs/InP self-assembled quantum dots
J. Groenen
Laboratoire de Physique des Solides, ESA 5477, Universite
´
P. Sabatier, F-31062 Toulouse Cedex 4, France
C. Priester
IEMN, De
´
partement ISEN, CNRS-UMR 8520, BP 69 F-59652 Villeneuve d’Ascq Cedex, France
R. Carles
Laboratoire de Physique des Solides, ESA 5477, Universite
´
P. Sabatier, F-31062 Toulouse Cedex 4, France
Received 14 June 1999
The strain distribution in self-assembled InAs/InP 001 quantum dots is calculated, using an atomistic
valence force-field description. Two typical dot shapes are considered. Strain relaxation is found to depend
much on the dot shape. From these modeling results we deduce the strain-induced phonon frequency shifts.
Unlike confinement, strain induces large frequency shifts. The calculations agree well with experimental results
obtained by Raman scattering. It is shown that alloying effects are small. Finally, we show that average strain
values can be obtained experimentally if one combines longitudinal and transverse optical-phonon Raman
scattering. S0163-18299910347-3
I. INTRODUCTION
Self-assembled quantum dot structures have been attract-
ing considerable attention in the past decade. Dots are ob-
tained during heteroepitaxy as a result from the elastic relax-
ation of misfit strain. Several studies were devoted to the
modeling of the strain inside the dots.
1–6
Calculations were
performed using either elastic continuum theory finite ele-
ment model
1–3
or an atomistic description valence force
field model VFF兲兴.
4–6
The strain field was shown to depend
strongly on dot shape. Thanks to the strain simulations, the
quantitative analysis of many experimental data has become
possible. In particular, they are very helpful to understand
the electronic properties.
2,3,5,6
Confinement effects were
shown see, for instance, Ref. 6 to depend strongly on dot
size and strain and thus on the dot shape.
Although phonons are efficient probes for investigating
low-dimensional structures, to date little work has been done
on phonons in self-assembled nanostructures.
2,7–16
Most of
the theoretical and experimental work on phonons in quan-
tum dots deals with unstrained systems. Only a few Raman
scattering investigations of self-assembled nanostructures
have been reported.
9–16
Valuable information about strain,
alloying, electronic properties, ...)hasbeen accessed using
resonant Raman scattering. In particular, Raman scattering
was shown to provide a means of determining independently
the residual strain and the alloy composition in SiGe/Si self-
assembled dots.
13
In this paper we shall address two issues. i How is the
strain distributed in InAs/InP 001 self-assembled dots? ii
How does it modify the optical-phonon frequencies? To an-
swer the first question, we use the VFF model.
17
A large
simulation cell is considered, in order to account for both the
strain distribution inside and around the dots. In contrast
with previous simulations of strain in capped dots, we shall
examine capped dots with truncated pyramidal shapes. The
strain-induced optical-phonon frequency shifts will be de-
rived from the calculated strain. We shall finally compare
these calculations to experimental results obtained by means
of resonant Raman scattering on InAs/InP 001 self-
assembled dots. In contrast with InAs/GaAs or
InP/In
x
Ga
1 x
P systems, the gap between the InAs and InP
optical-phonon frequencies is large. Consequently, the con-
finement of the InAs optical phonons inside the dots is very
efficient, providing us with local probes. Moreover, the dot-
related features can therefore be easily identified in the Ra-
man spectra.
II. CALCULATIONS
A. Strain distribution
It should be noted that the accuracy of the VFF model
goes beyond classical elasticity theory as it decribes the elas-
tic properties and the relaxation on the atomic scale.
18
The
strain elastic energy
E depends on the geometric deforma-
tions of bonds that each atom makes with its four nearest
neighbors. For each atom i of the zinc-blende structure, one
can write
E
i
j 1
4
3
8r
0
2
r
ij
2
r
0
2
2
j 1
4
k j 1
4
3
8r
0
2
r
ij
r
ik
r
0
2
3
2
.
1
r
ij
is the vector connecting the central atom i to one of its
four nearest neighbors j; r
0
is the unstrained bond length;
and
are, respectively, bond-bending and bond-stretching
elastic constants. They are related to the elastic constants
c
11
, c
12
, and c
44
of the continuum elasticity theory by the
following expressions:
c
11
3
a
, c
12
a
, c
44
4
␣␤
a
. 2
PHYSICAL REVIEW B 15 DECEMBER 1999-IVOLUME 60, NUMBER 23
PRB 60
0163-1829/99/6023/160135/$15.00 16 013 ©1999 The American Physical Society
a 4r
0
/
3 is the lattice constant. It is not possible to per-
fectly fit all three c
ij
’s with only two elastic constants
and
; the less ionic is the material, the better is the fit. Thus, this
method basically works with covalent bonding, but Coulomb
corrections could be introduced.
19
In practice, for the sake of
simplicity one usually simply uses Eq. 1.
20
This we do, and
choose to fit c
11
and c
12
and to drop c
44
, resulting in a
1020 % error range on the VFF effective c
44
.
We examine two typical InAs/InP 001 dot morpholo-
gies, corresponding to the ones already reported in Refs. 10
and 21. The misfit strain between InAs and InP equals
3.1%. The first morphology, labeled A, corresponds to 3
nm high and 25 nm wide dots, whereas the second one,
labeled B, corresponds to 7 nm high and 45 nm wide dots.
The dots have truncated pyramidal shapes, with 114 and
113 side facets for A and B, respectively. InAs islands are
formed on the top of a 1.5 monolayer ML wetting layer
WL and capped by a 25 nm InP layer.
10,21
For the sake of simplicity, we simulate pyramids with
square base the InAs/InP islands are in fact slightly elon-
gated along
11
¯
0
Ref. 21, and periodic boundary condi-
tions in the plane perpendicular to 001 are used. At the
bottom of the modelized cell, atoms are kept fixed, in order
to simulate the thick substrate. We calculate the atomic po-
sitions that minimize the total elastic energy. Once the posi-
tions of all atoms are known, the local deformation distribu-
tion is derived straightforwardly. Calculations have been
performed with and without WL 1 or 2 ML thick. Concern-
ing the strain field in the dots, no significant differences were
observed. The results we present here were obtained disre-
garding the WL. It has been shown that the subtrate and the
cladding layer are affected by the strain relaxation within the
dot. In particular, the strain field penetrates deeply into the
subtrate.
