Vous êtes sur la page 1sur 176

Jacak· Hawrylak ·W6js

Quantum Dots
Springer
Berlin
Heidelberg
New York
Barcelona
Budapest
Hong Kong
London
Milan
Paris
Santa Clara
Singapore
Tokyo
Lucjan Jacak
Pawel Hawrylak
Arkadiusz W6js

Quantum Dots

With 93 Figures

Springer
Professor Lucjan Jacak
Dr. Arkadiusz W6js
Wroclaw University of Technology
Wybrzeze Wyspianskiego 27
PL-50-370 Wroclaw, Poland

Dr. Pawel Hawrylak


Institute for Microstructural Sciences
National Research Council of Canada
Ottawa, Canada KIA OR6

Library of Congress Cataloging-in-Publication Data


Jacak, Lucjan, 1952-
Quantum dots I Lucjan Jacak, Pawel Hawrylak, Arkadiusz W6js.
p. cm.
Includes bibliographical references and index.

1. Semiconductors. 2. Quantum electronics. 1. Hawrylak, Pawel.


II. W6js,Arkadiusz, 1971- . III. Title.
QC61l.J215 1998 537.6'22 --dC21 97-32688 CIP
ISBN 13: 978-3-642-72004-8 e-ISBN 13: 978-3-642-72002-4
DOl: 10.1007/978-3-642-72002-4
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broad-
casting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this
publication or parts thereof is permitted only under the provisions of the German Copyright Law of
September 9, 1965, in its current version, and permission for use must always be obtained from
Springer-Verlag. Violations are liable for prosecution under the German Copyright Law.
© Springer-Verlag Berlin Heidelberg 1998

The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant pro-
tective laws and regulations and therefore free for general use.
Typesetting: Camera-ready copy by t~e authors using a Springer TEX macro package
Cover design: design & production GmbH, Heidelberg
SPIN 10644678 54/3144 - 5 43 21 0 - Printed on acid-free paper
Preface

We present an overview of the theoretical background and experimental re-


sults in the rapidly developing field of semiconductor quantum dots - systems
of dimensions as small as 10- 8 -10- 6 m (quasi-zero-dimensional) that contain
a small and controllable number (1-1000) of electrons.
The electronic structure of quantum dots, including the energy quan-
tization of the single-particle states (due to spatial confinement) and the
evolution of these (Fock-Darwin) states in an increasing external magnetic
field, is described. The properties of many-electron systems confined in a
dot are also studied. This includes the separation of the center-of-mass mo-
tion for the parabolic confining potential (and hence the insensitivity of the
transitions under far infrared radiation to the Coulomb interactions and the
number of particles - the generalized Kohn theorem) and the effects due to
Coulomb interactions (formation of the incompressible magic states at high
magnetic fields and their relation to composite jermions) , and finally the
spin-orbit interactions. In addition, the excitonic properties of quantum dots
are discussed, including the energy levels and the spectral function of a single
exciton, the relaxation of confined carriers, the metastable states and their
effect on the photoluminescence spectrum, the interaction of an exciton with
carriers, and exciton condensation.
The theoretical part of this work, which is based largely on original re-
sults obtained by the authors, has been supplemented with descriptions of
various methods of creating quantum-dot structures. It includes an extensive
overview of the experimental studies of quantum dots, such as work on far
infrared, photoluminescence and capacitance spectroscopies.
The authors would like to give sincere thanks to Prof. Jurij Krasnyj (Uni-
versity of Odessa) for the helpful discussions and cooperation that allowed
for some of the original results presented here to be obtained.

Wrodaw and Ottawa Lucjan Jacak


August 1997 Pawel Hawrylak
A rkadiusz W 6js
Contents

1. Introduction.............................................. 1

2. Creation and Structure of Quantum Dots. . . . . . . . . . . . . . . . . 5


2.1 Etching............................................... 5
2.2 Modulated Electric Field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 Interdiffusion Between the Barrier and the Quantum Well. .. 10
2.4 Semiconductor Microcrystals. . . . . . . . . . . . . . . . . . . . . . . . . . . .. 10
2.5 Selective Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 11
2.6 Self-Organized Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 12
3. Single-Particle States of Quantum Dots. . . . . . . . . . . . . . . . . .. 15
3.1 Density of States. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 15
3.2 Lateral Confinement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 16
3.3 Fock-Darwin Energy Levels. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 17
3.4 Selection Rules for Intraband Optical Transitions. . . . . . . . . .. 20
3.5 Valence-Band Holes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 21
3.5.1 Binding of Holes ................................. 21
3.5.2 Intersubband Mixing. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 23
4. Properties of an Interacting System ...................... 27
4.1 Generalized Kohn Theorem. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 27
4.1.1 Jacobi Coordinates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 27
4.1.2 Separation of the Center-of-Mass Motion. . . . . . . . . . .. 28
4.1.3 Independence of the Center-of-Mass Excitations
of the Number of Particles. . . . . . . . . . . . . . . . . . . . . . . .. 30
4.1.4 Decoupling of Far Infrared Radiation
and the Relative Motion .......................... 30
4.2 Effect of System Size. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 32
4.3 Effect of the Magnetic Field ............................. 33
4.4 Magic States of a Few-Electron System ............... ; . . .. 35
4.5 The Idea of Composite Fermions ......................... 38
4.6 Harmonic-Interaction Model. . . . . . . . . . . . . . . . . . . . . . . . . . . .. 40
4.7 Reconstruction of the Edge of a Compact Droplet. . . . . . . . .. 41
4.8 Heat Capacity ......................................... 46
4.9 Magnetization ......................................... 47
VIII Contents

5. Intraband Optical Transitions ............................ 51


5.1 Relation with the Kohn Theorem. . . . . . . . . . . . . . . . . . . . . . . .. 51
5.2 Measurements of Far Infrared Absorption. . . . . . . . . . . . . . . . .. 52

6. Interband Optical Transitions. . . . . . . . . . . . . . . . . . . . . . . . . . . .. 59


6.1 Idea of the Photoluminescence Experiment ................ 59
6.2 Carrier Relaxation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 61
6.3 Observation of Magic States of Quantum Dots
in the Absorption Spectrum. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 65
6.4 Interaction of the Exciton with an Additional Electron . . . . .. 69
6.5 Measurements of Photoluminescence. . . . . . . . . . . . . . . . . . . . .. 70

7. Capacitance Spectroscopy ............ . . . . . . . . . . . . . . . . . . .. 83


7.1 Idea of the Capacitance Experiment ...................... 83
7.2 Measurements by Means of Capacitance Spectroscopy. . . . . .. 84

8. Description of the Properties


of Self-Assembled Quantum Dots
Within the Band-Structure Model. . . . . . . . . . . . . . . . . . . . . . .. 97
8.1 Electronic Structure ........................ . . . . . . . . . . .. 97
8.2 Electron System in a Magnetic Field ...................... 103
8.3 Exciton in a Magnetic Field ............................. 112
8.4 Condensation of Excitons ................................ 118

9. Description of a Many-Electron Quantum Dot


with the Inclusion of the Spin-Orbit Interaction .......... 127
9.1 Quantum Dot in the Absence of a Magnetic Field .......... 129
9.2 Quantum Dot in a Magnetic Field ........................ 133

10. Description of an Exciton in a Quantum Dot


Within the Effective-Mass Approximation ................ 141
10.1 Exciton in the Absence of a Magnetic Field ................ 143
10.2 Exciton in a Magnetic Field ............................. 157

References .................................................... 165

Index ......................................................... 173


1. Introduction

Scientific research into electronic systems was limited for a long period of time
to naturally occurring isolated atoms or particles, metallic or semiconductor
crystals, or beams of beta radiation. Most of these are three-dimensional
systems, while an effective reduction of geometry to two or fewer dimensions
- by a strong spatial localization to a plane, line, or point (i.e., confinement
of an electron in at least one direction at the de Broglie wavelength) - occurs
only in the case of atoms and electrons localized on crystal imperfections
(e.g., on impurities).
The beginning of the 1970s marked the new era of research on electronic
structures of dimension limited to two, so-called quantum wells [29, 35]. The
quantum well is a very thin, fiat layer of semiconductor sandwiched between
two layers of another semiconductor with a higher conduction-band energy
[64]. The difference between the conduction-band energies of the two materi-
als binds the electrons in the thin layer. Since the effective mass of an electron
in a semiconductor is small, the de Broglie wavelength is relatively large. The
motion of electrons bound in a layer as thin as several crystalline mono lay-
ers is two-dimensional, and the excitations in the perpendicular direction are
strongly quantized. The material used most commonly for creating quantum
wells is gallium arsenide, GaAs, which in combination with a ternary solution
aluminum-gallium arsenide, AlxGal-xAs, serving as a barrier allows for the
creation of very thin epitaxial layers as a result of the almost equal lattice
constants.
The new, unusual properties of quasi-two-dimensional systems, which
promise applications mostly in electronics and opto-electronics, have at-
tracted the attention of many research laboratories. This in turn has resulted
in a rapid development of production technology and extensive research. The
discovery of the integer quantum Hall effect (IQHE) by Klaus von Klitz-
ing's research team [86] was awarded the Nobel Prize in 1985. The fractional
quantum Hall effect (FQHE) brought the Buckley Prizes to its discoverers
D. C. Tsui, H. C. Stormer, and A. C. Gossard [129]' and to Robert'Laughlin
for his theoretical work [92, 93, 94]. At present, the properties of the quasi-
two-dimensional systems are well investigated and understood, and quantum
wells have been produced and implemented for years in numerous devices,
2 1. Introduction

for instance, laser diodes used in CD players or microwave receivers used in


satellite television.
At the beginning of the 1980s, the rapid progress in technology, especially
very accurate lithographic techniques, made it possible to confine electrons in
a quasi-one-dimensional structure, the so-called quantum wire [109]. Quantum
wires are produced, for example, in the form of miniature strips, etched in a
sample containing a quantum well. The limited abilities of lithography mean
that their typical transverse dimensions are significantly larger than the depth
of the quantum well, reaching 10-500 nm.
Complete quantization of the electron's free motion is implemented by
trapping it in a quasi-zero-dimensional quantum dot. This was first achieved
by scientist from Texas Instruments Incorporated. Reed et al. [116] reported
the creation of a square quantum dot with a side length of 250 nm, etched by
means of lithography (see Fig. 2.2). Subsequent pUblications reporting the
creation of quantum dots in other research centers soon appeared: AT&T
Bell Laboratories [30, 128] and Bell Communication Research Incorporated
[79]. The diameters of these dots were already much smaller: 30-45 nm.
As a result of the strong confinement imposed in all three spatial dimen-
sions, quantum-dot systems are similar to atoms and therefore are frequently
referred to as the artificial atoms, superatoms, or quantum-dot atoms. What
makes quantum dots such unusual objects is, first of all, the possibility of
controlling their shape, their dimensions, the structure of energy levels, and
the number of confined electrons. It is possible, for instance, to create and
investigate such school models as a rectangular or parabolic potential well
binding one or several particles (with the same or opposite electric charges),
as well as the Landau quantization of motion of a single electron, the radiative
recombination from a few-particle system, and so on. The small number of
electrons in typical quantum dots, which facilitates carrying out the ab initio
calculations, makes these systems, mini-laboratories of many-body physics,
particularly attractive for theoretical physicists. Some intriguing effects char-
acteristic of two-dimensional electronic systems such as the formation of the
so-called composite fermions, which lead to the fractional quantum Hall ef-
fect, seem also to occur in quantum dots, where the possibility of numerically
solving the Schrodinger equation may be of great help in understanding them.
A fascinating idea of great promise is that of observing the formation of
the band structure in a crystal lattice consisting of a great number (~ 108 ) of
uniformly separated, identical quantum dots-artificial atoms with shapes and
sizes not encountered in nature. Some reports that describe the properties
of artificial molecules formed by a couple of interacting quantum dots have
already been published [57].
Current experiments concerned with quantum dots focus mainly on study-
ing their optical properties (absorption and emission of light in the visible or
far infrared range, and the Raman scattering of light) and electric properties
(capacitance and transport studies). Since quantum dots absorb and emit
1. Introduction 3

light in a very narrow spectral range, which is controlled, for instance, by an


applied magnetic field, it seems that they might very soon find application
in the construction of more efficient and more precisely controllable semicon-
ductor lasers. Both the first experimental results [84, 121] and the theoretical
expectations are very promising. The strong quantization of electron energy,
with parameters suitable for laser action, particularly in the so-called self-
assembled quantum dots, will probably allow quantum-dot-based lasers to
be able to work at higher temperatures and at lower injection currents [42].
What is also very promising is the possibility of an application of quantum
dots in a new generation of computers. The small dimensions and possibility
of dense packing of quantum-dot matrices could permit them to be used for
memory media of huge capacity.
Interesting information about the creation of the first quantum dots, their
properties, and possible applications is given in a popular paper by Reed [117].
A review of the unusual properties of quantum dots can also be found in the
article by Ashoori [5].
2. Creation and Structure of Quantum Dots

Unlike quantum wells, where the motion of carriers is restricted to a plane


through the crystallization of thin epitaxial layers [64], the creation of quan-
tum wires or dots, which confine the carriers to a space with at least two of
three dimensions limited to the range of the de Broglie wavelength, requires
far more advanced technology.

2.1 Etching

The earliest method of obtaining quantum dots was implemented by Reed


et al. [116]' who etched them in a structure containing two-dimensional elec-
tron gas. The steps of this process are shown in Fig. 2.1: The surface of
a sample containing one or more quantum wells is covered with a polymer
mask, and then partly exposed (Fig. 2.1a). The exposed pattern corresponds
to the shape of the created nanostructure. Because of the required high res-
olution, the mask is not exposed to visible light, but to the electron or ion
beam (electron/ion beam lithography). At the exposed areas the mask is re-
moved (Fig. 2.1 b). Later, the entire surface is covered with a thin metal layer
(Fig. 2.1c). Using a special solution, the polymer film and the protective metal
layer are removed, and a clean surface of the sample is obtained, except for
the previously exposed areas, where the metal layer remains (Fig. 2.1d). Next,
by chemically etching the areas not protected by the metal mask (Fig. 2.1e),
the slim pillars are created, containing the cut-out fragments of quantum
wells (Fig. 2.lf).
In this way, the motion of electrons, which is initially confined in the plane
of the quantum well, is further restricted to a small pillar with a diameter on
the order of 10-100 nm. A typical configuration of layers in such a structure
is illustrated in Fig. 2.2.
The chromium-doped GaAs base serves as the source of free carriers,
which flow into twenty GaAs quantum wells, created above the buffer layers
and separated by the AlGaAs barriers. The etching depth drops beneath the
interface between the last quantum well and the buffer layer. The gold mask
that remains after the etching process may serve as the electrode. The voltage
applied to the electrode controls the number of carriers confined in the dots.
6 2. Creation and Structure of Quantum Dots

Fig. 2. la-f. Process of


quantum dot etching [117]

Fig. 2.2. Configuration of


layers in a quantum dot
GaAs buffer
etched in a GaAs/ AIGaAs
Cr-doped GaAs substrate
superlattice; based on [116]
2.2 Modulated Electric Field 7

Fig. 2.3. Etched quantum dots: (a) diameter of 200 nm, GaAs/ AlGaAs well, elec-
tron scanning microscope picture [125]; (b) diameter of 30 nm, InGaAs/lnP well,
transmission electron microscope picture [128]

The simplicity of producing thin, homogeneous quantum wells makes


GaAs the most commonly used material for creating dots by means of etching
[30, 34, 79, 125, 116]. However, Temkin et al. [128] describe quantum dots
obtained by etching InGaAsflnP wells. Figure 2.3 presents pictures of real
dots obtained using this method.

2.2 Modulated Electric Field


Another method consists in the creation of miniature electrodes over the
surface of a quantum well by means of lithographic techniques, as shown in
Fig. 2.4. The application of an appropriate voltage to the electrodes produces
a spatially modulated electric field, which localizes the electrons within a
small area. The lateral confinement created in this way shows no edge defects,
which are characteristic of such etched structures. An electric gate can also be
created around the etched dot, thus allowing, at least partly, the elimination
of edge defects and additional squeezing of electrons.
The process of spreading a thin electrode over the surface of a quantum
well may produce both single quantum dots [3,4] and large arrays (matrices)
of dots [51, 97, 105, 122]. The modulation of electric potential, produced by
applying a voltage to an electrode, can be realized by the previous preparation
(using a lithographic technique) of a regular array of islets of nonmetallic
material (e.g., of the barrier material) on the surface of the sample. As a
8 2. Creation and Structure of Quantum Dots

Fig. 2.4. Quantum dot on the intersection of electrodes; four internal electrodes
localize the electrons, and four external ones serve as contacts for the electrons
tunneling to and from the dot [117]

result, the distance between the electrode (overlying the surface with the
islets) and the plane of the quantum well is modulated, and the electrons are
bound in small areas under the prepared islets: The photograph of a matrix
of such dots, together with the profile of the confining potential, is shown
in Fig. 2.5. Instead of modulating the distance between the electrode and
the well, it is also possible to build a pair of parallel, thin electrodes above
the well. The lower electrode can have regularly placed holes, which is where
quantum dots are to be created [1, 52]. This idea is shown in Fig. 2.6. If
a voltage is applied to the pair of electrodes, the result is a change both
in the dot size (Vgb) and the depth of the confining potential (Vgt). The
potential depth influences the number of confined electrons. However, when
the additional electrode is introduced between the quantum-well layer and
the doped layer, the number of electrons and the potential depth can be
changed independently.
A very advantageous feature of quantum dots whose electrons are confined
by the electric field produced by a set of electrodes is their smooth lateral
confinement, showing no edge effects. The possibility of controlling certain
parameters is also very important. In experimental reports one, can find in-
formation about these types of dots, which can be created on, for example,
gallium arsenide GaAs [3, 4, 51, 52, 97, 105], indium antimonide InSb [122],
and silicon Si [1].
2.2 Modulated Electric Field 9

Fig. 2.5. Quantum


~Ni[r dots created on InSb;
::::: re SIS t electrons confined by
the electric field (elec-
tron scanning micro-
scope picture). Bot-
tom: shape of the elec-
trode and the config-
uration of band edges
(valence and conduc-
tion bands) [122]

Fig. 2.6. Quantum


dots created on Si;
electrons confined by
the electric field (scan-
ning electron micro-
scope picture); length
of the white line: 1 /-tm.
Bottom: configuration
of electrodes (dimen-
sions: d 1 = 50 nm, a =
400 nm, t = 150 nm)
[1]
10 2. Creation and Structure of Quantum Dots

2.3 Interdiffusion Between the Barrier


and the Quantum Well
Brunner et al. [20] describe a method for obtaining quantum dots based on
a quantum-well material by local heating of a sample with a laser beam. A
parent material of a single, 3 nm thick GaAs quantum well was used, and this
was prepared using the molecular beam epitaxy method (MBE). It was then
placed between a pair of 20 nm thick Alo.35Gao.65As barriers. The topmost
10 nm thick GaAs cap layer was covered with a 100 nm coating of Si3N4 ,
protecting the surface against oxidation or melting by the laser beam. Mod-
ulation of the band gap in the quantum well was obtained by local heating
on the sample with an argon-laser beam with a power of 5.5 m W. The laser
beam was guided along a rectangular contour surrounding an unilluminated
area of diameter 300-1000 nm. At a temperature of about 1000°C a rapid
interdiffusion of Al and Ga atoms occurred between the well and the bar-
riers, which led to the creation of a local modulation of the material band
structure, i.e., to the creation of the potential barrier, which surrounds the
unilluminated interior of the rectangle. For larger dimensions of the illumi-
nated rectangle the obtained effective potential that confines the electrons
was flat inside the dot (this area will be called A - pure GaAs) , and steep
near the edge (area B - a solution of AIGaAs replaced pure GaAs due to
heating). With a decrease of the illuminated rectangle, the area A shrinks.
According to the authors of [20], for dimensions near 450 nm the effective
potential confining electrons is close to an isotropic parabola.
However, it should be mentioned that the details of the electron confining
potential in a quantum dot of any type cannot be measured directly (except
for the geometric dimensions) and are alternatively obtained through the
interpretation of various indirect effects, related to the electronic structure
of the object.