4
Consequently, to obtain a reliable and realistic
strain field, a rather thick nonfrozen substrate layer has to be
considered in the simulation. For that purpose, calculations
have been performed with up to 600 000 atoms and about
95% of the simulation cell corresponds to InP. A small scal-
ing factor 1.6 for A and 5 for B) remains between the actual
pyramid size and the one used in the simulation. One can
avoid this scaling factor but, in counterpart, one has to re-
duce the substrate layer thickness. We have checked, on
smaller systems, the validity of using such a scaling factor,
which is bound to the fact that the strain field depends much
on shape and not on size.
Typical local deformation distributions within the dots
along lines defined in Fig. 1are shown in Figs. 2 and 3 for
shape A. Except close to the pyramid boundaries, the strain
field is rather uniform and does not vary very rapidly. The
shear strain
ij
(i j) turns out to be significant at the facet
edges and pyramid boundaries and very small inside the
pyramid Fig. 3. Let us point out that the previous simula-
tions of strain in capped islands all correspond to untruncated
pyramids. Unlike for capped untruncated pyramids,
xx
and
yy
never change sign inside the pyramid.
2,5
Figure 2 clearly shows that the strain field penetrates deep
into the InP barriers. At the surface of the InP capping layer,
some strain is still present just above the dot compare lines
a and b). As this tensile strain is rather localized within the
xOy plane, it is able to promote vertical order when several
layers with dots are grown.
22
From the numerical local deformation values of all the
InAs cells, we have computed the average strain within the
FIG. 1. Schematic plot of a type-A pyramid. The origin O is the
center of the pyramid base plane. x
110
, y
11
¯
0
, and z
001
.
FIG. 2. Local deformations along the a, b, and c lines defined in
Fig. 1:
xx
solid lineand
zz
dashed line. z positions are normal-
ized with respect to the pyramid height (z 0 stands for the pyra-
mid bottom and z1 for the pyramid top. S denotes the sample
surface.
FIG. 3. Local deformations along the lines defined in Fig. 1: 1
solid line,2dashed line, and 3 dotted line. y positions are
normalized with respect to half of the pyramid base width (y 0
corresponds to the middle.
xy
and
xz
are not reported here as they
equal almost zero regardless of the y position.
xx
are not reported
either; they display almost constant values, given by
xx
yy
at
y0.
16 014 PRB 60
J. GROENEN, C. PRIESTER, AND R. CARLES
InAs dot. The average values of the diagonal strain compo-
nents in the InAs dot are reported in Table I the values
corresponding to a pseudomorphic two-dimensional 2D
layer are also given as a reference. According to symmetry
requirements, the average values of the shear strain compo-
nents vanish. The biaxial strain relationship,
zz
/
xx
⫽⫺2C
12
/C
11
, has often been assumed to be valid for flat
islands. It is noteworthy that, even for the rather flat islands
examined here, this relationship does not hold Table I. Re-
sidual strain depends much on shape. The B-type dots are
more relaxed than the A-type ones their height/width ratio
equals 0.155 and 0.12, respectively.
B. Optical-phonon spectra
One could, in principle, obtain the vibrational eigenmodes
from the diagonalization of the dynamical matrix once the
relaxation procedure described above has been performed.
In polar materials, both short-range interaction covalent
bonding and long-range Coulomb interaction have to be
taken into account. Such calculations have been performed
recently for free-standing GaP dots with up to 2000 atoms.
23
In our case, this procedure is, however, impractical: our sys-
tem, which includes necessarily a dot and a large part of the
matrix, contains too many atoms with regard to the simula-
tion capabilities.
Moreover, the dot vibrational eigenmodes are collective
excitations, involving all the atoms belonging to the dot. As
an approximation, we shall therefore consider that the
phonons experience the average strain field inside the dots.
Owing to the rather homogeneous strain field inside the dots
Figs. 2 and 3, this should provide us with reasonable re-
sults. The frequencies of the optical phonon in presence of
strain can be derived from the secular equation given in Ref.
24. The frequency shifts depend on the strain tensor
ij
and
the phonon deformation potentials K
˜
ij
. According to our
modeling, the average shear strain components can be disre-
garded. As
xx
yy
, the strain splits the optical phonons
into a singlet and a doublet component.
24,25
Their relative
frequency shifts are given by
0
S
1
2
K
˜
12
xx
yy
1
2
K
˜
11
zz
, 3
0
D
1
2
K
˜
11
xx
1
2
K
˜
12
yy
zz
. 4
The vibrations are along 001 for the singlet mode and in
the plane normal to 001 for the doublet modes.
25
Notice
that, depending on whether the longitudinal optical LO or
transverse optical TO deformation potentials are used, the
resolution of the secular equation provides us either with the
LO singlet and the LO doublet or with the TO singlet and the
TO doublet. Concerning Fig. 6 in Ref. 2, let us indicate that
the peaks assigned to LO and TO correspond in fact to the
histogram of the LO singlet and the LO doublet relative
changes in phonon energy. One can easily identify three
peaks for the dot: two having a similar location around 7%
i.e., the doublet and another one around 10% i.e., the sin-
glet.
According to Aoki et al.,
27
(K
˜
11
2K
˜
12
)
LO
InAs
⫽⫺6.4 and
(K
˜
11
2K
˜
12
)
TO
InAs
⫽⫺7.3. According to Yang et al.,
28
(K
˜
11
K
˜
12
)
TO
InAs
0.51. Unfortunately, (K
˜
11
K
˜
12
)
LO
InAs
has not
been measured. The corresponding TO value has been used
in most of the previous calculations. One has, however, to
note that in III-V compounds, (K
˜
11
K
˜
12
)
LO
is usually larger
than (K
˜
11
K
˜
12
)
TO
.
26
On the other hand, Tran et al. did in-
vestigate strained InAs/InP superlattices combining Raman
scattering and x-ray diffraction.
29
Their data can therefore be
used to estimate (K
˜
11
K
˜
12
)
LO
InAs
. One obtains (K
˜
11
K
˜
12
)
LO
InAs
0.92, which is about twice as large as the corre-
sponding TO value thus, one observes the same trend as for
GaAs and InP Ref. 26兲兴. One finally obtains K
˜
11
⫽⫺1.50
and K
˜
12
⫽⫺2.43 for the InAs LO and K
˜
11
⫽⫺2.09 and K
˜
12
⫽⫺2.60 for the InAs TO.
30
According to measurements per-
formed at room temperature on InAs 111,
0
equals
239.8 cm
1
and 218.8 cm
1
for the InAs LO and TO, re-
spectively.
In backscattering geometry from the 001 surface, only
the LO singlet is Raman-active. The calculated strain-
induced frequency shifts of the LO singlet are reported in
Table II. The shift expected for a pseudomorphic 2D layer is
also given. Although the strain relaxation in these systems is
quite different, the calculated strain-induced frequency shifts
Eq. 3兲兴 are rather similar.
The island heights are small and confinement may modify
the phonon frequencies. As the island widths are much
larger, the effects of lateral confinement on the phonon fre-
quencies can be neglected. From the island heights and the
optical-phonon dispersion relation
29
i.e., applying the usual
linear chain model
25
, we have calculated the confinement
induced frequency shifts
conf
. The values of
conf
de-
duced in this way for the first-order confined mode are re-
ported in Table II;
conf
is found to be very small for the 3
and 7 nm high islands. In our case, the frequency changes
related to dot size fluctuations are thus also small.
III. COMPARISON WITH EXPERIMENT
AND DISCUSSION
We present Raman spectra ofa2MLInAs single quan-
tum well sample SQW and a sample with A-type dots. De-
TABLE I. Average strain components for A and B dot shapes.
2D stands for a pseudomorphic InAs layer on InP.
Shape
xx
yy
zz
zz
/
xx
2D 3.1 3.1 3.4 1.1
A 2.81 2.80 2.67 0.95
B 2.62 2.60 2.18 0.83
TABLE II. Strain (S singlet) and confinement conf induced
LO frequency shifts in cm
1
) for A and B dot shapes. ‘‘2D’’
stands for a pseudomorphic 2 ML InAs layer in InP. ‘‘exp’’ denotes
experimental data.
Shape
S
conf
Sconf
exp
2D 12 48 9
A 11.7 0.2 11.5 12.1
B 11.3 0 11.3 11.8
PRB 60
16 015STRAIN DISTRIBUTION AND OPTICAL PHONONS IN . . .
tails about the sample growth can be found in Refs. 10 and
21. The spectra were recorded at room temperature with an
XY Dilor spectrometer equipped with a cooled charge
coupled device detector. Depending on the excitation energy
used incoming or outgoing resonance, the Raman spectra
display InAs confined phonon or/and interface IF mode
peaks. As we intend to discuss the effects of strain on con-
fined optical phonons, we only reported here Fig. 4spectra
displaying confined LO-related peaks scattering by TO
phonons is forbidden in backscattering geometry from 001
surfaces. The krypton laser lines we used are in incoming
resonance with either the SQW or the dot InAs E
1
-like tran-
sition 520.8 nm and 482.5 nm for Figs. 4a and 4b,
respectively.
10
We have been able to discriminate between
the island-related signal and the WL-related one using dif-
ferent polarization configurations and the 2 ML SQW sample
as a reference.
10
The InAs LO peaks related to the WL and
the islands are observed in the crossed-polarization and
parallel-polarization configurations, respectively. The InP
substrate LO peak is observed in the crossed-polarization
configuration. The peak observed in the parallel-polarization
configuration in the InP frequency range is attributed to a
symmetric InP-like interface mode IF.
10,29
Despite the dot
size fluctuations and the inhomogeneous strain fields, one
observes rather sharp Raman lines Fig. 4. This is likely due
to i the weak dependence on the dot size of both
S
and
conf
and iithe fact that the phonons are collective vibra-
tional modes the inhomogeneous strain field inside the dot
therefore does not induce significant line broadening.
The LO frequency shifts are reported in Table II. It is
noteworthy that the WL and the 2 ML SQW LO frequencies
are similar Fig. 4. The calculations
S conf
S
conf
account rather well for the experimental data.
Moreover, using grazing incidence, we were able to observe
the InAs TO doublet peak for dots with shape A). Its fre-
quency is shifted up by 7.2 cm
1
with respect to the bulk
InAs TO frequency. From the average strain components for
shape A) and the TO deformation potentials, we obtain Eq.
4兲兴
6.8 cm
1
, which is in good agreement with ex-
periment due to the very weak TO dispersion, confinement
effects are negligible.
Notice that we did not take into account in our simula-
tions that the dots are sometimes slightly elongated one ex-
pects a little less strain relaxation and higher phonon fre-
quencies. However, the main actual limitation we are
concerned with for either the comparison between the cal-
culations and the experimental results or the experimental
determination of the average strain valuesis due to our poor
knowledge of the phonon deformation potentials. The latter
are indeed difficult to measure accurately.
It is noteworthy that if one is able to measure both the dot
LO and TO frequencies, one can deduce the
xx
and
zz
average values by considering simultaneously Eqs. 3 and
4兲共without any numerical strain simulation. Considering
the experimental LO and TO frequency shifts reported here
for A-type dots Table II, one obtains
xx
⫽⫺2.92%,
zz
2.82%, and
zz
/
xx
⫽⫺0.93; these values are very close
to the ones predicted by the simulation Table I. The relative
differences between the experimental and calculated values
do not exceed 4%. This good agreement supports the as-
sumptions we made in particular the one concerning the
phonons probing the average strain field. Combining LO
and TO Raman scattering provides thus a means of measur-
ing average strain values in self-assembled dots.