2.4 Semiconductor Microcrystals


It is also possible to create quantum dots in the form of semiconductor micro-
crystals immersed in glass dielectric matrices. In the first experiment based on
that idea, carried out by Ekimov et al. [38], silicate glass with about 1% ad-
dition of the semiconducting phase (CdS, CuCI, CdSe, CuBr) was heated for
several hours at a temperature of several hundred degrees Celsius, which led
to the formation of appropriate microcrystals of almost equal sizes. Knowl-
edge of the dependence of the average crystal radius a on the temperature
and heating time
(2.1)
allowed for controlling their size. The radii of dots measured in different
samples varied in the range 1.2-38 nm. As a dielectric matrix, the alkaline
2.5 Selective Growth 11

chlorides can be used instead of glass [66]. The heating of such a matrix with
the addition of copper leads to the formation of CuCl microcrystals.

2.5 Selective Growth


Quantum dots can also be created through the selective growth of a semi-
conducting compound with a narrower band gap (e.g., GaAs) on the surface
of another compound with a wider band gap (e.g., A1GaAs) [45]. The re-
striction of growth to chosen areas is obtained by covering the surface of the
sample with a mask (Si02) and etching on it miniature triangles. On the
surface that is not covered with the mask the growth is then carried out
with the metal-organic chemical vapor deposition method (MOCVD), at a
temperature of 700-800°C. The crystals that are created have the shape of
tetrahedral pyramids, and hence when the first crystallized layers are the lay-
ers of the substrate compound (A1GaAs) and only the top of the pyramid is
created of GaAs, it is possible to obtain a dot of effective size below 100 nm.
The pictures of such dots and the configuration of layers GaAs/ A1GaAs are
shown in Fig. 2.7.
A different variant of the method of selective growth is described by
Lebens et al. [95]. Onto the 2 /Lm thick Alo.3sGao.62As substrate a layer

AI~S:?2
(111)8 GoAl
Fig. 2.7. Quantum dots created on the surface of GaAs in selective MOCVD growth
(scanning electron microscope pictures); width of the electron localization area at
the top of the pyramid is about 100 nm [45]
12 2. Creation and Structure of Quantum Dots

Fig. 2.8. Quantum dots (a), and quantum wires (b) created on the surface of GaAs
in MOVPE selective growth (scanning electron microscope pictures) [95]

of 10 nm thick GaAs was deposited and covered with a 20 nm Si3N4 mask.


The mask was later illuminated in chosen areas with the electron beam and
removed through the plasma etching. In the growth process, which is carried
out with the metal-organic vapor phase epitaxy method (MOVPE), GaAs
sedimented only outside the areas covered with the mask. The thickness of
the crystallizing GaAs layers was determined to be 100 nm. After covering
the newly created structure with a layer of Alo.2Gao.8As, quantum dots of
diameter 70-300 nm and quantum wires of width 90-300 nm and length
~ 0.1 mm were obtained. Pictures of such structures are shown in Fig. 2.8.

2.6 Self-Organized Growth

Petroff and DenBaars [110] describe another method for the self-crystallization
of quantum dots that does not require the creation of a mask. When the
lattice constants of the substrate and the crystallized material differ consid-
erably (7% in the case of GaAs and InAs, the most commonly used pair of
compounds), only the first deposited monolayers crystallize in the form of
epitaxial, strained layers with the lattice constant equal to that of the sub-
strate. When the critical thickness is exceeded, a significant strain occurring
in the layer leads to the break-down of such an ordered structure and to
the spontaneous creation of randomly distributed islets of regular shape and
similar sizes. The shape and average size of islets depend mainly on factors
such as the strain intensity in the layer as related to the misfit of lattice
constants, the temperature at which the growth occurs, and the growth rate.
The phase transition from the epitaxial structure to the random arrangement
of islets is called the Stranski-Krastanow transition [127]. Figure 2.9 presents
the dependence of the critical number of InGaAs monolayers deposited on
the GaAs substrate, at which the phase transition occurs, on the indium con-
centration in the solution (lower axis), or alternatively the misfit of lattice
constants (upper axis), i.e., the strain.
2.6 Self-Organized Growth 13

misfit strain with GaAs


0.02 0.03 0.04 0.05 0.06 0.07
35
30
25
20
15 Islands
10
5 Fig. 2.9. Dependence of the critical
thickness of the InGaAs layer, at which
o ~WWWW~Ull~llU~WWUil
the Stranski-Krastanow phase transition
0.2 0.3 0.4 0.5 0.6 0.7 O.B 0.9 1.0 takes place, on the indium concentration
composition: x (i.e., the strain) [110]

Fig. 2.10a-c. Evolution of islets - self-assembled quantum dots (white circles), in


the molecular beam epitaxy (MBE) growth of InAs on GaAs surface. Subsequent
pictures correspond to the increasing coverage of 1.6, 1.7, and 1.8 monolayer; size
of presented areas is 1 x 1p.m2 [110]

On the other hand, Fig. 2.10 presents subsequent phases of the creation of
islets for a pair of compounds with the maximum misfit of lattice constants:
InAs and GaAs, where according to Fig. 2.9 the transition occurs at the
1.8 monolayer deposition. When the process of crystallization is terminated
shortly after reaching the phase transition, the islets evolve to the state of
quasi-equilibrium, in which they assume the shape of pyramids [47,84, 103]
or flat, circular lenses [40,41, 114, 115] formed on a thin layer of InGaAs (the
wetting layer). When the inhomogeneous InGaAs layer is eventually covered
with another GaAs layer, a structure of a quantum well with significantly
increased thickness in very small areas is obtained.
Raymond et al. [114] report the growth of self-assembled dots in the shape
of lenses with ~ 36 nm diameter and ~ 4.4 nm height (with fluctuations of
5-10%). Marzin et al. [103] obtained dots in the shape of regular pyramids
with a square base of side ~ 24 nm and height ~ 2.8 nm (with fluctuations
of ~ 15%), and distance between neighboring dots ~ 55 nm.
14 2. Creation and Structure of Quantum Dots

The quantum dots formed in the Stranski-Krastanow phase transition are


called self-organized or self-assembled dots (SAD). The small sizes of the self-
assembled quantum dots (diameters in the range of 30 nm or even smaller),
homogeneity of their shapes and sizes in a macroscopic sample, perfect crystal
structure (without edge defects), and the fairly convenient growth process,
without the necessity of the precise deposition of electrodes or etching - are
among their greatest advantages. Thus there is great hope regarding their
future application in electronics and opto-electronics.
3. Single-Particle States of Quantum Dots

3.1 Density of States

The energy of an electron confined in an area as small as a quantum dot


is strongly quantized, i.e., the energy spectrum is discrete. In typical struc-
tures, with characteristic dimensions in the range of 10-100 nm, the distance
between neighboring energy levels is on the order of a few meV. As shown
in Fig. 3.1, the quantization of energy, or alternatively, the reduction of the
dimensionality of the system, is directly reflected in the dependence of the
density of states on energy.
The density of states for a three-dimensional system (bulk semiconductor)
has the form
dN ~E3/2 = E1/2. (3.1)
dE ex dE '
for a two-dimensional system (quantum well) is a step function,
dN d
dE ex dE L (E-ci) = L 1; (3.2)
ci<E ci<E

for a one-dimensional system (quantum wire) has a peculiarity,

dN ex ~ '""' (E _ c·)1/2 = '""' (E _ c)-1/2. (3.3)


dE dE L t L t ,
ci<E ci<E

energy
Fig. 3.1. Density of states as a function of energy in systems with different numbers
of spatial dimensions: 3D, bulk material; 2D, quantum well; ID, quantum wire; OD,
quantum dot
16 3. Single-Particle States of Quantum Dots

and for a zero-dimensional system (quantum dot) has the shape of 8-peaks,

(3.4)

Above, Ci are discrete energy levels, eis the Heaviside step function, and 8
is the Dirac function.
At low enough temperatures (several K) the energy of phonons is too low
to excite the electrons and the strong quantization of energy determines the
striking electronic properties of quantum dots.

3.2 Lateral Confinement


Since quantum dots are created mainly through producing a lateral con-
finement Vex, y) restricting the motion of the electrons, which are initially
confined in a very narrow quantum well (QW) by the potential VQw(z), they
usually have the shape of flat disks, with transverse dimensions considerably
exceeding their thickness. The energy of single-particle excitations across the
disk exceeds other characteristic energies in the system, and the confined
electrons can be considered as two-dimensional. The lateral potential of a
quantum dot differs significantly from the Coulomb potential binding elec-
trons in an atom - the latter has a peculiarity in the center. Depending on
the method used to create the dot, the lateral potential can be approximated
by a model potential. The potential of an etched dot with a considerable
radius is fairly close to a rectangular well with rounded edges. When a dot is
small (i.e., when its radius is comparable to the characteristic length of the
variation of the lateral potential near the edge), a good approximation offers
simple smooth potentials, such as a Gaussian well,
v (r) = - Vo exp ( _r2 / L 2 ), (3.5)
or a Poschl-Teller potential,
VCr) = -Vocosh- 2 (-r 2 /L 2 ). (3.6)
In the cases where the scattering states can be neglected, they may be re-
placed at the bottom with a parabolic well,
VCr) = Vo + kr2. (3.7)
Due to the independence of the positions of resonance peaks in the far infrared
(FIR) absorption on the number of confined electrons N (as a consequence of
the so-called generalized Kohn theorem, discussed in Sect. 4.1), unique for the
parabolic well, a comparison of FIR spectra obtained for different N can pro-
vide information about whether the parabolic well is a good approximation
of the actual potential.
Kumar et al. [88] described the results of a numeric, self-consistent solu-
tion of the Poisson and Schrodinger equations in the Hartree approximation
3.3 Fock-Darwin Energy Levels 17

for a quantum dot produced by a voltage applied to a metal electrode covering


a square GaAs layer with an area of 300 x 300 nm and a thickness of 30 nm,
created on the surface of a GaAs/ AlGaAs quantum well. A small number of
electrons (N::; 10) was considered. It follows from these calculations that de-
spite of the square shape of the electrode, the confining potential is to a good
approximation circularly symmetric, with a diameter of;:::; 100 nm, which is
considerably smaller than the size of the electrode. The separation between
the energy levels is almost independent of N, and the magnetic field evolution
of energy levels is similar to that of levels of a parabolic well (see Fig. 3.2).
This result explains the lack of dependence between excitation energy and
the number of particles measured in several experiments [I, 34, 105, 122], in-
dicating the parabolic profile of the confining potential. Therefore, this also
justifies the application of such an approximation in model calculations.

3.3 Fock-Darwin Energy Levels

The motion of a conduction-band electron confined in a two-dimensional


parabolic well in an external perpendicular magnetic field, in the effective-
mass approximation, is described by the following Hamiltonian:

H -1- ( p - -A . + -me)2
1 * wor
2 2
2m* c 2

-
2m*
1 *
p2- + -m
2
(2Wo + -w
1 2) r 2-
4 e
1 l
-w
2 e z, (3.8)

where m* is the effective mass, r is the position, p is the momentum, lz =


XPy - ypx is the projection of the angular momentum onto the field direction,
A is the vector potential ofthe magnetic field B (B = rotA, in the symmetric
gauge: A = ~B[y, -x,D]), and We = eB/(m*c) is the cyclotron frequency.
The eigenstates of such a Hamiltonian and the corresponding self-energies
were determined analytically by Fock [43] and Darwin [33] long before the
appearance of nanostructures. Using the complex variables
z = x + iy, z* =x - iy, (3.9)
and
8* = 8 x + iBy (3.10)
z 2'
two pairs of creation/annihilation operators (boson operators of single-
particle excitations) can be defined:
z* /(2lo) + 2l08z + _ z/(2lo) - 2l08;
a= V2 ' a - V2 '
b = z/( 2lo) + 2l08; b+ = z* /( 2l o) - 2l08z (3.11)
V2 ' V2 '
18 3. Single-Particle States of Quantum Dots

where

(3.12)

is the effective length scale, equal to the magnetic length in the absence of
the confining potential

lB = [ffi. (3.13)

The operators defined above fulfill the following commutation relations:


[a, a+] = [b, b+] = 1, [a, b] = 0, [a, b+] = 0, (3.14)
and hence the pairs of operators (a, a+) and (b, b+) are independent. The
Hamiltonian (3.8) can be written in the form

H = &.;+ (a+a +~) +&.;_ (b+b +~) , (3.15)

i.e., as a pair of independent harmonic oscillators, with a pair of characteristic


frequencies

(3.16)

The eigenstates of the Hamiltonian (3.8) are constructed by acting with the
creation operators on the ground state (a 100) = b 100) = 0):

In+n-) = 1 (a+r+ (b+r- 100). (3.17)


.jn+!n_!
The eigenenergies corresponding to the states In+n-) (Fock-Darwin energy
levels) are

(3.18)

In the absence of a magnetic field, these levels are degenerate, W+ = w_ =


wo, whereas a strong magnetic field (w c » wo) leads to the formation of
the structure of Landau energy levels, separated by the cyclotron energy
&.;+ ;::;::&.;c·
The system has a circular symmetry, so the operator of the angular mo-
mentum component along the symmetry axis, lz = XPy - YPx = z8z - z* 8; ,
commutates with H. The operator lz is diagonal in the basis of Fock-Darwin
states:
lz(n+n_) = n_ - n+. (3.19)
The Fock-Darwin states In+n-) can be also written in the form
* 1 [ZZ*]
'F exp + 412
(8z)n+(8;)n- [ZZ*]
. j ' , exp -2[2 . (3.20)
1>n+n_ (z, Z ) =
v L/Ir 0 n+.n_. °
3.3 Fock-Darwin Energy Levels 19

Representation in cylindrical coordinates (r,O) is often very useful. Due to


the circular symmetry of the Hamiltonian, the motion in the angular variable
separates (it is a free motion), and the appropriate component of the angular
momentum m is a good quantum number. The electron wave function can
hence be written in the form
(3.21)
where the angle-dependent function

'I/lm(O) = vk eimO (3.22)

is the eigenfunction of the operator of the angular momentum projection,


with eigenvalue m, while the radius-dependent function has the form

(3.23)

In the above expression .ch,:,1 denotes the Laguerre polynomials


1 dn ,.
.c1ml(z) = _z-Iml eZ __ (zn,+lm1e-z); (3.24)
nr m! dz n ,.
n = 0, 1, ... is the principal quantum number; m = -n, -n + 2, ... , n - 2, n is
the azimuthal quantum number; and nr = (n -JmJ)/2 is the radial quantum
number. The p,airs of quantum numbers (n, m) and (n+, 1L) are related by
n = n_ + n+, m = n_ - n+. (3.25)
The eigenenergies expressed by the quantum numbers (n, m) are
1
c(n, m) = h[2(n + 1) - 211wc m, (3.26)

where
1
[22 = w5 + -w~.
4
(3.27)

The evolution of the energy spectrum in an increasing magnetic field is


presented in Fig. 3.2. The Zeeman spin splitting (very small for GaAs) is
neglected here. The pairs of numbers on the vertical axis give the electron
eigenstates (n+, n_ ), while the straight dashed lines show the energies of
subsequent Landau levels (the levels in the absence of a parabolic potential
in the plane of a dot, or w_ = 0):

(3.28)

The thick line represents the tenth energy level. For the special values of the
magnetic field, corresponding to the integer ratio w+ : w_, the characteristic
of a zero magnetic field degeneracy of levels is restored. The crossings of
energy levels at these fields lead to the occurrence of rapid changes of slope
20 3. Single-Particle States of Quantum Dots

>.
f2> 12
(]) 11
c:
(])
10
/

40,31,22, ...
./
/
/

/
/ ""
30,21,12,03 ~-_*"~ " "
""
(i)+
r---t-----; 05 j;
20,11,02 ---,..'"---01-----104 :::I
03 g.
02
01 ~
=-
10,01 00 m..
1 "
1 "
1 //
1-__,.;/:"'-7/:"'---
00 /"
/ ,,/
1/ / "
~~-

magnetic field
Fig. 3.2. Evolution of the Fock-Darwin energy levels in a magnetic field; dashed
lines - the Landau energy levels; vertical arrows - allowed optical transitions

(small peaks) in the energy of higher single-particle levels (as was shown in
the example of the tenth energy level: at the point w+ : w_ = 2 the transition
between orbitals 11 and 03 occurs).