The rather good agreement between the calculations and
the experimental data suggests that the assumptions we made
are reasonable. It also suggests that alloying inside the is-
lands is not important. Optical phonons in InAsP alloys dis-
play the usual two-mode behavior: the InAs-like InP-like
LO frequencies decrease with decreasing In P content.
31
If
one considers InAsP dots instead of InAs dots in InP, the
calculations yield lower phonon frequencies due to the al-
loying induced frequency shift and the lower mismatch with
respect to InP which obviously do not account for the ex-
perimental data Table II. Moreover, it has been shown that
the formation of an intermediate InAsP alloy layer during the
growth of InAs/InP structures gives rise to additional inter-
face modes.
32
We do not observe the corresponding features
in the Raman spectra.
One may wonder whether one can find some evidence in
the Raman spectra for the strain inside the InP barriers. As
most of the InP LO Raman signal originates from unstrained
regions, we are not able to identify the contributions of the
strained InP. The shift one expects from the tensile strain
underneath the InAs dot Fig. 2and the InP LO deformation
potentials
33
is very small ( 0.5 cm
1
). On the other hand,
one expects the InP-like IF to be sensitive to the strain
around the islands.
14
However, as the IF frequencies depend
also on both the island size and shape,
23,34,35
it is not obvious
to obtain from the Raman spectra some reliable and quanti-
tative information concerning the strained InP.
IV. CONCLUSION
In summary, we calculated the strain distribution in typi-
cal self-assembled InAs/InP 001 dots using the valence-
FIG. 4. Raman spectra: a 2 ML InAs/InP SQW, b type A
dots, corresponding to 2.5 ML InAs deposited on InP and capped
by 25 nm InP. Spectra were recorded with the z(X,Y)z
¯
crossed-
polarization configuration solid line and with the z(X,X)z
¯
parallel-polarization configuration dashed line, with X
100
and
Y
010
. The dot-dashed line indicates the bulk InAs LO fre-
quency.
16 016 PRB 60
J. GROENEN, C. PRIESTER, AND R. CARLES
force-field method. The residual strain was shown to depend
much on the dot shape. Even for rather flat dots, the average
strain field is quite different from the 2D case. As the strain
field penetrates deeply into the subtrate, a large part of the
simulation cell has to be devoted to the subtrate.
We calculated the strain-induced frequency shifts, assum-
ing the InAs phonons experience the average strain field in-
side the dot. It is shown that confinement does not modify
much the dot phonon frequencies. Our calculations show
good agreement with experimental results obtained by Ra-
man scattering. Both the Raman spectra and the comparison
between the calculated frequencies and the experimental val-
ues indicate that alloying effects are small.
Unlike electronic spectra, the dot phonon frequencies do
not depend much on dot size. One can therefore analyze the
dot phonon frequencies by solely considering the residual
strain, and vice versa.
1
S. Christiansen, M. Albrecht, H. P. Strunk, and H. J. Maier, Appl.
Phys. Lett. 64, 3617 1994.
2
M. Grundmann, O. Stier, and D. Bimberg, Phys. Rev. B 52,
11 969 1995.
3
T. Benabbas, P. Francois, Y. Androussi, and A. Lefebvre, J. Appl.
Phys. 80, 2763 1996.
4
Y. Androussi, P. Franc¸ois, A. Lefebvre, C. Priester, I. Lefebvre,
G. Allan, M. Lannoo, J. M. Moison, N. Lebouche, and F. Bar-
the, in Evolution of Thin-Film and Surface Structure and Mor-
phology, edited by B. G. Demczyk, E. D. Williams, E. Gar-
funkel, B. M. Clemens, and J. E. Cuomo, Materials Research
Society, Symposia Proceedings No. 355 MRS, Pittsburgh,
1995, p. 569; C. Priester, I. Lefebvre, G. Allan, and M. Lannoo,
in Mechanisms of Thin Film Evolution, edited by S. M. Yalis-
ove, C. V. Thompson, and D. J. Eaglesham, Materials Research
Society Symposia Proceedings No. 317 MRS, Pittsburgh,
1994, p. 131.
5
H. Jiang and J. Singh, Phys. Rev. B 56, 4696 1997.
6
J. Kim, L.-W. Wang, and A. Zunger, Phys. Rev. B 57, R9408
1998, and references therein.
7
R. Heitz, M. Grundmann, N. N. Ledentsov, L. Eckey, M. Veit, D.
Bimberg, V. M. Ustinov, A. Yu Egorov, A. E. Zhukov, P. S.
Kop’ev, and Zh. I. Alferov, Appl. Phys. Lett. 68, 361 1996.
8
S. Fafard, R. Leon, D. Leonard, J. L. Merz, and P. M. Petroff,
Phys. Rev. B 52, 5752 1995.
9
B. R. Bennet, B. V. Shanabrook, and R. Magno, Appl. Phys. Lett.
68, 958 1996.
10
J. Groenen, A. Mlayah, R. Carles, A. Ponchet, A. Le Corre, and
S. Salau
¨
n, Appl. Phys. Lett. 69, 943 1996.
11
G. Armelles, T. Utzmeier, P. A. Postigo, J. C. Ferrer, P. Peiro
´
,
and A. Cornet, J. Appl. Phys. 81, 6339 1997.
12
P. D. Persans, P. W. Deelman, K. L. Stokes, L. J. Schowalter, A.
Byrne, and T. Thundat, Appl. Phys. Lett. 70, 472 1997.
13
J. Groenen, R. Carles, S. Christiansen, M. Albrecht, W. Dorsch,
H. P. Strunk, H. Wawra, and G. Wagner, Appl. Phys. Lett. 71,
3856 1997.