3.4 Selection Rules for Intraband Optical Transitions


Since the wavelength of light corresponding to the energy of single-particle
excitations of a quantum dot (E = 1 meV -> ,\ :::;::: 1 mm) is much larger
than the dot size « 111m), the oscillations of the electromagnetic field in-
duced by this radiation are almost homogeneous over the area of the dot.
The perturbation Hamiltonian, which describes the interaction between the
electron and the electromagnetic wave, can therefore be written in the dipole
approximation
He - m = eEor exp(iwt), (3.29)
where e is the electron charge and Eo is the amplitude of oscillations of the
electric field. In the first order of perturbation, the probability of the optical
3.5 Valence-Band Holes 21

transition (n+ n_) +-+ (n~ n'J is proportional to the squared perturbation
Hamiltonian,
A(n~n~,n+n_) ex (n~n~IHe-mln+n-) ex (n~n~1 r In+n-), (3.30)
while the respective oscillator strength [91] is

"
f (n+n_,n+n_ ) = TW n+n_,n+n_ ) 1A ("
2m * .(" ) 12
n+n_,n+n_, (3.31 )

where nw(n~n~, n+n_) = I€(n~n~) - €(n+n_)1 is the resonance transition


energy. By calculating the integral over the angle in the expression for am-
plitude A, the following selection rules are obtained:
n~ = n+ ± 1 or n~ = n_ ± I, (3.32)
together with the allowed resonance frequencies w±, indicated in Fig. 3.2 and
defined by (3.16).

3.5 Valence-Band Holes

3.5.1 Binding of Holes

Beside the conduction-band electrons, the valence-band holes are the second
type of carriers that can be bound in a quantum dot. The binding of holes,
initially confined to a two-dimensional quantum well, in the quantum dot is
possible by, for example, applying a spatially modulated electric field. The
dots that bind electrons in this way have been described in Sect. 2.2. Due to
the opposite electric charges of the two types of carriers, the lateral potential
defining dots of this kind cannot attract both electrons and holes at the same
time (either bind electrons and repel holes, or vice versa).
In the case of a quantum well it is possible to achieve the effect of bind-
ing both electrons and holes as a result of the appearance of steps in the
conduction and valence-band energies at the interface between the well and
surrounding barriers. As shown in Fig. 3.3, the energies of conduction and
valence bands can be arranged in two ways on the interface [64]. In a con-
travariant heterostructure (e.g., on the GaAs/ AIGaAs interface) both elec-
trons and holes are confined in the same material (here GaAs). However, in a
covariant heterostructure (e.g., on the InAs/GaSb interface) electrons accu-
mulate in one layer (here InAs), while holes flow into the other (here GaSb);
i.e., the barrier for electrons is the well for holes and vice versa.
An interesting problem is whether a quantum dot can bind both types of
carriers at the same time. In the case of a quantum well it was possible to
obtain the configuration of band energies on the well-barrier interface shown
in Fig. 3.3a, i.e., a contravariant heterostructure. In such a structure both
types of carriers are bound simultaneously. It should, however, be remem-
bered that the quantum well is a crystalline structure theoretically infinite in
22 3. Single-Particle States of Quantum Dots

~ (a) ~ (b)
Q)
c vacuum level ~ vacuum level
Q)
~ -~~r~:-------------------'
Ea

E.a L-'-~'- - --,L----',Il

configurational coordinate configurational coordinate


Fig. 3.3. Profile of the band-edge energies in a quantum well: (a) contravariant
heterostructure (GaAsjAlGaAs), (b) covariant heterostructure (InAsjGaSb), after
[64]. Indices 1 and 2 represent the material of the well and the barrier, respectively;
ECi, Evi, E gi , O"i (i = 1, 2) are the energies of conduction (CB) and valence (VB)
bands, band gaps, and the electron susceptibility in both materials, respectively;
EgQw is the effective band gap of the quantum well

two directions, which justifies its description in the band model. The quan-
tum dot, on the contrary, is a very small object, and using a band model to
describe its electronic properties seems to be a rather rough approximation.
When a pair of semiconductors is chosen as in Fig. 3.3a and the quantum
dot is created in the form of a small microcrystal of the material of the well
immersed in the material of the barrier (e.g., the self-assembled dots), then
the band description of such a structure should be approached with great
caution, as it is only an approximation. In this case it seems more appro-
priate to describe the quantum dot as a perturbation of the crystal field of
a surrounding semiconductor, and to search for the electron-hole structure
of the dot in the effective-mass approximation [16]. The applicability of the
effective-mass approximation requires the condition of a small variation of
the perturbing potential on the interatomic distance. This seems to be true
particularly for the dots created by interdiffusion (see Sect. 2.3), but is rather
questionable for the self-assembled dots (see Sect. 2.6). Conversly, the idea of
a simultaneous binding of electrons and holes in the self-assembled dots has
been supported by recent pseudopotential calculations [44]. However, if the
effective-mass approximation is accepted, then the bare potential (electric by
nature) that attracts a hole has to repel an electron, and vice versa.
The charge-neutral electron-hole pair can, however, be bound in such a
potential as a result of the mutual Coulomb attraction between the particles.
It should be stressed that the two approaches, using (i) the approximation
of the band structure of the material forming the quantum dot and (ii) the
3.5 Valence-Band Holes 23

effective-mass approximation, i.e., two completely different starting points,


will eventually lead to a similar main result, the binding of an exciton in
the dot, and differ only in certain details (see Chaps. 9 and 10). This resem-
bles to a certain degree the situation where starting with both weak- and
tight-binding approximations - the band structure of a crystal is eventually
reached.
The two types of carriers can be introduced into the quantum dot by cre-
ating the excitons (electron-hole pairs) using light of an appropriate wave-
length. The lifetime of an exciton bound in a dot is finite due to the nonzero
overlap of the electron and hole wave functions, and the observed emission of
a photon as a result of a radiative recombination (e.g., wavelength, the shift
between absorption and emission energy) reflects the electronic structure of
the system. Absorption and emission will be described more extensively in
Chap. 6.

3.5.2 Intersubband Mixing

Unlike what we see in the case of conduction electrons, a realistic description


of valence-band holes requires the inclusion of a larger number of subbands.
The valence band is built from atomic p-type orbitals, and thus its states
carry an internal (band) angular momentum equal to unity. In a system
with circular symmetry (bulk semiconductor, quantum well, circular quantum
dot, etc.) both the angular momentum length j and its projection along the
axis of symmetry j z are good quantum numbers of a single, noninteracting
particle. Denoting the spin angular momentum by (T, internal (band) angular
momentum by m, and the orbital angular momentum of a particle by l, we
have
j = (T +m + l. (3.33)
In general, the lengths and projections of (T, m, and l separately are not good
quantum numbers.
The single-particle states of the valence band with l = 0 are classified in
the following way, according to the values of (j, j z) = (I (T + m I, 0"z + m z ) :
- (~, ±~), heavy-hole subband,
- (~, ±~), light-hole subband,
- C!, ±~), spin-orbit split-off subband.
In the simplest approach (the effective-mass approximation) all subbands
are treated independently. Each described sub band is then characterized by
a pair of effective masses m~ and m'i, defining the dynamics along and across
the axis of symmetry z). In a two-dimensional system (e.g., a quantum well
from which quantum dots are typically etched), the mass m'i, which de-
scribes the dynamics in the direction transverse to the plane of the system,
24 3. Single-Particle States of Quantum Dots

can be neglected, and the only difference between such a simplified descrip-
tion of holes and that of electrons consists in the occurrence of a number of
subbands (a number of distinguishable types of particles, characterized by
various masses).
A more accurate description of holes requires taking into account the
interaction between the sub bands (intersubband mixing). The model Lut-
tinger Hamiltonian HL [98] describes the interaction between the heavy-
hole and light-hole sub bands (u and m oriented in the same direction:
Iii = lu + ml = ~). In the basis of jz = {+~, -~, +~, -D,
the Hamiltonian
HL is expressed as follows (in the so-called circularly symmetric approxima-
tion):

(3.34)

The operators that appear in the above expression are defined as follows (we
use here the atomic units where the effective Bohr radius aB and the effective
Rydberg Ry are equal to unity):

(3.35)
"2 1""
where k± = -i(ox ± iOy ), kz = -ioz, and k = 'i(k+k- + Lk+).
A A "A

The effective masses for heavy holes (h) and light holes (1) along the
axis of symmetry (z) and in the perpendicular plane (r) are defined by the
Luttinger parameters of a given material (e.g., for GaAs 1'1 = 6.8,1'2 = 1.9,
and 1'3 = 2.9):
-1
m hz = 1'1 - 21'2, -1
mhr = 1'1 + 1'2,
m -1
1z = 1'1 + 21'2, mlr-1 = 1'1 -1'2· (3 .36)
Figure 3.4 presents the dispersion curves for the 8.5 nm thick GaAs quan-
tum well, and the discrete energy levels of the quantum dot created in the
form of a disk 40 nm in diameter etched out of such a well. The results were
obtained by diagonalization of the Luttinger Hamiltonian (3.34), i.e., in the
band-structure approximation, taking band parameters appropriate for the
bulk material of the quantum dot (GaAs).
In the case of a quantum well, the continuous energy spectrum splits into
discrete subspaces (quantum-well subbands), which correspond to different
discrete states of motion in the direction perpendicular to the plane of the
well (uncoupled from the in-plane motion). In Fig. 3.4 we have plotted energy
E as a function of the in-plane wave vector k for a couple of the lowest
3.5 Valence-Band Holes 25

..
20~------~--~--~~~~------~

18

16

14

~ 12
>-
0)

~ 10
s:::
G> Fig. 3.4. Energy as a
8 function of wave vec-
tor for a valence-band
hole confined in a quan-
6
tum well (lines) and
lines - OW in a quantum dot (cir-
4 cles); thin lines - the
dots - 00 effective-mass approxi-
2,.~~~~~~~~~~~~~~~ mation; thick lines -
o 2 3 4 5 6 7 8 9 10 results including inter-
wave vector (a-~) subband mixing

quantum-well subbands. In the case of a quantum dot, the entire energy


spectrum is discrete. For each discrete state the average value of the in-plane
wave vector was calculated. In Fig. 3.4 the discrete states are represented by
circles showing eigenenergy E and the average wave vector k.
The thin lines correspond to noninteracting subbands, and hence to
parabolic (free particle) dispersions, with different effective masses for light
and heavy holes. The thick lines represent the eigenenergies of the Luttinger
Hamiltonian. Above a certain value of the wave vector (kaB ~ 2, aB is the
Bohr radius) the discrepancy between the two solutions (i.e., the effects of
intersubband mixing) cannot be neglected. Small dimensions of the quantum
dot lead to high average values of the wave vector k, and for the dot of 40 nm
diameter almost all the states lie within an area of strong interaction between
the subbands of light and heavy holes.
The significance of the intersubband interaction means that the single-
band approximation should be discarded as far as the description of valence-
band holes confined in the quantum dot is concerned. As a consequence, the
generalized Kohn theorem, described further in Sect. 4.1, does not hold for
holes [18, 19, 31, 32]- As shown in the experiment by Bayer et al. [8], the (spin-
orbit) interaction between the sub bands of the valence band considerably
increases with a decrease of dot dimensions_
However, certain factors, for example the strain present in the self-
assembled quantum dots, may lead to a considerable weakening of the in-
tersubband interaction, and then the single-band approximation for valence-
band holes can still be justified (see Chap. 8).
4. Properties of an Interacting System

4.1 Generalized Kohn Theorem

4.1.1 Jacobi Coordinates

The advantage of the explicit separation of the center-of-mass motion from


the relative dynamics in the description of quantum dots was first demon-
strated by Laughlin for the case of three electrons [93]. The theory of a
many-electron dot in the generalized Jacobi coordinates, utilizing both single-
particle and many-particle, symmetries was formulated by Hawrylak [55]. In
this chapter we present a derivation based on the standard Jacobi coordi-
nates.
Let us consider a system of N identical particles (with equal masses). The
transformation from the system of variables (positions and momenta) that
describe the motion of individual particles,
(4.1)
into the system of N - 1 independent relative variables that describe the
motion of particles with respect to the center of mass of the system:
(4.2)
and the pair of variables that describe the motion of the center of mass of
the system,
(4.3)
is given by the Jacobi matrix U == Ofl/or, defined as
1 -1 0 0 0
1 1
"2 "2 -1 0
1 1 1
"3 "3 "3 -1
U= (4.4)
1 1
N-l N-l
-1
1 1
N N
i.e.,
28 4. Properties of an Interacting System

t if j < i + 1,
Uij ={ -1 if j = i + 1, (4.5)
o ifj>i+1.

The pair of matrices U and W describe transformations of coordinates and


associated momenta from one system to the other:
(!=UT, 7l'=W T p,
T = W(!, P = U T 7l'. (4.8)
The Jacobi variables fulfill the same commutation relations as T and p:
[(!i,7l'j] = Dij in. (4.9)

4.1.2 Separation of the Center-of-Mass Motion

Let us consider a system of N interacting electrons described by the Hamil-


tonian
N
H L [aD + alTi + a2Pi + a3T; + a4P; + a5 T i X Pi]
i=l
N

+ L Vint (T i - Tj ), (4.10)
i<j

where aD, ... ,005 are constants, and Vint is the interaction potential.
Using (4.5) and (4.7), subsequent components of the Hamiltonian (4.10)
can be expressed in terms of the Jacobi variables (unless indicated, the sum-
mations over i, j, k, and l run from 1 to N):
1. The term linear in coordinates:
LTi = 'L:Wij(!j == N(!N (4.11)
ij
4.1 Generalized Kohn Theorem 29

2. The term linear in momenta:


LPi = L Uji'Trj = 'TrN (4.12)
ij
3. The term quadratic in coordinates:

LT; = LWijWikejek = L~ie; (4.13)


ijk
where
i
Hi if i < N,
(4.14)
N ifi = N
4. The term quadratic in momenta:

LP;= 2.=k UjiUki'Trj'Trk = L :i'Tr; (4.15)


""J "
(~i as above)
5. The term linear in angular momenta:

L[Ti x Pi]a = LLcabe[Wijej]b[Uki'Trk]e = Lei x 'Tri (4.16)


ijk be
where [.. .]a denotes the vector's ath component and Cabe is the com-
pletely antisymmetric tensor (c123 = 1, and an odd permutation of in-
dices changes the sign)
6. The term containing the differences of coordinates (in Vint}:
N-l
Ti - Tj = L Wikek - L Wjlel = L (Wik - Wjk)ek (4.17)
k I k=l

Denoting the center-of-mass variables as in (4.3), the Hamiltonian (4.10) can


be written in the form

H {
1 2 + Ct5RxP }
Cto+CtlNR+Ct2P+Ct3NR2 +Ct4NP

+ {~
8 [Ct3i+1ei
i i+1
+ Ct4-i-'Tri 2
+ Ct5ei x 'Tri 2 ] }
+ Hint ,(4.18)

where Hint is the total interaction Hamiltonian, dependent only on the rela-
tive coordinates {el,"" eN-l}. As indicated in (4.18) by the large braces,
the Hamiltonian H can be split into a sum of two Hamiltonians: ,Hem' ex-
pressed in terms of the center-of-mass variables, and H re [, expressed in terms
of the relative variables:
(4.19)
Due to the separation of variables, these two Hamiltonians commutate:
30 4. Properties of an Interacting System

(4.20)
Thus, if the system of electrons is described by the Hamiltonian as in (4.10),
i.e., if the single-particle Hamiltonian contains only the terms linear and
quadratic in coordinates and momenta and linear in angular momenta (the
elements of types const, r, r2, p, p2, and r x p) and if the interaction potential
depends only on the differences of individual coordinates, then the motion of
the center of mass of the system decouples from the relative motion.

4.1.3 Independence of the Center-of-Mass Excitations


of the Number of Particles

Let us consider a system of N interacting electrons moving in a parabolic po-


tential well and in a perpendicular magnetic field. The single-particle Hamil-
tonian is given by (3.8). Let us also assume that the interaction potential
depends only on the differences of coordinates (for the Coulomb interaction
this requires that the medium have a uniform dielectric constant). In such a
case, by virtue of the previous point, the center-of-mass Hamiltonian sepa-
rates and has the form

Hem
p2
= 2M + 2M
1 (2 + 41 w2) R 2-
Wo e
1
2 WeLz , (4.21 )

where M = Nm is the total mass of the system, and Lz is the projection of


the center-of-mass angular momentum L = R x P onto the direction of the
magnetic field. The eigenenergies of the Hamiltonian (4.21) do not depend on
mass M (see Sect. 3.3), but only on Wo and We. Hence, the excitation spectrum
of the Hamiltonian (4.21) does not depend on the number of electrons N
and is identical to that of a single electron (the Fock-Darwin spectrum; see
Sect. 3.3).