14
Y. A. Pusep, G. Zanelatto, S. W. da Silva, J. C. Galzerani, P. P.
Galzerani, P. P. Gonzalez-Borrero, A. I. Toropov, and P. Bas-
maji, Phys. Rev. B 58, 1770 1998.
15
S. H. Kwok, P. Y. Yu, C. H. Tung, Y. H. Zhang, M. F. Li, C. S.
Peng, and J. M. Zhou, Phys. Rev. B 59, 4980 1999.
16
A. A. Sirenko, M. K. Zundel, T. Ruf, K. Eberl, and M. Cardona,
Phys. Rev. B 59, 4980 1999.
17
M. J. P. Musgrave and J. A. Pople, Proc. R. Soc. London, Ser. A
268, 464 1962.
18
P. N. Keating, Phys. Rev. 145, 637 1966.
19
R. M. Martin, Phys. Rev. B 1, 4005 1970.
20
J. L. Martins and A. Zunger, Phys. Rev. B 30, 6217 1984.
21
A. Ponchet, A. Le Corre, H. L’Haridon, B. Lambert, and S. Sa-
lau
¨
n, Appl. Phys. Lett. 67, 1850 1995.
22
See, for instance, Q. Xie, A. Madhukar, P. Chen, and N. Koba-
yashi, Phys. Rev. Lett. 75, 2542 1995.
23
H. Fu, V. Ozolin
¨
s, and A. Zunger, Phys. Rev. B 59, 2881 1999.
24
F. Cerdeira, C. J. Buchenauer, F. H. Pollak, and M. Cardona,
Phys. Rev. B 5, 580 1972.
25
B. Jusserand and M. Cardona, in Light Scattering in Solids V,
Topics in Applied Physics Vol. 66, edited by M. Cardona and G.
Gu
¨
ntherodt Springer, New York, 1989.
26
See Table 3.1 in Ref. 25; (K
˜
11
K
˜
12
)
LO
/(K
˜
11
K
˜
12
)
TO
equals 2.3
and 1.7 for GaAs and InP, respectively.
27
K. Aoki, E. Anastassakis, and M. Cardona, Phys. Rev. B 30, 681
1984.
28
M. J. Yang, R. J. Wagner, B. V. Shanabrook, W. J. Moore, J. R.
Waterman, C. H. Yang, and M. Fatemi, Appl. Phys. Lett. 63,
3434 1993.
29
C. A. Tran, J. L. Bredner, R. Leonelli M. Jouanne, and R. A.
Masut, Phys. Rev. B 49, 11 268 1994; Superlattices Micro-
struct. 15, 391 1994.
30
Notice that if one calculates the strain-induced LO frequency
shifts using the (K
˜
ij
)
LO
derived from (K
˜
11
2K
˜
12
)
LO
and (K
˜
11
K
˜
12
)
TO
instead of the (K
˜
11
K
˜
12
)
LO
value we deduced from
Ref. 29, one would obtain shifts which are typically 25% lower
than those reported here Table II.
31
R. Carles, N. Saint-Cricq, J. B. Renucci, and R. J. Nicholas, J.
Phys. C 13, 899 1980.
32
L. G. Quagliano, B. Jusserand, and D. Orani, Phys. Rev. B 56,
4919 1997.
33
E. Anastassakis, Y. S. Raptis, M. Hunermann, W. Richter, and M.
Cardona, Phys. Rev. B 38, 7702 1988.
34
P. A. Knipp and T. L. Reinecke, Phys. Rev. B 46,103101992.
35
M. P. Chamberlain, C. Trallero-Giner, and M. Cardona, Phys.
Rev. B 51, 1680 1995.
PRB 60
16 017STRAIN DISTRIBUTION AND OPTICAL PHONONS IN . . .
... The strain also exerts an influence on the vibrational properties of such materials. Raman spectroscopy and IR spectroscopy offer effective tools for studying the strain in semiconductors and semiconductor nanostructures [10][11][12][13][14][15][16][17][18][19][20][21][22][23]. It is known that a biaxial stress alters the symmetry of zinc-blende-and diamond-type crystals, leading to a splitting of optical phonons at the center of the Brillouin zone into singlet and doublet modes [11]. ...
... Fig. 4 shows the Raman spectra of InAlAs QDs registered using different scattering geometries in the range of wavenumbers typical of In-As optical vibrations. Two peaks are observed around 220 cm − 1 and 250 cm − 1 , which were previously attributed to light scattering by TO and LO phonons in In x Al 1-x As alloys [17,18,25,38]. It is well known that vibrations in solid alloys are localized vibrations, and they cannot be considered as purely longitudinal or purely transverse modes. ...
... Anyway, in this case the splitting between TO Z and TO Y' phonons can be observed as well. It is known that the relation between the strains ε zz and ε xx in QDs differ from this relation in the flat layers of In x Al 1-x As [17]. As a result, it becomes impossible to accurately evaluate the strain in this case. ...
Article
The splitting of long-wave optical phonons that arises in biaxially stressed zinc-blende- and diamond-type crystals due to violation of cubic symmetry in such crystals was studied both theoretically and experimentally (using micro-Raman technique). The anglular dispersion of optical phonons near the center of the Brillouin zone was calculated for split modes in biaxially stressed (001)-oriented films. The results obtained were used to analyze the effects due to strain and alloying on optical-phonon frequencies in stressed Ge, InGaAs and InAlAs films grown on GaAs (001) substrates. The developed approach permits a more precise determination of both the composition and the biaxial strain in AIIIBV alloy films based on an analysis of Raman spectra taken from such films.
... This indicates that phonon modes of interest are related to the InP matrix. The LO phonon energy of bulk InP of 30 meV lies within this range [128]. The large distribution of obtained values is related to the fact that in the experiment, broadening does not originate only from the phonon effects but also from the spectral diffusion. ...
... quant [127]. This indicates that phonon modes of interest are related to the InP mat phonon energy of bulk InP of 30 meV lies within this range [128]. The large d of obtained values is related to the fact that in the experiment, broadening doe nate only from the phonon effects but also from the spectral diffusion. ...