4.1.4 Decoupling of Far Infrared Radiation


and the Relative Motion

The interaction of an electromagnetic wave in the far infrared (FIR) range


with electrons confined in a quantum dot can be described in a dipole approx-
imation (see Sect. 3.4). For the N-electron system, the interaction Hamilto-
nian is the sum of the single-particle Hamiltonians (3.29):

H e- m = L eEOri exp(iwt) , (4.22)

where Eo is the amplitude of oscillations of the electric field. This can be


expressed in terms of the center-of-mass variables (Q = N e is the total electric
charge):
He- m = QEoRexp(iwt). (4.23)
4.1 Generalized Kohn Theorem 31

Since H e - m does not depend on the relative variables, H e - m commutates with


H re1 , and the far infrared (FIR) light cannot directly excite the relative motion
of the system and so affects merely the center-of-mass motion. Therefore, it is
not possible to change the state of the relative motion through the interaction
of the system with the FIR radiation.
Consequently, in the process of absorption (or emission) of the FIR photon
the initial and the final electronic states can differ from one another only
by the center-of-mass excitation. The FIR absorption can be used only for
measuring the center-of-mass excitations, equal to the excitations of a single
electron confined in a dot, whereas the number of confined electrons and the
interaction between them has no influence on the values of resonance energies
(positions of absorption peaks).
This property was first pointed out by Maksym and Chakraborty [101]
and also by Bakshi et al. [6]. However, even earlier Brey et al. had proved
that the FIR resonance frequencies of a two-dimensional parabolic quantum
well in a magnetic field depend neither on the form of the interaction between
electrons nor on their number [17]. This is a generalization of the previous
result by Kohn, who proved that under certain assumptions the cyclotron-
resonance frequency of an electronic system does not depend on the form of
the interaction potential [87].
The inclusion of the electronic spin consists in adding a pair of terms to
the single-particle Hamiltonian (3.8): the Zeeman energy 9fJ,B(}"B (where 9 is
the giro magnetic factor and fJ,B is the Bohr magneton), which describes the
interaction of the spin with the magnetic field, and the term that describes
the spin-orbit coupling. The Zeeman term does not affect the separation of
the center-of mass and relative motions. Moreover, since the absorption of
an FIR photon excites an electron without changing its spin, this term also
has no influence on the FIR resonance energies. On the other hand, the term
that describes the spin-orbit coupling introduces the interaction between the
center-of-mass and relative motions, which are no longer independent. This
breaks the assumptions of the generalized Kohn theorem.
For a single conduction-band electron (the conduction band is built of
atomic s-type orbitals, which have no internal angular momentum) the en-
ergy of the spin-or bit interaction is low. However, the effect of this interaction
increases with an increase in the number of electrons [70] (see also Chap. 9). In
the case of the valence-band holes (the valence band is built of atomic p-type
orbitals, which have an internal angular momentum) the interaction between
sub bands with different orientations of the internal (band) angular momen-
tum and the spin is significant (especially between the pair of light-hole and
heavy-hole subbands, which are close in energy; see Sect. 3.5). Consequently,
the breaking of the Kohn theorem has a strong effect even for a single hole
[8, 18, 19, 31, 32].
32 4. Properties of an Interacting System

4.2 Effect of System Size

Although according to the generalized Kohn theorem the measurements of


far infrared (FIR) absorption have very low sensitivity to the interactions
between electrons confined in a quantum dot, other experiments, described
further in this monograph, allow an important role of the many-body effects
in the system to be unraveled.
In natural equivalents of quantum dots, the many-electron atoms, the
characteristic energy of single-particle excitations in the confining Coulomb
potential of a nucleus is many times larger than the characteristic energy
of interactions between the electrons (due to the large electric charge of the
nucleus, concentrated in a point). As a result, the ground state of the elec-
tronic system in an atom is in the first place determined by quantization of
the single-electron motion in the Coulomb field of the nucleus (spatial quan-
tization of the electron kinetic energy). In atoms with consecutive atomic
numbers, the electrons fill up the subsequent single-particle states according
to the Pauli exclusion principle. The interaction between electrons is not cru-
cial unless a degenerate shell is populated. The electrons in a partly filled
shell form an almost independent system, moving in the average electric field
generated by the nucleus and the lower, completely filled shells. Their orbital
and spin configuration is a result of the minimization of the total interaction
energy, according to the Hund principles:
1. Maximization of the total spin S {:} minimization of the exchange-
interaction energy
2. Maximization of the total orbital angular momentum L {:} minimization
of the direct-interaction energy
3. Maximization or minimization (depending on the sign of the parameter
defining the spin-orbit coupling) of the total angular momentum J =
L + S {:} minimization of the spin-orbit interaction energy
Since the dimensions of quantum dots are visibly larger than those of atoms,
the Coulomb interaction is of much greater significance here. According to
the Heisenberg uncertainty relation, the single-particle excitation energy de-
pends on the size as E: ex 1/ L2, while the Coulomb interaction energy behaves
as Vc ex 1/ L. In very small quantum dots (e.g., self-assembled dots L ;::::20-
40 nm) the relation between E: and Vc is similar to that in atoms, and the
ground state satisfies the (slightly modified) Hund rules (see Sect. 8.2 and
[137]). On the other hand, as shown by Bryant [23] and later by other au-
thors [54, 55, 101, 102, 93], in larger dots the interaction between electrons
determines the ground state of the system. The ground state ,is a result of
the competition between two opposite effects: The Coulomb repulsion favors
pushing the electrons apart from one another, whereas the external con-
finement squeezes electrons in the center of the dot and prevents a large
separation between them.
4.3 Effect of the Magnetic Field 33

4.3 Effect of the Magnetic Field

The significance of the interaction between the carriers is clearly revealed in


the dependence of the energy spectrum of the system on the magnetic field.
The possibility of obtaining very strong magnetic fields (a few T), strong
enough for the cyclotron frequency We = eB/(m*c) to considerably exceed
the resonance frequencies of the confining potential W = E/ti (E is excitation
energy without a magnetic field), enables a continuous transition from the
regime of spatial quantization with excitation energy on the order of several
meV to the regime of magnetic-field (Landau) quantization.
Under the influence of a magnetic field, the electrons are subject to an
additional squeezing in the plane perpendicular to the field. An additional
term ~mw;r2 appears in the Hamiltonian (3.8), and the characteristic length
scale lo, defined in (3.12), shrinks. In strong fields, where the magnetic length
lB becomes smaller than the dot diameter (e.g., in a field B = 1 T we
have lB = 25 nm), a dominating single-particle potential is the parabolic
Landau potential, and lo ;:::; lB. Following the decrease of the dot size, the
characteristic energy of electron-electron Coulomb repulsion varies with the
applied field as e2 /lo ex vB.
On the other hand, the characteristic kinetic energy scale, which is gov-
erned by the noninteracting excitation gap w_ (see (3.16)), decreases with
an increasing magnetic field, from the bare interlevel separation of the con-
fining potential Wo in the absence of a field to zero in the Landau regime (see
Fig. 3.2).
In a quantizing magnetic field and without any additional confining po-
tential, the Coulomb repulsion leads to pushing the electrons apart as far as
possible. The electrons occupy the single-particle states in the lowest Landau
level, with individual angular momenta li limited only by the dimensions of
the sample. However, if a system of electrons is subject both to a strong mag-
netic field B and to a confining potential Vc,onf, a pair of contrary trends meet:
the Coulomb repulsion, with the energy scale of e2 /lo, pushes the electrons
apart; and the confinement, scaled by w_, attracts them toward the center of
the quantum dot. As a result, for a given magnetic field and confinement, the
ground state is a compromise between both these trends. Tuning the mag-
netic field varies the two energy scales and induces transitions in the ground
state [54, 101].
The energy of each eigenstate can be viewed as the sum of the (Coulomb)
interaction energy and the single-particle energy (the average values of the
interaction potential and the noninteracting Hamiltonian, respectively, in a
given state). If the confining potential is in the form of an isotropic 'parabola,
and the magnetic field is strong enough to justify neglecting the mixing with
higher Landau levels and to polarize all electronic spins, then the single-
particle energy is a constant within each total-angular-momentum (L tot , a
good quantum number) subspace: Ltotnw- + const.
34 4. Properties of an Interacting System

Thus the eigenstates of the total Hamiltonian are equal to the eigen-
states of the interaction potential alone, and the inclusion or variation of the
noninteracting terms of the Hamiltonian (i.e., the magnetic field or confine-
ment) leads only to (i) rescaling of the characteristic length lo, which scales
the wave functions and the Coulomb energy, and (ii) adding the appropriate
single-particle energy, which depends on the total angular momentum and,
through the kinetic-energy scale w_ (see (3.16)), on the magnetic field and
confinement.
More importantly, since in the units where the characteristic length lo
and the characteristic interaction energy e21lo are equal to unity, the wave
functions and their Coulomb energies do not depend on the noninteracting
Hamiltonian. Furthermore, since the single-particle energy is constant within
each eigensubspace labeled by the total angular momentum, the variation of
the noninteracting Hamiltonian does not rearrange the energy levels within
each subspace. It only rigidly shifts the entire energy spectra corresponding
to these eigensubspaces with respect to one another. The relative shift of the
energy spectra corresponding to the different angular momenta comes from
the fact that the variation of the magnetic field or confinement changes the
characteristic single-particle energy w_ on the scale of the Coulomb energy
e2 l l o.
The ground state within each angular-momentum subspace does not de-
pend on the magnetic field or confinement. It is simply the ground state of
the interaction Hamiltonian in this subspace (calculated, e.g., by Girvin and
Jach [46]). However, the absolute ground state of the system changes when-
ever it is pushed up above the lowest-energy state from another subspace.

ne=3 8=10T
no Landau level
mixing ,."
,.-'
, ....

20 .,', ,.".. total


.\.".....,.~ ........
> ?-\ ".".

-w
Cl)
"
E \... ,.".(
1,\ ,.-" single
10 "',
I ,'," electron

.,.,."
,../
/a"·

··i· . ·, ......
I

interaction
,
. . .......
Fig. 4.1. Contributions to the
ground-state energy of three elec-
trons as a function of the total angu-
lar momentum J; the confining po-
10 20 tential is parabolic (nwo = 4 meV)
J [101]
4.4 Magic States of a Few-Electron System 35

As shown in Fig. 4.1 for three electrons, the Coulomb energy decreases and
the single-particle energy increases (linearly) with the angular momentum
L tot . Adding the two energies together, One ohtains the total energy as a
function of angular momentum - a dependence with a minimum at a certain
value of L tot . One can choose the energy units of e2 /lo, such that the depen-
dence of the Coulomb energy On L tot is independent of the magnetic field
and confinement. The single-particle energy is linear in L tot , and the slope
is governed by the magnetic field and confinement. Therefore, these two pa-
rameters (magnetic field and confinement) control the minimum-energy value
of L tot . Alternatively, by varying the magnetic field and/or confinement, the
ground state can be driven through a sequence of subspaces, or values of
angular momentum.

4.4 Magic States of a Few-Electron System

Let us concentrate on the ideal case of the parabolic confinement. To be


accurate, the interaction energy of the system does not depend directly on
the total angular momentum L tot , but On the relative angular momenta lij,
which are related to the average distances between the electrons. Due to the
separation of the center-of-mass and relative variables (see Sect. 4.1.2), the
interaction energy is independent of the center-of-mass angular momentum
Lcm. Hence, further considerations can be restricted to the case of Lcm = o.
Due to the identity of the particles, which requires appropriate symme-
tries of the total wave fUnction, not all values of the total angular momen-
tum L tot = L li can be realized under the condition of vanishing Lcm. For
instance, in the case of two spin-polarized electrons, L tot - Lcm has to be
an odd number. In the case of three spin-polarized electrons, it is given by
L tot - Lcm = 3a + 2b, where a:?: 1 and b:?: 0 are integers.
Let us look at the dependence of the total energy on the total angular mo-
mentum L tot , i.e., the total-energy spectrum resolved into the total-angular-
momentum subspaces. Due to the required symmetries of the wave function,
at certain values of L tot , all the low-lying energy levels correspond to the
states of excited center-of-mass (cm) motion. We shall denote these angular
momenta by L;ot + 1. In particular, the lowest-energy state at each of these
values of L tot (in each of these total-angular-momentum subs paces) has its
center-of-mass motion excited. In other words, due to the constraint imposed
on the wave function symmetry, it may not be possible to construct a many-
electron state with the relative angular momentum L;ot + 1 and energy that
is lower than that of the lowest-energy state with the smaller relative angular
momentum (L;ot).
On the other hand, to each state in the L;ocsubspace a center-of-mass ex-
citation can be attached, which increases its angular momentum by One [i.e.,
moves this state to the (L;ot + 1)-subspace]. In the absence of confinement (in
36 4. Properties of an Interacting System

the degenerate Landau level) the center-of-mass excitation energy vanishes,


and the pair of states that differ only by this excitation are degenerate.
The lowest-energy state at a given total angular momentum L tot that has
its center-of-mass motion in the ground state lies below the lowest-energy
state with L tot - 1. In other words, around this value of L tot the total energy
decreases as a function of L tot . This is a natural result, since the total angular
momentum is indirectly related to the total size of the system, and the total
energy is the energy of the electron-electron repulsion.
However, at angular momenta L;ot + 1, where the lowest-energy state has
its center-of-mass motion in the excited state and hence is degenerate with
the lowest-energy state at L;ot, the total energy does not change under an
increase of L tot . In other words, due to imposed constraints, the increase of
the total angular momentum from L;ot to L;ot + 1 does not allow for the
construction of a state with a lower energy (a larger area).
Which values of angular momentum appear in the sequence of magic
states with magic angular momenta L;ot depends on the number of electrons.
For the few-electron systems the magic states were found numerically through
the exact diagonalization of the Hamiltonian, for instance, by Girvin and
Jach [46]. In Fig. 4.1 the three-electron lowest-energy levels in the magic
subspaces are indicated by arrows. If at a given magnetic field the Coulomb
energy (with the magic fine structure) is added to the single-particle energy
(linear in Ltod, the total-energy curve is obtained, with the downward peaks
in the magic angular momenta L;ot. For any combination of the magnetic
field and confinement (provided that the field is strong enough to confine the
electrons in the lowest Landau level) the ground state is one of the magic
states indicated by arrows. Hence, the variation of the magnetic field and/or
confinement does not drive the ground state through all neighboring values
of L tot , but through a sequence of magic values L;ot, and so all other values
in between are skipped.
Figure 4.2a shows the dependence of the ground-state energy of three
electrons on the magnetic field, for both spin configurations, S = ±~ (spin-
polarized system) and S = ±! (spin-unpolarized system), without taking
into account the Zeeman splitting - the spin oscillations are visible: ~ ~ !.
Inserted frames contain full energy spectra determined at a magnetic field of
B = 10 T (left: S = ±~, and right: S = ±!) - a clear division of the two
lowest Landau levels is visible.
Figure 4.2b presents the angular momentum and the projection of the
total spin onto the magnetic field as a function of the magnetic field intensity.
The Zeeman energy, appropriate for GaAs (0.03 meV /T), leads to the spin
polarization of the system at all fields B 2: 2.5 T, except for a small area at
5-6 T. The orbital angular momentum for the polarized system assumes only
the magic values L = 3, 6, 9, 12, ....
Magic states of a quantum dot, due to the occurrence of a nonzero energy
gap in the excitation spectrum, are referred to as the incompressible states
4.4 Magic States of a Few-Electron System 37

5.8"'--'---"1;:::====::::;--:;;:::===::;;::::;'

-
>
II)

.§. 5.6
w
25 25 - _
- --~-
__-

w \ 12
\
>- \
EJ
II) \ I
c
II)

(a)
5.4~---.---.---.---.----.---.---.---,
2 4 6 8 10 12 14 16 18
magnetic field B (T)
~ 15,--------------------------------,
E (b)
.a 12
cII)
9
I
E
I
..
o 6
E
3
.!!!
:::s
OJ
1=-::-£', - - - , I -
04=~~--_r--~--_,--_,~--r_--~--~
- - - - - - - - - - - •
c 2 4 6 8 10 12 18
CIS 14 16
magnetic field B (T)
Fig. 4.2. (a) Magnetic-field evolution of the ground-state energy of three spin-
polarized (continuous line) and spin-unpolarized electrons (dotted line); without the
Zeeman energy. (b) Total angular momentum (continuous line) and spin projection
(dotted line) of three electrons as a function of the magnetic field; with the Zeeman
energy [54J

[93]. A finite energy is required in order to change the state of the system.
For example, a variation of the confining potential at a fixed magnetic field
does not change the area occupied by the electrons unless the ground-state
transition occurs. As noticed by Laughlin [93, 94], the series of magic fill-
ing factors 1I* can be attributed to the series of incompressible magic states.
The sequence of magic ground states, or the magic values of angular mo-
mentum or the filling factor, depends on the number of electrons N but not
on any material parameters or external parameters like the magnetic field
or confinement. The occurrence of the magic series is rather a manifestation
of very special properties of an interacting electron (fermion) system in two
dimensions, which appear also in the fractional quantum Hall effect (FQHE).
38 4. Properties of an Interacting System

Numerical calculations using the exact-diagonalization techniques were


carried out for few-electron systems by Laughlin [93] (three spin-polarized
electrons in the lowest Landau level), Maksym and Chakraborty [101] (three
and four spin-polarized electrons), Hawrylak [54, 55] (two and three electrons,
all spin configurations), and Yang et al. [138] (five and six electrons).
It might be noted here that if mixing with higher Landau levels can be ne-
glected, or if only a finite number of lowest Landau levels is included (which is
absolutely justified in strong magnetic fields), then due to the commutation
between the total angular momentum L tot and the (circularly symmetric)
total Hamiltonian H, the numerical diagonalization can be carried out inde-
pendently in separate submatrices H, each with a finite dimension, labeled
with consecutive values of L tot .
It is possible therefore to obtain numerically accurate eigenstates of the
Hamiltonian, provided that the submatrix dimension does not exceed the
computer's capabilities. This dimension grows with the number of particles
and the angular momentum; in the lowest Landau level, at filling factors
v ::21, the computations can be carried out easily for up to seven or eight
electrons (the dimension of the maximum Hilbert subspace is on the order of
a few thousand).
A basis that spans the many-electron Hilbert space and at the same time
conveniently resolves the total angular momentum L tot can be constructed
from the antisymmetrized products of single-particle states (the Slater de-
terminants). In such a basis the many-body interaction (Coulomb) matrix
elements are easily expressed in terms of the two-body matrix elements. In
the basis of the Fock-Darwin single-particle states (the eigenstates of the
noninteracting Hamiltonian) the two-body Coulomb matrix elements can be
evaluated analytically. The respective expressions are given, for example, in
[132] (see (4.32) for the lowest Landau level). For an arbitrary basis (e.g., for
an arbitrary confining potential) the matrix elements have to be integrated
numerically.
In order to perform the calculation for large numbers of electrons, exclud-
ing the exact diagonalization in a complete basis, one can limit the space
of single-particle states used in construction of the many-body space to the
states expected to give the largest contribution to the actual low-energy eigen-
states (in order to achieve good convergence, the choice of which states are to
be kept or discarded has to depend on confinement and the magnetic field)
and perform the diagonalization in such a truncated basis [60, 61].