Article
Full-text available
We present an experimental study on the optical quality of InAs/InP quantum dots (QDs). Investigated structures have application relevance due to emission in the 3rd telecommunication window. The nanostructures are grown by ripening-assisted molecular beam epitaxy. This leads to their unique properties, i.e., low spatial density and in-plane shape symmetry. These are advantageous for non-classical light generation for quantum technologies applications. As a measure of the internal quantum efficiency, the discrepancy between calculated and experimentally determined photon extraction efficiency is used. The investigated nanostructures exhibit close to ideal emission efficiency proving their high structural quality. The thermal stability of emission is investigated by means of microphotoluminescence. This allows to determine the maximal operation temperature of the device and reveal the main emission quenching channels. Emission quenching is predominantly caused by the transition of holes and electrons to higher QD’s levels. Additionally, these carriers could further leave the confinement potential via the dense ladder of QD states. Single QD emission is observed up to temperatures of about 100 K, comparable to the best results obtained for epitaxial QDs in this spectral range. The fundamental limit for the emission rate is the excitation radiative lifetime, which spreads from below 0.5 to almost 1.9 ns (GHz operation) without any clear spectral dispersion. Furthermore, carrier dynamics is also determined using time-correlated single-photon counting.
... The TO phonon frequency is shifted up by 7.1 cm −1 with respect to bulk InAs TO frequency (218.8 cm −1 ). Groenen et al have studied the strain distribution and optical phonons in InAs/InP self-assembled quantum dots [59]. From the average strain components and TO deformation potentials, they have calculated the TO frequency shift of 6.8 cm −1 , which is comparable to the shift of InAs QDisks in our InP/InAs superlattice nanowire. ...
... Hence we estimate the strain in the InAs QDisk with ε zz of 2.6%-3.5% and ε xx of -(2.8%-3.2%) from Groenen's theoretical calculation [59]. The result is also agreement with that analyzed from InAs/ InP high-resolution TEM images through geometrical phase analysis method [60]. ...
Article
Full-text available
We report diameter-tailored luminescence in telecom band of InP/InAs multi-heterostructure nanowires with continuously-modulated diameter from microscale to nanoscale. By using the self-catalyzed vapor-solid-liquid approach, we tune the indium particle size, and consequently the InP/InAs nanowire diameter, during growth by modulating the flow rate of the indium source material. This technique allows a high degree of continuous tuning in a wide scale from microscale to nanoscale. Hence it offers an original way to bridge the gap between microscale-featured photolithographic and nanoscale-featured nanolithographic processes and to incorporate InAs quantum disks with tunable diameters into a single InP/InAs quantum heterostructure nanowire. We realized site-defined nanowires with nanoscale diameters initiated from site-defined microscale-diameter particles made with a conventional photolithographic process. The luminescence wavelength from InAs quantum disks is directly connected to the nanowire diameter, by which the strain in the InAs quantum disks is tailored. This work provides new opportunities in the fabrication and design of nanowire devices that extends beyond what is achievable with the current technologies and enables the nanowire shape to be engineered thus offering the potential to broaden the application range of nanowire devices.
... The VFF presented in section 3.3.2 has been frequently used for the calculation of strained heterostructures and specially for quantum dots [51,52,55,60,61] as input for electronic or vibrational band structure calculations. As pointed out before, the extension of the VFF (in the following called XVFF) made to account for the pressure dependence of the bulk modulus is already important when dealing with strongly strained heterostructures to correctly describe the internal strain as well as the one induced by the external pressure. ...
... The application of this simple model for the calculation of the strain in lattice mismatched heterostructures has been demonstrated by many authors[51][52][53]. However, considering the large strain existent in e.g. ...
... However, in single QDs, s is mostly distributed at the interface between the QD and the host material where the nuclear spins have weak influence on the electron spin. More importantly, owing to typical symmetry of C 2v or higher, the shear-strain components at opposite interfaces have different signs such that its effect on both the NQI and anisotropic exchange interaction is averaged out [48]. The QDS investigated in this work has symmetry of C s or lower such that a net effect of shear strain is preserved. ...
Article
Full-text available
Dynamic nuclear polarization (DNP) is well recognized as being important in spintronics and quantum-information processing. DNP gives rise to high nuclear spin polarization that not only can prolong electron-spin lifetime by generating an Overhauser field (OHF), but also has fertilized the idea to explore nuclear spin qubits. In strained quantum-dot structures (QDSs), a nuclear spin is coupled to a strain field via its quadrupole moment. It has been shown that such nuclear quadrupole interaction (NQI) can be used to achieve appreciable DNP and hence electron-spin polarization. Here, we uncover magneto-optical anomalies from a series of laterally arranged (In,Ga)As QDSs that arise from the NQI and DNP in these nanostructures. We find that the principal axis of NQI in symmetry-lowered QDSs significantly deviates from the growth direction, resulting in tilting of OHF with an angle exceeding 37°. The resulting transverse component of OHF is probed with respect to the crystallographic orientations and its influence on the DNP and ensemble spin dephasing is analyzed. We show that a high-symmetry electronic confinement potential for excitons does not guarantee a high-symmetry NQI for nuclei within the same nano-object, thereby calling for correlated optimization in the symmetry of the electronic confinement potential and that of the nuclear spin bath. Our results underline the role of oblique NQI in electron-spin decoherence and depolarization, which has so far largely been overlooked. This work thus sheds light on design rules for engineering the electronic and spin landscape of QDSs for better performance of DNP desirable for applications in spintronics and quantum computation.
... Actually, the frequency shifts of optical phonons are dependent on the strain tensor ε ij and the phonon deformation potentialsK ij . 27,28 The relative frequency shift Δω induced by the strain with respect to the frequency ω 0 of the LO phonon is given by 29,30 Δω ...