4.5 The Idea of Composite Fermions


The interpretation of the incompressible magic states that occur in the spec-
trum of electrons confined in a quantum dot under a strong magnetic field
was given by Laughlin [93] and developed by Jain [73, 77] as an analogy to
the fractional quantum Hall effect. The magic states of strongly interacting
4.5 The Idea of Composite Fermions 39

electrons were there identified with the so-called compact states of weakly
interacting composite fermions (see the monograph by Jacak et al. [69]).
In the circular representation the presence of.a parabolic confining poten-
tial leads only to the rescaling of a magnetic field,

(4.24)

(i.e., to the rescaling of the length unit, according to (3.12)), but does not
change the form of single-electron wave functions. Hence, a spatially con-
strained system in the magnetic field B is equivalent to a free two-dimensional
system in the effective magnetic field B.
The transformation of an electron system (ES) into a composite-fermion
system (CFS) is performed by attaching to each electron an equal and even
number of elementary fluxes of a magnetic field. This can be achieved formally
by multiplying the ES wave function by the so-called Jastrow factor

lfh = p II (Zj - Zk)2m pL *, (4.25)


j<k

where N is the number of particles, pI is the wave function of the noninter-


acting ES with total angular momentum L *, P L is the CFS wave function
with total angular momentum L = L * + mN (N - 1), and P is the operator
projecting onto the lowest Landau level.
Let us define, following Jain, the compact states of the noninteracting ES,
in which the electrons occupy only the orbitals with the lowest values of the
angular momentum (internal orbitals), allowed by the Pauli exclusion prin-
ciple, in the lowest s + 1 Landau levels. The compact state is hence defined
by the system of numbers [No, N l , ... , N s], I:: Ns = N; the electrons occupy
the Ns lowest levels within the sth Landau level. The compact CFS state,
denoted by [No, N l , ... , NslcFl is a result of the transformation (attaching
fluxes) of the corresponding compact ES state [No, Nl, ... , Nsl. Among oth-
ers, the so-called Laughlin states [NlcF belong to this class.
As shown in the exact numeric calculations [73, 77], the eigenstates of the
interacting ES are very close to the respective noninteracting compact states
CFS. For instance, for N = 6 a sequence of ground states with L = 35, 39,
and 45 gives the following overlaps with the appropriate compact CFS states:
0.991 for [4,21 cF , 0.994 for [5, 11cF, and 0.986 for [6,01cF' This means that
the interaction between electrons is the reason for the formation of composite
fermions.
The electronic system described in terms of composite fermions is an ef-
fectively free (noninteracting) system. It should, however 1 be stressed that the
system of noninteracting composite fermions is a many-body system. Each of
its quantum states is nonlocal and differs strongly from the antisymmetrized
product of single-particle electronic states, i.e., the appropriate Slater deter-
minant.
40 4. Properties of an Interacting System

Jain's idea elegantly explains the peculiar behavior of a system of two-


dimensional electrons confined in a quantum dot and allows one to determine
with high accuracy the ground state of interacting ES (as a compact CFS
state with the lowest energy).

4.6 Harmonic-Interaction Model

Johnson and Payne [75] employed the approximation of a harmonic interac-


tion between electrons

V(Ti, Tj) = 2Vo - ~m*.n2ITi - Tj12. (4.26)

The approximation was justified by a finite broadening of electron wave func-


tions across the quantum well. If the electron positions T i, T j are treated as
two-dimensional vectors in the plane of the well, this broadening leads to
the saturation of the interaction at small distances: V(Ti,Tj) -> 2Vo < 00.
Taking into account the broadening of wave functions in the z-direction, in
the form of a factor B(z), the operator of the Coulomb interaction that acts
on two-dimensional wave functions of a variable T == (x, y) can be written as

V(TI - T2) = -E
e2 J J
dZ I dZ 2 1
ZI -
B*(ZI)e(Z2)
12 1
Z2 + TI - T2
12 < 00. (4.27)

Moreover, Johnson and Payne wrote that the potential of the actual interac-
tion differs from the bare Coulomb potential (ex: ITi -Tjl-I) due to screening.
The parameters Vo and .n should be chosen depending on the dimensions of
the quantum dot. A discussion of this model is also presented in [74]. In such
an approximation the Hamiltonian of the relative motion of N electrons con-
fined in a parabolic dot can be diagonalized analytically, and the obtained
ground state is a generalization of Laughlin's function. The excitation spec-
trum of the system was determined in [76] as a function of the value of the
parameter .n.
A substantial difference, which was introduced by replacing the Coulomb
l/r-type interaction by the harmonic interaction, is evident in the appear-
ance of additional degeneracies: degenerate now are the subspaces of states
with given values of the total angular momentum L tot and the center-of-mass
angular momentum Lcm. Denoting the total angular momentum of the rela-
tive motion of electrons by Lrel = L tot - L cm , in a strong magnetic field (in
the lowest Landau level) the operator of the total harmonic interaction that
acts on the N-electron state can be written as V = 2NVo - ~m*.n2 D2, where

D2 == L ITi - Tjl2 = constLrel· (4.28)


i<j
Hence, the total Hamiltonian can be expressed as
H = clLtot + C2Lrel + C3, (4.29)
4.7 Reconstruction of the Edge of a Compact Droplet 41

where Cl, C2, and C3 are constants. The Hamiltonian is therefore diagonal in
the basis of any eigenfunctions of L tot and Lrel:
(4.30)
Thus in the case of three electrons, for example, instead of Laughlin's magic
state with angular momentum L tot = 9 (v = !)[93], in the harmonic-
interaction approximation a pair of degenerate states is obtained spanning a
two-dimensional space: L tot = Lrel = 9.

4.7 Reconstruction of the Edge of a Compact Droplet


Numerical analysis of the quantum-dot models has shown that in a system
of many (>10) electrons confined in a quantum dot in a strong magnetic
field the transitions between the magic states are related to consecutive re-
constructions of the edge of the system, while the density of electrons inside
the dot remains unchanged [60, 61]. At relatively strong fields, all the elec-
trons are spin-polarized and occupy the lowest Landau level. It was found
that over a wide range of the magnetic field the confined electrons form a
compact droplet of maximum density, occupying subsequent orbitals with the
lowest available values of angular momentum [27, 28, 60, 61, 85, 100, 106].
The values of the magnetic field and electron density are similar to those in
the case of the lowest completely filled Landau level (filling factor v = 1) in a
bulk material. In the lowest Landau level the single-particle (noninteracting)
excitation energy from the state of a compact droplet is a linear function
of angular momentum. This configuration is hence a simple realization of a
chiral Luttinger liquid [27, 28, 78, 60, 61]. The interaction between electrons
leads to the occurrence of a minimum in the dependence of excitation energy
on the angular momentum. The corresponding excitation is then referred to
as a magneto-roton. The magneto-roton energy decreases with an increase of
the magnetic field. After reaching the critical value, a transition in the ground
state (droplet reconstruction) occurs, to another ground state with a smaller
filling factor. In the case of a smaller number of electrons, a compact droplet
breaks, forming a ring; i.e., the transition consists in introducing the holes
(vacancies) into the bulk of the droplet. For a larger number of electrons, the
compact droplet is replaced in the transition by the system composed of a
smaller droplet surrounded by a ring, with the holes (vacancies) introduced
not into the center, but rather near the droplet edge. Such a transformation
is then called the edge reconstruction. A further increase of the magnetic field
drives the system through a sequence of incompressible states (with nonzero
excitation energy) with increasing angular momentum.
In the model discussed, the electrons are confined in a two-dimensional
parabolic potential of a quantum well. A magnetic field is applied perpendic-
ularly to the plane of the well. The single-particle (noninteracting) Hamilto-
nian has form of (3.8), with its eigenstates in the form of the states of a pair
42 4. Properties of an Interacting System

of uncoupled harmonic oscillators with the pair of frequencies w±, multiplied


by the spin function In+n_O') [55, 54] (see Sect. 3.3). Concentrating here on
the case of strong magnetic fields (w+ ~ We »w_), we assume that all elec-
trons are spin-polarized and confined to the lowest Landau level. The space
of single-particle states is hence restricted to the states 1m) == 10, m, 1), which
are uniquely defined by the angular momentum m, with energies (disregard-
ing the zero-motion energy) em = mw_.
Denoting the operators of creation and annihilation of an electron in the
state 1m) by c;t and Cm, respectively, the Hamiltonian of the N-electron
system can be written in the form

(4.32)

where Eo = .j2;=Ry* aB/lo is the energy scale, (Ry* = e2 /2Ea Bis the effective
Rydberg, a B = di 2 /m*e 2 is the effective Bohr radius, the length scale to is
defined by (3.12), and P = ml + m2 - PI - P2 (compare (8.24)).
Due to the circular symmetry of the quantum dot, the total angular mo-
mentum L = Li mi of the system is a good quantum number, which labels
the eigenstates. In a compact droplet composed of N + 1 electrons, N + 1
single-particle states with the lowest available angular momenta are occupied

IG N + I ) = (flo c~ ) Ivac) , (4.33)

where Ivac) stands for vacuum, and the total angular momentum of this
compact state is Lo = ~N(N + 1). Since (in the lowest Landau level, with
the total polarization of spins) it is not possible to construct another state of
N + 1 electrons with this angular momentum L o, the compact state (4.33) is
an exact eigenstate of the interacting system. Further in this section we shall
assume that the magnetic field, confinement, and the number of electrons
assume such values that the ground state of the system is the compact state.
Following the bosonization procedure described in [27, 28, 60, 61, 126]
we can construct the excited many-electron states with angular momen-
tum L = Lo + M by moving the electrons from the inside of the com-
pact droplet (m :::; N) to the outside (m > N) or, alternatively, creating
4.7 Reconstruction of the Edge of a Compact Droplet 43

the two-particle electron-hole excitations. In the basis of elementary excita-


tions Ih k1 ) = C~ Ckl IG N +1) (one pair, let us denote this state by IP) and
Il2hk2kl) = c~c~ Ck2Cki IGN+t} (two pairs, 2P), the eigenstate with index p
will be written as
(4.34)
12>II,k2>kl
12+h-k2-k1=M

In the above equation the electrons are excited to states li > N outside the
droplet and leave the holes in states k i ::; N inside the droplet. The values of
coefficients A and B, i.e., the full excitation spectrum of the system of N + 1
electrons, are obtained through the numerical diagonalization of the Hamil-
tonian performed in the basis of IP and 2P excitations. The diagonalization
in the basis consisting of IP and 2P excitations is accurate for M < 9, since
at M = 9 the first three electron-hole pair (3P) excitation occurs. For M 2: 9
this procedure is an approximation, since the diagonalization is performed in
a truncated basis.
The approximation that consists in a numerical diagonalization of the
Hamiltonian in a truncated (finite) basis (e.g., the so-called configuration
interaction method) is equivalent to the variational method - with the trial
function in the form of a linear combination of basis functions. However, the
diagonalization enables an easy improvement of accuracy by increasing the
basis size. This can be done by adding consecutive functions and testing the
accuracy through studying the convergence of results with the growth of the
basis size.
Besides this, and in contradiction to the variational method, as a result
of the diagonalization of the Hamiltonian, the complete energy spectrum, to-
gether with the complete orthogonal set of eigenfunctions, is obtained. With
the appropriate selection of the basis, a number of functions obtained in
this procedure can be good approximations of the exact eigenfunctions. For
instance, when the basis is constructed by choosing the lowest-energy eigen-
states (states below the cut-off energy) of a simplified Hamiltonian (e.g.,
the noninteracting Hamiltonian), the accuracy of the full-Hamiltonian eigen-
functions calculated in the diagonalization procedure can be controlled by
sweeping the critical cut-off energy. The best approximation in such a case is
obtained for the ground state and the low-energy excited states.
Matrix elements of the Coulomb interaction in the basis of IP and 2P exci-
tations are expressed through the two-body matrix elements in the following
way:
- IP-IP matrix element, which describes a single magneto-roton:
(k~ l~ IVeelhk 1) 81~ 11 8k~ kl {E~F - E~F}
+{(l~klIVeelhkD - (l~klIVeelk~h)}, ( 4.35)
where the Hartree-Fock energy is defined as
44 4. Properties of an Interacting System

EfIF = 2) (lmIVeel ml ) - (lmIVeel lm)). (4.36)


m

The diagonal part of a matrix element defines a difference in the Hartree-


Fock energy of an electron in the state h (above Fermi level) and of holes
in the state kl (inside a droplet). The off-diagonal part defines a difference
between the energy of the attractive direct interaction and the repulsive
exchange interaction.
- 2P-2P matrix element, which describes the energy of the magneto-roton
pairs, including possible bound states:
(k~ k~l~ l~IVeell2hk2kl)
OI;1201~llOk;k20k~kl {En F + E~F - E~F - E~F}
+{ Ok~ kl Ok;k2 - Ok~ k20k;kJ{ (l~l~ IVee Ih l 2) - (l~l~ [Veel l 2h)}
+{OI~llOI;12 - OI~1201;zJ{(klk2IVeelk~kD - (klk2[Veelk~k~)}
+OI~1t Ok~kl {(l~k2IVeell2k~) - (l~k2IVeelk~l2)}
+Ol~ It Ok~ k2 {(l~klIVeelk~h) - (l~klIVee 1l2k~)}
+OI;120k~k2{(l~klIVeelk~h) - (l~klIVeelllk~)}
+OI;120k;k2{(l~klIVeelhkU - (l~klIVeelk~lt)}. (4.37)
This matrix element defines the Hartree-Fock energies of two electron-hole
pairs, as well as the direct and exchange interaction between them.
- 1P-2P matrix element, which describes the decay of a magneto-roton pair
into a single magneto-roton:
(k~ l~ [Veelhhk2k 1 )
ok~kl{(l~k2IVeelhl2) - (l~k2IVeell2h)}
+ok~k2{(l~klIVeelhlt) - (l~kl[Veelhl2)}
+Ol~ d (k2klIVeelk~ l2) - (klk2IVeelk~ l2)}
+ol~12{(klk2IVeelk~l2) - (k2klIVeelk~l2)}' (4.38)
Figure 4.3 presents the evolution of the energy spectrum of a system of 15
electrons in an increasing magnetic field. The spectrum is decomposed into
subspectra corresponding to different values of the total angular momentum
L tot . Three diagrams correspond to the three values of the magnetic field.
A quantity M is defined as M = L tot - Lo, where Lo corresponds to a
filling factor 1/ = 1. In the state 1/ = 1 all electrons are spin polarized in
the lowest Landau level, which occupies the orbitals with the lowest values
of angular momentum allowed by the Pauli exclusion principle. The total
angular momentum in this state is Lo = 0 + 1 + 2 + ... + (N - 1). As
shown in Fig. 4.3a, over a certain range of the value of the magnetic field
(B ~ 2 T) the ground state of the quantum dot is the compact droplet
(M = 0). The excitations of the system correspond to the creation of pairs of
quasi-particles (with positive angular momenta, measured by M): an electron
4.7 Reconstruction of the Edge of a Compact Droplet 45

0.4 (a)
N=15
0.3 - ~ -
o
W
UJ 0 .2

0.1

0.0 B = 2.0 T
0.4 (b)
0 .3
o
W
....... 0.2 Fig. 4.3. Evolution
W

--
0.1
of the excitation spec-
trum of a parabolic
0.0 B = 2.5 T quantum dot contain-
ing 15 electrons in
(c) a magnetic field: (a)
0 .2 B = 2.0 T, the ground
state is a compact
o 0.1 droplet with M = 0;
W
....... (b) B = 2.5 T, a mini-
W 0.0
mum develops at M =
-0.1 7 (magneto-roton); (c)
B = 3.0 T B = 3.0 T, the ground
-0.2 ..J.,-.,...,....,...,....,...,....,...,....,...,....,...,....,...,....,....,...,....,....,...,....,...,....,...,....,....,............... state has M = 24
o 5 10 15 20 25 30 (two bound magneto-
M rotons) [61]

outside and a hole inside the droplet. With an increase of the magnetic field
(Fig. 4.3b) we observe the appearance of a characteristic minimum in the
dependence of the excitation energy on the angular momentum (M = 7 for
B = 2.5 T). Figure 4.4a presents the occupation coefficients of the single-
particle orbitals ((c;;:"Cm) ~ radial charge density), which correspond to the
first excited state. Such excitation, which is located near the system edge,
is referred to as the edge magneto-rotan. A further increase of the magnetic
field results in a decrease of magneto-roton energy (due to the increasing
energy of repulsion between the electrons, compared to the difference between
kinetic energies of neighboring single-particle orbitals). For a certain field
value (B ~ 3 T) this energy becomes negative, and a change of the dot ground
state occurs (binding of a magneto-roton inside a droplet). This transition,
occurring in larger dots, is equivalent to a transition between subsequent
magic states in the few-electron dot. Further transitions in the ground state of
the quantum dot, induced by increasing the magnetic field, correspond to the
binding of subsequent magneto-rotons, i.e., the decrease of the filling factor v.
Figure 4.3c presents a situation where the droplet binds two magneto-rotons
46 4. Properties of an Interacting System

B = 2.5 T
M=7

(a)

(b)
Fig. 4.4. Occupation coeffi-
cients of the single-particle or-
bitals with angular momentum
m, for the magneto-roton min-
ima from Fig. 4.3, (a) M = 7
and (b) M = 24; empty bars -
o 6 10 16 20 the compact droplet with M =
m 0[61]

in the ground state (B = 3 T, M = 24). The corresponding diagram of the


occupation coefficients of single-particle orbitals is shown in Fig. 4.4b.

4.8 Heat Capacity

As shown in Sect. 4.1, the measurement of the FIR light absorption does not
allow for the investigation of interactions between electrons. The quantities
that can be measured (at least in principle) and in which the occurrence of
magic values of angular momentum is reflected, are, for instance, the ther-
modynamic quantities such as the heat capacity and (see the next section)
magnetization. The electronic heat capacity at a constant volume C v , which
is defined as the derivative of the average energy with respect to temper-
ature d (H) /dT, was calculated by Maksym and Chakraborty [101]. Their
results for the system of three and four electrons are shown in Fig. 4.5. In
the calculations only the lowest Landau level was taken into account, and the
Zeeman energy splitting was neglected. However, as the authors claim, this
had no qualitative influence on the results obtained. The confining potential
was assumed to be a parabolic well with characteristic energy 1lv.;o = 4 meV.
The dotted line in Fig. 4.5 represents the result for the noninteracting sys-
tem. The oscillating curve drawn with a solid line shows the results with the
inclusion of interactions. The difference between the two is considerable. The
4.9 Magnetization 47
0.3r------------,O.2r-----------,
n=4 T=3K J=22 ~ n=3 T=3K
i' 0.2
:> ,r---
GI 0.1
E
-;0.1
o
0.2:=::=~======:::, 0.2~=======~
n=4 T=1K J=22 •
n=3 T=1K
ro----.l
J=12
r---- ..
.r-------""
0.1 J=9 •

Fig. 4.5. Heat capacity of three and four electrons in a parabolic quantum dot
[101]

dashed step-like curve shows the values of angular momentum in the ground
state of the interacting system.
At a lower temperature Cv is close to zero, far from the critical magnetic-
field values B* for which a transition to the state with another magic angular
momentum L;ot takes place. It then increases in the vicinity of B* and falls
back to zero at the field equal exactly to B*. The reason for this behavior
is that at a very low temperature the contribution to the thermodynamic
average comes mainly from the two eigenstates with neighboring values L;ot.
At the magnetic field far from B*, a wide energy gap forbids thermal ex-
citation, and the system remains practically in the ground state. On the
other hand, when the field equals any of the values B*, the energies of both
strongly contributing states cross each other, and their thermal mixing does
not change the average energy. The only situation where considerable values
of heat capacity are observed occurs when the difference between the two
lowest energies is on the order of kBT.
At a higher temperature, where the thermal broadening is wide enough
for the thermal mixing of states with higher energies (at any value of B),
the shape of the heat-capacity curve changes slightly. However, the minima
remain at the transition values B*, for which the separation between the
ground-state energy and the next differing energy is the greatest.

4.9 Magnetization
In [102]' Maksym and Chakraborty present the dependence of the magnetiza-
tion M on the magnetic field B. It was calculated for quantum dots containing
three and four electrons. The magnetization can be calculated in two ways:
It is possible to determine the matrix elements of the magnetization operator
48 4. Properties of an Interacting System
2 r------------------------,
j::'O
0
3;

-
CI)
-1
E
-2
~ -2

-3 -4
10 10

..,
B
12
8
B
4
4
2
0 0
2
en I
I
0 0
0 II 10 16 0 5 10
B (T) B (T)

Fig. 4.6. Dependence of magnetization M, total angular momentum J, and total


spin projection S on a magnetic field B for three (a) and four (b) interacting
electrons in a quantum dot; M is the thermodynamical average at temperature
T = 0.1 K, and J and S are the ground-state values [102J

M = --e L r· X (p. + eA) (4.39)


2m* .' • •
or to differentiate the eigenenergies with respect to the magnetic field. In a
finite basis both methods are equivalent. However, the latter is more suitable
for numerical calculations, where it gives more accurate results at the same
limitation of the basis. In order to obtain the thermodynamic average
M(T) = (M), (4.40)
in the cited work first the free energy F was calculated, and then it was
differentiated with respect to the magnetic field. The obtained dependence
of magnetization on the magnetic field, together with the changes in the
ground-state total angular momentum and spin, is presented in Fig. 4.6. The
temperature used for the calculation was 0.1 K and the confining potential
was the parabola with characteristic energy !U.u o = 4 meV. A dot-dashed line
shows the magnetization of a noninteracting system.
The basis used for the diagonalization of the Hamiltonian 'consisted of
the states in which at most one electron occupied a state in the first excited
Landau level while the other remained in the lowest (ground) Landau level:
:L ni :::; 1. As the authors claim, the obtained accuracy is very good. The inset
in Fig. 4.6a presents a comparison between the main result for :L ni :::; 1 (solid
line) and the result for :L ni :::; 2 (dotted line).
4.9 Magnetization 49

In order to understand the reason for the dramatic difference between


the magnetization of the interacting and noninteracting systems shown in
Fig. 4.6, we should write out, following Maksym and Chakraborty, the
ground-state energy of the system of N electrons (for simplicity, spin po-
larized and limited to the lowest Landau level):

E = Ltot tu.v_ + ~N li(w_ + w+) + Vint(Ltot ) + g* JlBBa z. (4.41)

The frequencies w± have been defined in Sect. 3.3, and Vint is the interaction
energy in the ground state. In the case of GaAs, for which the calculations
were done, the spin contribution is very small (LlE = 0.03 meV /T, which
gives a 1 % contribution to M) and can be neglected. The first two components
describe a noninteracting system and after differentiation give a continuous
magnetization curve, marked in Fig. 4.6 with a dot-dashed line. The third
component, which describes the interactions, gives only a small correction to
M. However, due to the sharp changes of L tot (and hence also of Vind, this
correction is discontinuous and clearly visible at low temperatures.
Very similar dependencies to those presented in Fig. 4.6 for N = 3 and 4
were also demonstrated for N = 2 by Wagner et al. [131]. The magnetization
of a many-electron quantum dot and its connection with the spin-orbit inter-
action was also studied by Jacak et al. [70] (see Chap. 9). In the framework
of the Hartree-Fock approximation the analytical expression for total energy
as a function of a magnetic field was found. The respective dependencies are
shown in Figs. 9.2-4. For a magnetic field B at which the magnetization van-
ishes, Be/BB = 0, the transition between the paramagnetic and diamagnetic
states occurs. For a GaAs dot (tu.vo = 5.4 meV) a critical field is ;:::: 2 T, which
is in good agreement with the experimental results of Ashoori et al. [4].
5. Intraband Optical Transitions

5.1 Relation with the Kohn Theorem


The characteristic single-particle excitation energy of typical quantum dots
lies in the far infrared (FIR) range. Depending on the dot dimensions, this
energy varies between a fraction to tens of meV. As described previously, in
Sect. 4.1, due to the fact that the wavelength of light from this range (~ 1
mm) considerably exceeds the dot diameter « 1p,m) , the FIR radiation can-
not couple directly to the internal (relative) motion of the system of confined
electrons, but only to their center-of-mass motion. Indirect interaction of the
FIR light with the relative motion of electrons, which would allow for the
observation of the effects of electron-electron interactions in the FIR exper-
iment, requires a coupling between the relative (reI) and the center-of-mass
(cm) motions.
Let us first assume that the spin-orbit coupling can be neglected. The pair
of corresponding Hamiltonians for a free (without the lateral confinement)
two-dimensional system, Hem and H re [, commutate. This result is also not
affected by the presence of a perpendicular magnetic field (see Sect. 4.2.1).
Hence, only the lateral potential that confines the electrons in the quantum
dot can introduce a cm-rel coupling. However, the potentials that define real
quantum dots are typically close to a circularly symmetric parabolic well
(even if the quantum dot itself is not exactly circular) due to the deflection
of the band edge at the system borders, and the parabolic potential does
not introduce the cm-rel coupling (see Sect. 4.1.2). As a result (neglecting
the spin-orbit coupling), the positions of absorption peaks correspond to a
reasonable approximation to the excitation energies of a noninteracting sys-
tem (the Fock-Darwin levels; see Fig. 3.2). These energies depend neither on
the number of confined electrons nor on the presence or form of the inter-
actions between them (however, the number of confined electrons affects the
intensities of peaks).
It should be mentioned, however, that a weak dependence of 'resonance
FIR energies on the number of electrons, which was observed in experiments
(e.g., [63]), shows the presence of a slight cm-rel coupling in real systems
(see Fig. 5.1; the slight shift of the positions of the minima on the consec-
utive curves demonstrates a weak breaking of the Kohn theorem). It is also
visible in Fig. 5.5, where the values of resonance energies in the two graphs
52 5. Intraband Optical Transitions

0.99
::>
t:: 0.98 Fig. 5.1. Transmission of the
>C!J FIR light of wavelength 118 J-tm
i=" (Le., energy 10.5 meV) for dif-
0.97 ferent gate potentials VG. On
-0.720············
average, three electrons per dot
0.96 -0.715 •...........•. are bound at VG = -0.715 and
-0.720 V, and two electrons per
5.95 6.00 6.05 6.10 dot are bound at VG = -0.730
Magnetic field (8) and -0.735 V [63]

(which correspond to various numbers of electrons in the same quantum dot)


differ considerably: In the absence of the magnetic field (B = 0) they are
23 meV for N = 25, and 32 meV for N = 210. The reason for such cou-
pling seems to be the spin-orbit interaction. The inclusion of the spin-orbit
interaction (see Chap. 9) leads to the dependence of resonance energies on
the number of electrons, and also to their splitting: The two features were
very clearly demonstrated in the experiment of Demel et al. [34] (see the
description on page 54). The inclusion of the spin-orbit interaction allows
also for the explanation of the characteristic transition between the param-
agnetic and diamagnetic states. This was observed by means of capacitance
spectroscopy in the experiment of Ashoori et al. [4] (see the description on
page 88). Significant is the fact that a quantitatively good description of the
two independent effects, observed by means of two different techniques (FIR
absorption and capacitance spectroscopies), was reached at the same value of
the spin-orbit coupling constant. Moreover, this value is on the same order
of magnitude as that characterizing the atoms of which the dot is composed.
Let us also mention that an example of a quantum dot where the Kohn
theorem is broken due to the nonparabolicity of the confining potential is
the system of electrons bound to an ionized donor D+ by the bare Coulomb
potential [56]. A positively charged donor D+ can bind one or two electrons,
forming a charge-neutral donor DO or a negatively charged donor D-.

5.2 Measurements of Far Infrared Absorption

Information regarding the first measurements of FIR absorption in a system


of quantum dots was published by Sikorski and Merkt [122]. The system in-
vestigated was composed of a regular array of about 108 dots. The dots were
created on the surface of InSb using electrodes of special shape, and their di-
mensions were 3x3 mm 2 (see Fig. 2.5). The advantage of the InSb material is
the small effective mass of a conduction electron (0.014me) and, consequently,
a large separation between the single-electron energy levels of the dot. The
5.2 Measurements of Far Infrared Absorption 53

0.3 AVgM firo=10.4meV

0.2

0.1

(a)
0.0 ! - - - - - ' - - - - ' - -......- -.......- -......- - i
0.3

-
-
~

0.1
p-lnSb (111)
T=4K

(b)

firo=3.2meV

Fig. 5.2. Dependence of the


0.1 FIR absorption spectrum of
(c) the quantum-dot array on
the magnetic field; the max-
0.0 LJ3=-,-_-=:i:::::::::::::C==::J::~~!!d ima correspond to the reso-
o 0.5 1.0 B (T) 2.0 2.5
nance transitions [122]

number of electrons in each dot (0-20) was controlled by changing the gate
voltage. Below the threshold (turn-on) voltage VT = -98 V the dots were
empty. The measurements were carried out at helium temperature :::::0 4 K),
and at magnetic fields of 0-4 T. The measured quantity was a relative drop
in the transmission of light through a sample containing the quantum dots
populated with electrons (gate voltage VG > VT), compared to the drop of
transmission through the sample with empty dots (gate voltage VT ):
T(VG) - T(VT )
t= (5.1)
T(VT)
The energy of the laser beam f1w was fixed, while the magnetic field intensity
was increased continuously. The sample graphs published in [122] are shown
in Fig. 5.2. The maxima at the field B = 1 T on the curves for. the fixed
energy of f1w = 10.4 meV correspond to the resonance excitation energy of
10.4 meV in the field of 1 T. The average numbers of electrons per dot for
three different values of the gate voltage VG = 3, 8, and 18 V are N = 3, 9,
and 20, respectively. The identical positions of maxima for different values
of VG indicate that the actual confining potential is parabolic. The absence
54 5. Intraband Optical Transitions

30>15D13
16 n
.. .
20
10
Q)

~
S 5

..r::.

~ 20 00 5 15
.§.. J)"V (V) Fig. 5.3. Dependence of the FIR
p-lnSb(111) resonance energies of a quantum
8 T=4K
..s::: dot on the magnetic field; points
!Ng=8V
-EMA
- the experiment; lines - the
Fock-Darwin energies w+ and
w_. Inset: dependence of the FIR
resonance energy on the number
00~--~10~-===::::::::~3~0::==~40'
B (T)
of electrons (without a magnetic
field) [122]

of maxima on the curves for the energy /U.;J = 7.6 meV comes from the fact
that the magnetic field B for which 7.6 meV is the excitation energy is close
to zero (as can be seen in Fig. 5.3). On the curves drawn for /U.;J = 3.2 meV
the maxima appear again, at B = 1.5 T, as a mode with lower energy.
Demel et al. [34] present the results of measurements carried out on the
array of larger dots, each containing 25-210 electrons. The sample investi-
gated was etched from a modulation-doped GaAs/ AlGaAs heterostructure;
the quantum dots were created inside the columns in the shape of 600 nm
wide squares with rounded corners. Due to the deflection of the band edge
at the boundaries, the actual diameter of the area in which electrons were
confined was considerably smaller than that of the column, and depending
on the number of electrons varied between 100 and 160 nm. The Fermi level
in the obtained structure was below the first discrete energy level in the dots.
By controlling the intensity of illumination falling on the sample, the average
number of electrons per dot N (excited to the conduction band by the falling
photons) was varied between 0 and 210. The value of N was determined from
the intensity of absorption. The area from which the signal was collected was
3 x 2 mm 2 and contained about 107 dots. The measurements were carried out
using the Fourier spectrometer, at a magnetic field fixed in the range 0-15 T,
and a temperature of 2.2 K. The measured quantity was the transmission of
light through the sample. The reported absorption spectra for N = 210 are
shown in Fig. 5.4, and Fig. 5.5 gives the corresponding dependence of the
excitation energy on the magnetic field.
An interesting result, visible in Fig. 5.5, is the splitting of the resonance
spectrum and the anticrossing of respective energy curves. According to the
authors of the original work [34], the plasma excitations in the system may be
the reason for this effect. Chakraborty et al. [25] suggested that similar behav-
ior may result from the Coulomb interaction between neighboring dots, which
breaks the rotational symmetry of a single dot (the numerical calculations
presented were made for a pair of coupled dots, each containing three or four
5.2 Measurements of Far Infrared Absorption 55

100~ ~r /
100 ~Ylil- l2+
12.4T
ir-r
100 'I'1f1.5T 5.4T B=
...... 8.4T 11.4T
~
e....100
;if> ~T
~
W
100
m
i="

\1
T

100

k~
Fig. 5.4. FIR absorption spectra of
99
the array of quantum dots, for different
magnetic fields. ES in the schematic of
o 50 100 150 200 the sample shows the electron localiza-
Wave number (cm- 1) tion areas [34J

200 (a)

150

---'~
....

---
8
100

50
N=210
R=160nm

200 (b)

150
......-

---
'8
~100

N=25
8 R=100nm
50
Fig. 5.5. Evolution of
the FIR resonance ener-
°0 2 4 8 8 10 12 14 16
gies in a magnetic field:
(a) 210 electrons, (b)
Magnetic Field B (T) 25 electrons [34J
56 5. Intraband Optical Transitions

electrons). However, in the cited experiment of Demel et al. the large distance
between the neighboring dots (R:: 1J.Lm) seems to exclude this explanation. A
different interpretation of the observed form of FIR spectrum was given by
Gudmundson et al. [49], in which the higher-energy modes were interpreted as
the following transitions between the Fock-Darwin levels: (n, m) - 7 (n, m+2)
and (n, m) - 7 (n + 1, m + 1) (see Fig. 3.2). For a perfectly parabolic confining
potential V(x, y) these transitions are forbidden (see Sect. 3.4). However, as
shown in [49]' the observation of these forbidden transitions may be a result
of deviations from the ideal parabolic confinement (the anharmonic terms).
A similar interpretation was also given by Shikin et al. [120].
It seems, however, that both the splitting of modes and their anticrossing
are the manifestation of the spin-orbit interaction (see Chap. 9). As men-
tioned previously, such an interpretation is also supported by the fact that
for the same value of the spin-orbit coupling constant, the proposed model
provides a very good description of the dependence of energy (and the chemi-
cal potential) ofthe system on the number of electrons and the magnetic field.
These dependencies were measured independently by Ashoori et al. [4] (see
also the description on page 88). As shown by Darnhofer and Rossler [31] on
the example of two electrons confined in an InSb quantum dot, the inclusion
of the spin-orbit interaction within the band model leads to a qualitatively
similar excitation spectrum.
Alsmeier et al. [1] describe measurements carried out on the array of dots
created on the surface of silicon. The confinement of electrons was provided
by a pair of electrodes placed over the silicon surface, as shown in Fig. 2.6.
The lower electrode had regularly spaced, round holes that had a diameter
of 150 nm. The application of the appropriate voltages VCT (top-gate volt-
age) and VCB (bottom-gate voltage) resulted in the creation of a spatially
modulated electric field. The advantage of the system described is the possi-
bility of continuous, independent control of the dot diameters (in the range
40-150 nm) and the number of electrons per dot (in the range 20-350). The
measurements were carried out at a temperature of 2 K and at a constant
magnetic field of intensity 0-12.5 T. Similarly, as in the work described in
previous paragraphs, the directly measured quantity was the relative drop of
the transmission of light through the sample. The curves obtained for 350
and 140 confined electrons are presented in Fig. 5.6. The maxima correspond
to the resonance excitation energies. Also, the evolution of these energies in
a magnetic field is visible. This evolution roughly follows that of the lowest
Fock-Darwin levels shown in Fig. 3.2.
Meurer et al. [105] describe the application of the FIR absorption mea-
surement in the determination of the charging energy of the quantum dot (the
energy required to add a single electron to the dot). The sample investigated
consisted of an array of quantum dots, obtained from a single GaAs/ AIGaAs
quantum well by the application of a voltage Vc to the electrode covering
the square, about 100 nm high cubes of GaAs created on the surface of the
5.2 Measurements of Far Infrared Absorption 57

(a) (b) Vgt =30V, Vgb=-2.25V


Vss=18V B (T)

o
~

~o
!::: o
!<i
I 0
o

o o

o
o
o
o

30 50 70 90 110 130 30 50 70 90 110 130


wave number (cm-1) wave number (cm-1)