Article
The influence of an ultrathin GaP (or GaAs) sublayer on the nonradiative intraband relaxation in InAs/InGaAsP/InP quantum dots (QDs) is investigated. It is found that, based on our studies, the QDs with some heights (e.g., 1.5 nm) and GaP sublayer thicknesses (e.g., 1.03 monolayers) present high states degeneracy in the first excited state (ES) with respect to ground state (GS), which suggests that the Auger relaxation is triggered more easily. We also find that the energy difference of the ES and GS decreases with increasing sublayer thickness, which suggests that the electron-phonon interaction is affected. This work further presents a study of intraband relaxation for an InAs/InP QD with GaP or GaAs sublayer. It is found that there is a critical thickness of the GaP sublayer: when the sublayer is with less than the critical thickness, the intraband relaxation is only determined by one- longitudinal optical (LO) phonon or two-LO phonons, dependent on QD heights. However, with GaAs sublayer, QDs do not have the above feature. This finding may be helpful for designing and optimizing high-speed QD devices.
Article
Full-text available
We report Raman scattering experiments and aluminum (Al) ratio on self-assembled InAlAs quantum dots (QDs) in AlGaAs matrix with different coverage thickness and growth time. The Raman feature assigned to LO phonons in the dots, exhibit a downward frequency shift compared to the one of InAs/AlGaAs QDs. This shift is caused by the strain effects. It is also found that the Al intermixing from the barrier towards the QDs is more important in InAlAs/AlGaAs than in InAs/AlGaAs QDs. We have found that the decrease of the coverage thickness and growth time of the InAlAs/AlGaAs QDs leads to broadening of the Raman peak assigned to LO phonons. This result is attributed to the increase of the In concentration fluctuation in the QDs. In the second part of this report we have investigated the effect of aluminum (Al) concentration on size of In1−xAlxAs/Al0.3Ga0.7As QDs (x=0.28, 0.38, 0.5) by photoluminescence measurements.
Article
We provide direct experimental evidence for the effect of a phonon bottleneck on exciton and spin generation in self-assembled In0.5Ga0.5As quantum dots (QDs). With the aid of tunable laser spectroscopy, we resolve and identify efficient exciton generation channels in the QDs mediated by longitudinal-optical (LO) phonons from an otherwise inhomogeneously broadened QD emission background that suffers from the phonon bottleneck effect in exciton generation. Spin-generation efficiency is found to be enhanced under the LO-assisted excitation condition due to suppressed spin relaxation accompanying accelerated exciton generation. These findings underline the importance of fine-tuning QD energy levels that will benefit potential spin-optoelectronic applications of QDs by reducing spin loss due to the phonon bottleneck.
Article
Full-text available
This paper presents a study of rapid thermal annealing (RTA) effects on self-assembled In(Ga)As/GaAs quantum dots (QDs) using Raman scattering and photoluminescence (PL) spectroscopy. The PL dependence on excitation shows the existence of two size distributions of QDs (small and large) after annealing. However, Raman scattering shows one longitudinal optical (LO) phonon peak of the QD. We have presented two simple models (Raman scattering model (RSM) and PL model) to determine both composition and size of QDs. First, in the RSM, we have taken into account the strain effects. Second, in the PL model, we have assumed a conical QD shape, and we have taken into account the electron-hole confinement, strain effect, and the electron-hole Coulombic interaction. A Gaussian distribution has been introduced to describe the random distribution of indium content for different QDs. We initiated our study with the simulation of the Raman spectra by the RSM to determine the value of indium content of In xGa 1-x As QDs before and after annealing. We have introduced this value into the PL model to determine the QD size. Finally, we found a weak FWHM of composition distribution that shows a homogeneous indium composition in different QDs in both as-grown and annealed QDs.
Article
The minimum energy atomic configurations of stacked GaAs/InAs/GaAs quantum dot structures are calculated by using the conjugate gradient energy minimization with the Stillinger-Weber potentials. The islands are assumed to be either of pyramidal shape or of trapezoidal shape. The numerical results for the five-layer stacked structures show that the normal strains exhibit stepwise up-and-down profiles through the vertical centerlines of the islands and intermediate layers. Next, the molecular dynamics method with the Tersoff potential is applied to single Ge/Si and Si/Ge/Si structures with pyramidal islands in order to investigate the effect of temperature. It is found that there is a considerable difference between the normal strain in the direction perpendicular to the island base obtained by the conjugate gradient minimization with the Stillinger-Weber potential and that obtained by the molecular dynamics method at 800 K.
Article
Full-text available
Using a valence force field, we predict the symmetric lattice distortions around isovalent impurities in 64 semiconductor-impurity systems. For the five systems for which extended x-ray absorption fine-structure (EXAFS) data are available, the results are in excellent agreement with experiment. Our theory also explains quantitatively, without adjustable parameters, the observed bond-length variations in solid solutions A1-xBxC of semiconductor alloys, as well as their excess enthalpies of mixing.
Article
Full-text available
Using a pseudopotential plane-wave approach, we have calculated the electronic structure of strained InAs pyramidal quantum dots embedded in a GaAs matrix, for a few height (h)-to-base(b) ratios, corresponding to different facet orientations \{101\}, \{113\}, and \{105\}. We find that the dot shape (not just size) has a significant effect on its electronic structure. In particular, while the binding energies of the ground electron and hole states increase with the pyramid volumes (b2h), the splitting of the p-like conduction states increases with facet orientation (h/b), and the p-to-s splitting of the conduction states decreases as the base size (b) increases. We also find that there are up to six bound electron states (12 counting the spin), and that all degeneracies other than spin, are removed. This is in accord with the conclusion of electron-addition capacitance data, but in contrast with simple k.p calculations, which predict only a single electron level.