Fig. 5.6. FIR absorption spectra of an array of quantum dots containing (a) 350
and (b) 140 electrons at various magnetic fields [1]

quantum-well structure. The varying distance between the electrode and the
quantum well led to the creation of a modulated electric potential, laterally
confining electrons in regularly spaced dots. The number of electrons in a
single dot N = 1, 2, 3, 4 was controlled by a change of voltage Vc. The
transmission of light through the sample was measured at a temperature of
1.2 K using a Fourier spectrometer. Figure 5.7 presents a characteristic de-
pendence of the excitation energies on the magnetic field obtained for four
electrons confined in each dot. It should be noted that contrary to the result
of Demel et al. [34], a splitting of the spectrum was not observed here. This
is in accordance with the interpretation of such splitting as a consequence of
spin-orbit coupling [70] (see Chap. 9), the effect of which increases with the
number of particles and is much weaker for several electrons.
On the other hand, the graph in Fig. 5.8 shows the dependence of the
average number of electrons per dot on the applied voltage Vc. The number
of electrons was determined from the integrated relative difference in trans-
mission: N ex: Ji1TIT. Sharp steps in the curve correspond to the binding of
consecutive electrons in the dot, which occurs whenever the potential of the
electrode reaches the ground-state energy of a subsequent number of elec-
trons: N = 1, 2, 3, 4 (i.e., the Fermi level passes through the consecutive
discrete energy levels). According to the authors of the experiment, it is very
surprising that such a great number of dots (~ 108 ) can be simultaneously
58 5. Intraband Optical Transitions

.
125r---------------------------
.....~.
"./"
e"
100 • ...."''''...,..
...
4 e/dot
~

";"E 75 •....../ . / Gate


~ .........
8 50 ...•........ Vgrt"J~~~~i$j~ ES Fig. 5.7. FIR resonance ener-
........... gies of an array of quantum dots,
25 ...... each containing four electrons, as
f.'-...-.__._ &-Iayer a function of the magnetic field.
o --......-------.--.-.--- . ---.-- ....-. Inset: schematic of the sample; ES
2 3 456 789 shows the regions of electron lo-
Magnetic field (T) calization [105]

Ul
c
0
4
1:5 I , ! I
Q)
3
Q)

...
<0-
0
! 1 ! q3
1
Q) 2
..c
E
::l
Z Fig. 5.S. Dependence of the av-
erage number of electrons per dot,
-0.76 -0.74 -0.72 -0.70 obtained from the FIR spectra, on
Gate voltage (V) the gate voltage [105]

populated with equal numbers of electrons. The reason for this is the signif-
icant charging energy that is required to add a next electron to the system
(~ 15 meV, which exceeds the energy of thermal excitation kBT). The charg-
ing energy measures the Coulomb repulsion between the added electron and
the electrons that are initially confined in the dot.
6. Interband Optical Transitions

6.1 Idea of the Photoluminescence Experiment

The photoluminescence measurements are the basic tool that allow for the
investigation of discrete energy levels of quantum dots. For this reason this
is usually the first step in the studies of quantum dots obtained by vari-
ous methods (described previously in Chap. 2). The idea of the photolumi-
nescence experiment is presented in Fig. 6.l. A laser beam of appropriate
wavelength excites the electrons from the valence band (VB) to the conduc-
tion band (CB) and creates electron-hole pairs (rin ----+ e + h). These pairs
can be excited directly into the discrete levels in the quantum dot, or above
the discrete levels; to the two-dimensional quantum-well continuum or (even
higher excitation energy) to the bulk semiconductor continuum. A fraction of
the generated particles relax nonradiatively and fall to the ground state or to
weakly excited states in quantum dots. The electron-hole pairs confined in the
dots recombine, emitting photons, which are then registered (e + h ----+ ')'out).

-.0 I
increa sing excitation power.

CDI
;:1
.!!!I
<D
~I m
n.


§I
liiil a
::::J
00

CD ++ -- c:r Fig. 6.1. The photolumines-


~ :Ei rout ... Yout ...
Ol
::::J
a.
cence experiment: measurement
.!!! of the intensity of outgoing
co
~ ~

(laser)
...
"0


[II Ol beam rout as a function of the
"0
--0-0-- wavelength rout at the con-
:::r stant wavelength of the incom-
o
1 m
00
ing beam rin (PL spectrum), or
I on the wavelength rin at the
1 constant wavelength rout (PLE
~O spectrum)
60 6. Interband Optical Transitions

Two main types of photoluminescence experiments can be distinguished


that are used for measurements of the spectrum of emission and absorption
of light by the system, respectively.
In the first case (photoluminescence, PL) the wavelength of the light ex-
citing electrons is small, and the electron-hole pairs (excitons) are created
high above the discrete energy levels of the dots. The measured spectrum,
dependence of the radiation intensity lout on energy Eout for the outgoing
beam, provides information on the structure of energy levels of the quantum
dots. As shown in Fig. 6.1, depending On the power of the laser beam Pexc ,
directly related to the exciton creation rate dn/dt in the fixed surface of the
sample A,
1 dn
Pexc ex A dt' (6.1)

two situations are possible:


(i) The number of created excitons is so small that a single exciton at
most is confined at a time in each dot (each confined exciton recombines be-
fore the capture of the next one). When the relaxation rate is much higher
than the recombination rate (typical for the bulk material or the quantum
well), almost all excitons relax to the ground state before the recombination,
and a single peak dominates the PL spectrum. The reason for the occurrence
of higher-energy peaks in the PL spectrum may be, however, the appearance
of metastable states in the exciton energy spectrum, for which the character-
istic relaxation lifetime is at least comparable with the lifetime of radiative
recombination. A significant fraction of excitons that relax to the lowest dis-
crete states in the dots stay in the metastable excited states and never reach
the ground state. An additional higher-energy peak in the PL spectrum cor-
responds to the radiative recombination from a metastable state. An analysis
of this effect was presented by Jacak et al. in [68, 71] (see also Chap. 10).
(ii) The number of created pairs is large enough to reach the state of
quasi-equilibrium, in which each dot is (on average) populated with a num-
ber (N) of bound excitons. In such a case, the recombination from the initial
(ground) state of N excitons is possible either to the final ground state or to
final excited states of N -1 excitons. In consequence, higher-energy peaks ap-
pear in the PL spectrum, the positions of which correspond to the excitation
energies of the (final) system of N -1 excitons. In a noninteracting system (or
if, due to a strong spatial quantization, the many-particle states are approxi-
mately the noninteracting states) this corresponds to the situation presented
in Fig. 6.1: due to the Pauli exclusion principle, some of the particles occupy
the higher-energy levels, and the recombination of electron-hole pairs with
different total single-particle energies is possible.
In the second case (photoluminescence excitation, PLE) the energy of the
incoming (laser-beam) photon Ein is varied, and only those outgoing photons
lout are registered that have their energy equal to a fixed constant (below the
energy of the lowest discrete level). The measured spectrum, dependence of
6.2 Carrier Relaxation 61

the intensity of emitted radiation {'out on the energies of the incoming photons
E in , is related to the number of electron-hole pairs excited at a given energy
Ein. This is because the number of electron-hole pairs that relax to the state
from which the recombination is registered depends on the total number
of created pairs. On the other hand, the probability of creation of a pair
(absorption of a photon) is proportional to a joint (optical) density of states
of the electron-hole pair. Hence, a peak in the PLE spectrum corresponds to
a peak in the optical density of states, which in the case of a quantum dot
translates to a discrete energy level.
The condition of obtaining clear peaks in the photoluminescence spec-
trum, which are characteristic of systems with a strong spatial confinement
(with a discrete spectrum of energy levels), is the lack of thermal broadening.
Therefore, the temperature at which the measurements are usually taken is
on the order of a few K, so that it satisfies the condition kBT « E: (E: is the
separation of neighboring energy levels). However, in the case of very small
self-assembled dots (SAD) the energy quantization is so strong (E: ~ 45 meV)
that Raymond et al. managed to register the PL signal even at room tem-
perature [114].
Due to the small intensity of radiation that is emitted by a single dot,
the measurements are usually taken from a large system (array) of dots. The
inhomogeneity of the individual dot sizes leads to an additional broadening
of peaks.

6.2 Carrier Relaxation

Particularly in the case of the emission spectrum (PL), the possibility of


observing higher-energy peaks, which correspond to the recombination from
excited states of a quantum dot, depends on the relation between the rates of
a pair of competing processes: the nonradiative relaxation of carriers and the
radiative recombination of electron-hole pairs. These problems were discussed
in [9, 10, 11, 12]. The most significant mechanisms of the carrier relaxation
in semiconductor structures are the interaction with longitudinal phonons:
optical LO and acoustic LA (carrier scattering on the transverse phonons TO
and TA is elastic [16]), emission of an FIR photon, and the Auger processes.
The effective relaxation through the Auger processes, discussed extensively in
[10], requires the presence of a high-concentration gas of free (not confined)
electrons, which interact electrically with electrons confined in the quantum
dot. Below we present the results presented in [9, 11, 12] and connected with
the relaxation through the interaction with phonons.
In the first order of the perturbation, the rate of relaxation through the
emission of a phonon is given by the golden Fermi rule:

(6.2)
62 6. Interband Optical Transitions

where Ii) and If) stand for the initial and final states of the exciton, respec-
tively, q is the wave vector of the emitted phonon; mu q = /iqc s is the phonon
energy, Cs is the velocity of sound (3700 mjs for GaAs) , and nB represents
the Bose-Einstein distribution of phonons in the crystal lattice at tempera-
ture T (at zero temperature we have nB = 0). The electron-phonon coupling
operator W has the form
(6.3)
The relaxation of an exciton, which is confined in a quantum dot, through
the emission of an optical phonon LO is rather inefficient. The phonon energy
ELO is almost constant (weakly dependent on the wave vector), and there-
fore the first-order process (with the emission of a single phonon) requires a
precise fitting of the distance between the two discrete energy levels of the
confined exciton, and the photon energy ELO. The interaction of the exciton
with LO phonons is obviously also possible through higher-order processes
(e.g., through a simultaneous emission ofLO and LA phonons). However the
efficiency of such combined processes is very limited. Moreover, energy ELo
is fairly high (~ 30 meV for GaAs) and excludes the relaxation between a
pair of closer excitonic energy levels.
Considering an acoustic phonon LA, in the simplest case of the electron-
phonon interaction, the parameter a in (6.3) is expressed by the deformation
potential D (8.6 meV for GaAs):
2 D2/iq
a (q) = -2-'
{lC
(6.4)
s

where {l is the density ofthe crystal (5.3 gjcm 3 for GaAs). A similar equation
describes the parameter a which governs the hole-phonon interaction (the
only difference is that due to the anisotropy of the hole effective mass, a
strongly depends on the direction of the phonon wave vector q). Figure 6.2
presents a comparison of the rates of the radiative recombination and of the
relaxation from the first excited state from which the recombination is allowed
by the selection rules (zero total orbital angular momentum of the pair), as a
function of the characteristic frequency Wo of the parabolic confining potential
of the dot ~m*w5r2.
For small phonon wave vectors lengths q the probability of relaxation
increases with q, and hence with the phonon energy, equal to the distance
between the ground and excited states of the exciton (~ muo). The reason
for such behavior is the relation W IX yfq (see also (6.3) and (6.4)).
However, if the phonon wavelength drops below the smallest characteristic
dot dimension (Le., below the thickness of a quantum well, from which the
dot was created), the matrix element I (fleiqrli) I, which appears in (6.3),
decays rapidly (integral of a quickly oscillating function) independently of
the direction of the emitted phonon (direction q). In effect, the relaxation
rate itself rapidly decreases with q. This means that the exciton confined in
6.2 Carrier Relaxation 63

80~-----------------------,

FIRST EXCITED
j=O EXCITON
60

Fig. 6.2. Relaxation rate (solid


line) and radiative recombination

--
20 rate (dashed line) of an exciton in
a quantum dot, from the first op-
RADIATIVE DECAY tically active excited state (with
zero angular momentum), as a
O~--~----~---r~--r---,
o 2 4 6 8 10 function of the characteristic fre-
quency of the confining potential
[11]

a quantum dot of diameter L is not able to emit a phonon with wavelength


A < L, i.e., with an energy Iiw q > E, where the critical value is
E _ 27rn
(6.5)
- Cs L .
For instance, for a quantum dot made in a 3 nm thick GaAs quantum well,
it gives E = 5.1 meV. As shown in Fig. 6.2, at roughly the same energy the
probability of nonradiative relaxation drops below the probability of radia-
tive recombination. As a result, in small dots there is no efficient mechanism
of relaxation through the interaction with phonons. Such an explanation was
proposed in order to explain the structure of luminescence peaks (appearance
and then vanishing of the additional peak) shown in Fig. 6.3. This interpre-
tation, however, does not seem complete, as it overlooks the possibility of a
radiative transition between different excitonic states.
A remaining alternative mechanism of recombination is the emission of a
low-energy photon (in the far infrared range). The emission of FIR photons
may lead to the effective freeing of excited excitonic states in a quantum dot.
However, for the states with an appropriate symmetry, these transitions may
be forbidden by the angular-momentum selection rules. This happens when
both the electron and the hole have equal (zero) angular momenta (are in the
s-type states) both in the ground state and in the excited state of the exciton
(electron and hole angular momenta are not the good quantum numbers sep-
arately. However, the eigenstate of the electron-hole pair may be very close
to a noncorrelated state [89, 90]). Since the emitted photon carries the angu-
lar momentum (see Sect. 3.4), the transition between the electron and hole
s-type states is forbidden and in such a case the excited state of an exciton
is metastable. Hence, when the channel of relaxation through the interaction
with phonons is blocked, the photoluminescence spectrum is determined by
the stability of the excited states against the emission of an FIR photon. As
64 6. Interband Optical Transitions

Inlerdiffused Region
[J 25nm 10.6s
Dol Size
w=
1000 A Si3N4
100A GaAs Onm
200 A AIGaAs (Inlerdiffused
30A GaAs Region)
200 A

~
(x=0.35) Q tu 0 t

---------=
uan m CB

L ~
0

E
I I
~
C
x.y
Dot Sizew

.
::I
.ri
.!!.
~
enz x50 400nm

w
I-
Z
W
U 450nm
Z
W
u
en
w 500nm
z
:2
::::>
...J
e
l-
1000nm

e Fig. 6.3. Photolumines-


:J: cence spectra of single quan-
Cl. QW(Ref.)
x1 tum dots; numbers on the
right-hand side of the curves
1.68 1 .70 1. 72 1. 74 1.76 denote the dot diameters
ENERGY (meV) [20]

shown by Jacak et al. [68, 71] (see also Chap. 10), the investigation of the
stability of these states allows for an interpretation of the above-described
experiment of Bockelmann et al. (and also other results of luminescence mea-
surements for dots of different types). According to this interpretation, the
additional peak in the photoluminescence spectrum corresponds to a recom-
bination from a weakly excited metastable state of an exciton. It is shown
that the distance between the pair of peaks and also the height and broaden-
ing of the higher-energy peak, depend strongly on the dot size. The results of
the calculation, presented briefly in Chap. 10, show that the evolution of the
photoluminescence spectrum corresponds very well to the experimental de-
pendence in Fig. 6.3. For a large dot the spectrum is dominated by a strong
peak, which corresponds to the recombination from the ground state. The
decrease of the dot size is initially accompanied by the lowering of the excita-
tion energy of a metastable state, and in turn by a visible strengthening and
6.3 Observation of Magic States of Quantum Dots in the Absorption Spectrum 65

narrowing of the additional peak. At a certain critical dot size this additional
peak becomes comparable with the main peak. Above the critical size the
repeated increase of the main peak is observed, as well as a decrease of the
additional peak. At the same time the displacement of peaks toward higher
energies, which accompanies the decrease of the dot size, is observed, similar
to that observed in the experiment (see Fig. 6.3).