Article
Full-text available
The phonon structure of GaP quantum dots is studied using an atomistic potential model. The dot eigenmodes are obtained from a direct diagonalization of the dynamical matrix and classified using an efficient dual-space analysis method. Our calculations provide a theoretical explanantion for several experimental observations. (1) Depending on the spatial localization, the phonon modes of dots are either dot-interior (bulklike) or surfacelike. (2) The frequencies of the dot-interior modes can be qualitatively described by the “truncated crystal method” using a single branch and a single wave vector of the bulk-phonon dispersion. In contrast, the surface modes cannot be described by this model. (3) The dot-interior modes have a dominant bulk parentage from a specific part of the Brillouin zone, while the surface modes do not. (4) The frequencies of the bulklike Γ-derived longitudinal optical (LO) and transverse optical (TO) phonon modes are found to decrease with decreasing dot size. This decrease reflects the downward dispersion of the bulk optical-phonon branches away from the Γ point. (5) The surface modes located between the bulk TO- and LO-phonon bands have a significant bulk Γ character, and are thus Raman detectable. (6) The dot-interior modes exhibit only a slight LO/TO mode mixing, while the surfacelike modes show a strong mode mixing.
Article
We report a Raman study of confined In As-like and InP-like longitudinal optical and interface phonons in highly strained short-period superlattices grown coherently on (001) InP substrates by Atomic Layer Epitaxy. A resonant enhancement of the InAs Raman scattering cross-section at the E1 gap of InAs appears as the InAs layer thickness increases beyond 2 monolayers. This enhancement indicates that the scattering involves E1 -gap excitons strongly confined in the InAs wells when this gap is created. InAs-like confined longitudinal-optic phonons of A1 and B2 symmetries are observed in both z(x,x)z¯ polarized and z(x,y)z¯ depolarized configurations under InAs E1-gap resonant excitation. This relaxation of the selection rules is analyzed in terms of resonant impurity-induced scattering. It indicates that the symmetry of the exciton wave functions at the L point, which is different from those at the Gamma point, allows intraband coupling of heavy-hole and light-hole states via the deformation-potential interaction.
Article
After demonstrating inconsistencies in some of the better known elasticity calculations, an alternative method of imposing the necessary invariance conditions on the strain energy of a crystal is presented. The new method is equivalent to the Born-Huang procedure but, in addition to providing further insight, also offers one or two operational advantages. For example, it demonstrates that all purely first-neighbor interactions are central only. The method is applied to the calculation of the elasticity of a two-constant model of the diamond type of crystal, and this predicts the relation 2c44(c11+c12)=(c11-c12)(c11+3c12), which is very well satisfied by the experimental data for diamond, silicon, and germanium.
Article
We have utilized resonant Raman scattering to investigate the phonon modes of self-organized Ge quantum dots grown by molecular-beam epitaxy. Both Ge-Ge and Si-Ge phonon modes are found to exhibit strong enhancements at the E1 exciton. The strain in the quantum dots deduced from the phonon energies is consistent with the results of high-resolution transmission electron microscopy. An upper bound on the confinement energy of the E1 exciton in quantum dots was deduced. The enhancement strength in the Si-Ge phonon indicates strong interaction between this mode and the E1 exciton of the Ge dots.
Article
Strained epitaxy has been shown to produce pyramidal-shaped quantum dot structures by single-step epitaxy. In this paper we examine the strain tensor in these quantum dots using a valence force field model. We use an eight-band k⋅p formalism to find the electronic spectra in the highly strained dots. Results obtained for the conduction-band spectra using the effective-mass approach are shown to have serious errors. This is particularly true for excited states in the conduction band. The dependence of the electronic spectra on the quantum dot size and shape is also reported along with comparisons with published experimental results.
Article
The first-order Raman scattering by TO and LO phonons has been measured in GaSb, InAs, and InSb under hydrostatic pressures up to their phase transitions. The Raman frequencies increase nearly linearly while the LO-TO splitting decreases with increasing pressure. The scattering intensities display strong enhancement as the E1 band gaps approach the laser frequencies with increasing pressure. The measured volume dependence of the Raman frequencies and the transverse effective charges is interpreted by means of pseudopotential calculations. The resonance behavior is discussed in terms of resonant Raman scattering near the E1 gaps.
Article
In this paper we report measurements of the effects of large static uniaxial stress along [001], [111], and [110] on the frequency of the k⃗≈0 optical phonons in Ge, GaAs, GaSb, InAs, and ZnSe using first-order Raman scattering. In the absence of stress, the first-order Stokes-Raman spectrum of diamond-type materials exhibits a single peak which corresponds to the k⃗≈0 triply degenerate optical phonons (F2g or Γ25′) while the zinc-blende materials exhibit two peaks, corresponding to the k⃗≈0 LO and TO phonons. The application of the uniaxial stress causes polarization-dependent splittings and/or shifts which are linear in the stress. From these observed splittings and shifts we have obtained experimental values for the phenomenological coefficients (p, q, r) which describe the changes in the "spring constant" of these optical phonons with strain. Comparison of the experimental values is made with several theoretical considerations based on bond-stretching and bond-bending interactions between atoms. The shift due to the hydrostatic component of the strain yields a value for the mode-Grüneisen parameter, which is compared with the results of hydrostatic-pressure measurements. For the zinc-blende-type materials, the doubly degenerate TO-phonon line exhibits both a splitting and shift with stress, while only a shift is observed for the singlet LO-phonon line. In the case of the III-V compounds, one of the split TO lines has a stress dependence equal to that of the LO-phonon line, while this is not the case for the group II-VI material (ZnSe) we have investigated. This latter result is interpreted in terms of the stress dependence of the effective charge.
Article
We present a detailed investigation of the vibrational modes of single thin InAs layers directly grown on InP substrate. The Raman spectra have revealed the presence of features besides the longitudinal-optical and transverse-optical modes related to InP and InAs. We have succeeded in distinctly detecting Raman modes due to the presence of an intermixed InxAs1-xP layer resulting from an incorporation of arsenic in the InP substrate. Moreover, our Raman data show unusual intense features related to the lattice vibrations of interface bonds. The observed energies of these interface modes are in good agreement with the calculation based on a modulated dielectric model. This work clearly demonstrates that the vibrational properties of heterostructures can be sensitive to the structure of the interface and that Raman spectroscopy is a powerful tool to investigate the crystal structure of heterostructures as well as the interfacial chemistry.