6.3 Observation of Magic States of Quantum Dots


in the Absorption Spectrum
The interband absorption spectrum (in the range of visible light) may allow
for a direct observation of transitions between the magic states (in the few-
electron quantum dots), or the breakup of the compact droplet at a filling
factor v = 1 (in larger systems) [58, 60, 61]. However, no such experiment
has been successfully carried out so far. Figure 6.4 presents the single-particle
states of a quantum dot in a strong magnetic field. CB and VB denote the
conduction and valence bands, respectively. The forbidden energy gap be-
tween the two bands is disregarded. Restricting ourselves to the lowest Lan-
dau level, let us label the orbitals with a single quantum number M, equal to
the orbital angular momentum. Prior to the photon absorption, the valence

>-
~ Fermi level
- cb

(j)
c
ill

o~~~~~~~~~W+~~~~~
- - -______ vb

o 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Fermi level -
-
>- Fig. 6.4. Schematic dia-
....
Ol
holes
gram of the single-particle
(j)
c energy levels of a quan-
ill tum dot in a strong mag-
netic field; vertical arrows
- allowed interband opti-
cal transitions. Upper di-
agram: compact droplet;
o 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 lower diagram: recon-
M structed droplet [60, 61]
66 6. Interband Optical Transitions

band is completely filled, and the occupied electronic states in the conduction
band (states occupied by the confined electrons) are marked with rectangles.
Some of allowed optical transitions of an electron from the valence band to
the conduction band are marked with vertical arrows: the orbital angular
momentum of the excited electron does not change, and the orbital to which
the electron is promoted is empty (the Pauli exclusion principle).
The upper and lower diagrams show two possible situations. When the
electrons form a compact droplet (upper diagram), the excitation of an elec-
tron is possible only above the Fermi level, i.e., outside the area occupied
by the dot in the real space (outside the droplet). The kinetic energy of this
excitation is high, and the attraction between the hole and the electron sys-
tem is small. Hence, the energy of creation of such an exciton is high. When
the electrons are in the state with a lower filling factor (a magic state with
a higher total angular momentum for several electrons, or a many-electron
system with a reconstructed edge), it is possible to excite an electron below
the Fermi level, inside the physical area of the electron droplet. The energy
of the photon required to promote a valence-band electron inside the droplet
is lower than that required for the excitation outside the droplet. Thus, it
should be expected that the transition in the ground state of the quantum
dot, induced by a magnetic field, will result in the occurrence of a new strong
low-energy peak in the absorption spectrum (PLE). This new peak would
emerge below the main peak, which corresponds to the excitation near the
edge of the droplet. Obviously, the actual electronic droplet is an interact-
ing system, while what is shown in Fig. 6.4 disregards electron correlations.
However, as shown in the realistic calculations for the interacting system,
the conclusion, which suggested the appearance of a new peak indicating a
change of the ground state of a dot, remains true. It is assumed here that all
electrons are spin polarized (the calculations show that three electrons form
a polarized system, also a polarized compact droplet, over a wide range of a
magic field [54]), while the spin of the excited fourth electron depends on the
polarization of light.
The exact absorption spectra, determined for a parabolic dot contain-
ing three electrons, are presented in Fig. 6.5. The upper diagram shows the
spectrum in a magnetic field of 2 T, when the ground state corresponds to
v = 1 and angular momentum L = 3. The strong peak with the lowest en-
ergy indicates the creation of an exciton at the droplet edge. For a larger
number of electrons this corresponds to the Fermi edge singularity [53]. The
lower diagram presents a spectrum in a magnetic field B = 4 T, for which
the electrons form a ring with a higher magic angular momentum L = 6. A
new peak corresponding to the creation of an exciton inside the bulk of the
droplet is visible. The evolution of the absorption spectra of three electrons
in an increasing magnetic field is presented in Fig. 6.6. The circles represent
absorption peaks, and their areas are proportional to the individual transition
intensities. Since the circularly polarized light excites only electrons with a
6.3 Observation of Magic States of Quantum Dots in the Absorption Spectrum 67

4 compact dot
B=2Tesla, R=3
c 3
o
e-O
'.0:;

In 2
..0
«

04-------~--~~~~~~~~~

4 reconstructed dot
B=4Tesla, R=6
3 Fig. 6.5. Absorption
c
o spectra of a quantum
e-
'.0:;
dot containing three el-
O ectrons. Upper diagram:
2
In
..0
compact state with L =
« 3 (droplet), lower dia-
gram: reconstructed
state with L = 6 (ring) .
Insets: approximate elec-
tron density dis-
04-~~~~~~~~~~~~--~ tributions Insets: appro-
-2.0 -1.5 -0.5 0.0 ximate electron density
distributions [61]

defined spin projection, two spin configurations of four final-state conduction


electrons are possible: The projection of the total spin along the field direc-
tion can be Sz = 2 (an excited electron has the same spin as the polarized
system of three electrons in the initial state), upper diagram; or Sz = 1 (an
excited electron has the opposite spin), lower diagram.
For Sz = 2, at magnetic fields below 2.5 T, the peak marked as a for
B = 2 T dominates, which corresponds to the creation of an exciton at the
droplet edge. At a field of 2.5 T a change of the ground state of three electrons
to the next magic state occurs, and this change is reflected in the spectrum
by the appearance of a new peak, marked for B = 2.5 T as b. A subsequent
change of the ground state of the dot occurs at B = 4 T. In addition to the
occurrence of a new peak b, which signals the first transition, the oscillation
of the main peak (a) is observed at each subsequent transition.
For Sz = 1 the Pauli principle does not exclude the creation of an exciton
inside a droplet even in weak magnetic fields (peak c for B = 2 T), since the
promoted electron can be distinguished from the three initial-state electrons
68 6. Interband Optical 'Transitions

-too --..-------T"""O.....-::.:-----.-""T"-,.-,'.i"'"-rr.
• • ! ; : :.
-rT."'!,,~"T'
• !
',' -,:r,.-,

••• ii,

!
i ~

, O
, , 'f· t
: t.
-1.25 : :: :

e a • • ~ •.•.•• !~,~~~~
ee@.
......
0
-1.50
. : :. .'
WO
:;;::~:
:::::::
::i.
•••••••
I
~ -t75
• •
• • •
-2.00

R=3 R=6

-1.25

WO -1.50
:::::::

r R~:
::i.
I
~ -1.75

-2.00 b) 1
R=3 R=6
23456
magnetic field, Tesla
Fig. 6.6a,b. Magnetic-field evolution of the absorption spectra of three electrons
confined in a quantum dot. Areas of circles correspond to intensities of individual
transitions. The two diagrams correspond to the two circular polarizations of light
(spin projection of four electrons in the final state): (a) 0"+ (Sz = 2) and (b) 0"_
(Sz = 1) [61]

by its opposite spin projection. As a result, the changes of the ground state do
not cause such dramatic changes in the absorption spectrum as was observed
for Sz = 2. The striking difference between the spectra for the two polar-
izations of light at low magnetic fields, when the three initial-state electrons
form a compact droplet, should be experimentally observable.
6.4 Interaction of the Exciton with an Additional Electron 69

6.4 Interaction of the Exciton


with an Additional Electron

A similar difference between the absorption spectra, which are measured at


two different polarizations of light, occurs for any number of spin polarized
electrons. In [132]' the spectra have been determined for a single confined
electron in a strong magnetic field, presented in Fig. 6.7 (calculated in the
band-structure model for the material of the quantum dot). The addition
of an exciton X into the system leads to the creation of the X- complex
(charged exciton), in which two light particles (electrons) are bound to one
heavy particle (hole). Two light polarizations, u+ and u_, correspond to the
polarized (Sz = 1) and unpolarized (Sz = 0) electrons, respectively. The
electron-exciton binding energy is positive (stable X- complex even in the
absence of the external confining potential) only in the state Sz = o. In the
case of polarized electrons (Sz = 0), in a strong magnetic field, a magneto-
exciton in the ground state does not interact with the additional electron
(binding energy equals zero). Consequently, the energy of the first absorp-
tion peak for X- is either smaller or equal to the energy of the first peak
for a neutral exciton X. Similar spectra were measured in a quantum well
(B < 11 T), where the excitons were probably localized on the fluctuations

10
8=10T X

-
::::I
~ 10
>- X Ct

-
:t::
en
c:
Q)

c: 5
c:
0
;:;
...0
Co
10 -
en X a+
.Q Fig. 6.7. Absorption spectra of an
CIS
empty dot (X is created) and of a
5 dot containing one electron (X- is
created) in a magnetic field of 10 T.
The two spectra for X - correspond to
two circular polarizations of light 0"+
Or-----~~~4-~~-L~~ and 0"_ (two spin-polarized and spin-
-15 -10 -5 o unpolarized electrons in the final state)
energy (meV) [132]
70 6. Interband Optical Transitions

of quantum-well thickness [82, 83]. The difference between the 0'+ and 0'_
spectra allowed for the identification of the low-energy peak as that corre-
sponding to the creation of X- . It should be added that in the magnetic field
of 10 T a weak confining potential (in [132] assumed as Wo = 2.5 meV) has
only a small effect on the obtained X- binding energy. Hence, the spectra
calculated in [132] describe both quantum dots and quantum wells.
The above-mentioned lack of interaction between an exciton and an elec-
tron in a strong magnetic field results from a more general property. This
involves a two-dimensional system of electrons and holes in a strong mag-
netic field, when all particles are spin polarized and the wave functions of
electrons and holes are identical (in spite of different effective masses). Such
a system has the eigenstate P, in which the excitons interact neither among
themselves nor with the fluid of excess carriers (e.g., electrons). Hence, the
excitons form the phase that is, in a sense, analogous to the Bose-Einstein
condensate [24, 96, 99, 108]. The state P is the state with the lowest en-
ergy in a subspace with total orbital angular momentum equal to that of the
ground state of excess electrons. This subspace is reached when the excitons
are created in an electron system that was initially in its ground state.
The vanishing of the interaction energy for two electrons and one hole
means that the electron-electron exchange energy, which attracts the photo-
created exciton to the initial-state electron, compensates exactly for the weak-
ening of the attraction within this created electron-hole pair, which has to be
created outside the orbital initially occupied by the first electron. As a result
the energy of the final-state X- complex is equal to the sum of electron and
exciton ground-state energies.

6.5 Measurements of Photoluminescence


The measured emission PL spectra (and the corresponding absorption PLE
spectra) of quantum dots are characterized by considerably greater intensity
per unit volume than the spectra of two-dimensional quantum wells. For
instance, Kash et al. [79] reported a 5-fold difference in intensity between the
photoluminescence peaks from dots and wells, corresponding to the ground-
state emission (exciton recombining from its lowest-energy state), and as
great as 50- to 100-fold difference for the higher-energy peaks. The occurrence
of the (additional) lateral potential confining particles in the quantum dot
results in the maximum in the photoluminescence spectrum of quantum dots
shifting slightly toward higher energies, compared to the maximum in the
spectrum of the quantum well from which the dots were made. An,illustration
of this effect is presented in Fig. 6.8 taken from Temkin et al. [128]. It shows
the photoluminescence spectra measured for a quantum well, quantum dot,
and quantum wire. The occurrence of sharp peaks in the photoluminescence
spectrum of quantum dots is characteristic of systems with a completely
discrete energy spectrum, which then distinguishes a zero-dimensional system
6.5 Measurements of Photoluminescence 71

T=6K 8E=(14.,!) meV


1= 6328A --I I--
100A,lnP
=====50A InGaAs
InP Buffer
InP Substrate

QUANTUM WELL Fig. 6.S. Photolumines-


cence spectra of quan-
tum wells, wires and dots.
Configuration of layers in
the quantum well, from
which the dots and wires
were made, and the di-
1.37 1.42 mensions of structures are
WAVELENGTH (J.un) given [128]

(a) Dot spectra (b) Line spectra

X5
~
(/)
c:
2c:

Fig. 6.9. Photoluminescence spectra of: (a) quantum dots with diameters (ii) d =
140 nm, (iii) d = no nm, (iv) d = 80 nm; (b) quantum wires with widths (ii)
d = 160 nm, (iii) d = 140 nm, (iv) d = 100 nm. In both diagrams, (i) is the
spectrum of the quantum well from which the dots and wires were made [95]

from other cases. A direct reason for the appearance of a large number of small
peaks, rather than a few strong and clearly separated ones at the energies that
correspond to the discrete energy levels, may be the inhomogeneity of sizes
of individual quantum dots. Figure 6.9 shows the graphs of Lebens et al. [95],
which present the changes in the photoluminescence spectra of quantum dots
and wires, which appear in decreasing sizes. The displacement of the main
maximum toward higher energies is visible. Also, the gradual emergence of
small peaks becomes more and more visible with an increase in the separation
between energy levels (compared to the widths of the peaks).
72 6. Interband Optical Transitions

Zrenner et al. [139] describe an experiment in which the photolumines-


cence spectrum allows for the identification of a natural formation of quan-
tum dots as a result of the inhomogeneous thickness of a thin GaAs quantum
well. For a quantum well with a thickness of 10-12 monolayers, the fluctua-
tions of the overall potential, which appear due to the local changes of the
well thickness by 1-2 atomic layers, were estimated to about 36 meV for
the conduction-band electrons and about 7 meV for the valence-band holes
[140] (this estimation assumed the band-structure approximation within the
dot area). These fluctuations are thus strong enough for the localization of
excitons to be observed.
The quantum well used in the experiment was made of adjacent thin
layers of GaAs (3 nm) and AlAs (4 nm), placed between 40 nm thick bar-
riers of Alo.4sGao.52As. Such a structure is called a coupled quantum well
(CQW). AlAs is a compound with an indirect band gap, with minimum en-
ergy in the conduction band at point X. A respectively strong electric field,
which is produced by the applied voltage and directed from the GaAs layer
toward the AlAs layer, lowers the X-point energy in the AlAs layer below
the r-point energy in the GaAs layer. As a result, the exciton created op-
tically at point r (the absorbed photon has a very small momentum, and
the excitation between the valence and conduction bands occurs with almost
conserved momentum, i.e., at point r - a direct process) very quickly relaxes
nonradiatively to the lowest-energy state. In this state the electron and the
hole are separated both in the momentum space (X-r) and in the real space
(AIAs-GaAs). Due to the electron-hole separation, the lifetime of such an
indirect exciton is enlarged by three orders of magnitude (500 ns compared
to 0.5 ns for a direct exciton). Due to its long lifetime, a mobile exciton
can, prior to the recombination, flow into a region where an islet with locally
increased thickness of the GaAs layer is formed (fluctuations on the order
of two monolayers). Then it can relax to the state in which both particles
(electron and hole) are again in the same (GaAs) layer. By controlling the
energy of point X in the AlAs layer, compared to the energy of point r in
the GaAs layer, over a wide range (about 70 meV), it is possible to realize
a resonant transition of excitons from the separated state (X-r) directly to
one of the discrete states in a quantum dot (the differences between the en-
ergy and the quasi-momentum x-r have to be equal to the energy and the
quasi-momentum of the appropriate phonon). Thus, the point X in the AlAs
layer serves as an internal reservoir of excitons.
Figure 6.10 presents the photoluminescence spectra from this paper. The
spectrum (a) was measured by collecting the signal from the area with di-
ameter of 100 Mm. Two maxima can be distinguished: a stronger 0ne, which
corresponds to the indirect recombination from the lowest-energy state (at
point X), and a much weaker one, which corresponds to the direct recombi-
nation at point r. On the other hand, in spectrum (b), measured for a much
smaller area of the sample (about 2 Mm in diameter), several sharp peaks
6.5 Measurements of Photoluminescence 73

>-
~
~ 1700
W • PL-Experiment
....:. - Calculation
Il. 1660L-_........:G::::aAs/.'-:::::.iI\I.::./>s~28::.::.8:.o:AJ:::.:50::..:.JA
-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5

GaAslAlAs 30A/50A E~- HHo Fig. 6.10. Photolumi-


II Vs=-O.2V j nescence spectra of a
c T=4.8 K GaAs/ AlAs coupled quan-

o
::I
o "L =632.8 nm tum well, collected from
the areas of diameters: (a)
....I E~- HHo dL = lOOp,m and (b) dL =
Il.
j (a) 2p,m. Sharp peaks in (b)
correspond to the recom-
bination of excitons local-
ized on the fluctuations
of the quantum-well thick-
ness. Insets: configuration
of band edges, and the de-
(b) pendence of the conduc-
tion-band energy in points
1700 1710 1750 1760 r and X on voltage VB
[139]

marked with arrows are clearly visible. These peaks are characteristic of the
recombination of an exciton from discrete energy levels, i.e., the localization
of an exciton in a small area.
Zrenner et al. [140] present a more recent experiment carried out by the
same group. They report a successful observation of photoluminescence peaks
that correspond to the optical transitions from the system of one and two
excitons confined in a single quantum dot in a perpendicular magnetic field.
The sample used in the experiment (coupled quantum well GaAs/AIAs) and
the relaxation of carriers to the quantum dot have been described above
(Fig. 6.10 and the description in the text). In order to obtain a signal from
a single dot, several conditions had to be fulfilled. The signal was collected
from an area limited to about 1 p,m 2 by the application of a high-resolution
detecting system (space-resolved photoluminescence). Such an area contains
about 100-1000 quantum dots, and in order to select only a single dot (the
one in which the exciton energy is the lowest), the voltage drop VB, which
was applied across the well, was varied. The excitons from the state with the
hole at point X (AlAs) can relax only to the dots in GaAs with energies below
the critical value, which depends on VB. By lowering VB to the appropriate
value, it was possible to obtain a situation in which the ground state of only
one dot is occupied in the sampled area. The enlargement of the exciton
lifetime obtained in the coupled-well structure is necessary in this situation
74 6. Interband Optical Transitions

GaAs/AIAs 30Al40A
AL=632.8nm
PL=20J.1W
T=8K

i~------~~~~~
r:::
::J I-----..I..--.=..~~Ir"""'"

£I---------~~~~V~
ro
~I-------~~~~

~ 1-----.-::::"-xtI-IrMd
::J
O~--~~~~~~~~--=---------~
()
~r---~~~~Hft-r--~--------;

~~---"~--~H*--=--------~

Fig. 6.11. Photolumines-


cence spectrum of quantum
dots formed as the fluctu-
ations of thickness of the
GaAs/ AlAs quantum well,
DO-XX as a function of the volt-
age VB. Arrows - peaks cor-
responding to the ground-
DD-X Vs=-D.5
state recombination of one
1710 1715 1720 1725 1730 1735 1740 1745 1750 1755 (OD-X) and two (OD-XX)
Energy (meV) excitons [140]

to allow the exciton to relax from the AlAs layer to the selected quantum dot
before recombination. Figure 6.11 presents the measured dependence of the
photoluminescence spectrum on voltage VB. Lowering of the X-point energy
in GaAs below the r-point energy takes place at VB = 0.35 meV. Starting
from that value, a shift of the main peak toward lower energies (the Stark
shift) is observed. For the values of VB close to zero, the filling of many
dots with various sizes becomes possible, and the spectrum is composed of
many separate peaks. Below VB = -0.35 a single peak is observed, which
corresponds to the emission of a photon from the single lowest-energy dot
containing one exciton.
Analyzing the spectrum a little below VB = -0.35, the authors observed a
peak emerging for VB > -0.35 at energy about 1720 meV and identified it as
the emission from the system of two excitons. Probably there are in fact two
very close peaks, corresponding to two spin configurations: a singlet (vanish
6.5 Measurements of Photoluminescence 75

both the total spin of the pair of electrons, ISel = IShl = 0, and the total
spin of the pair of holes, Se,z = Sh,z = 0) and a triplet (ISel = IShl = 1 and
Se,z = -Sh,z = 0, ±1), split by the electron-electron and hole-hole exchange
interactions.
Zrenner et al. [140] studied also the magnetic-field evolution of energies
of the photoluminescence peaks for the emission from a system of one or
two excitons in the dot (peaks are marked with arrows in Fig. 6.11). The
measurements were carried out in the magnetic fields 0-12 T. Each of the
three peaks (a single exciton and a pair of excitons in a singlet and triplet
spin configuration) in a nonzero magnetic field splits into two peaks, which
correspond to the opposite spin projections of all particles onto the field
direction. The magnitude of the splitting is the Zeeman energy, equal to
81.7 /LeV IT.
Brunner et al. [21] also describe the measurement of a photoluminescence
signal from a system of two excitons confined in a quantum dot. Similarly,
as in the work described above, the investigated dots were formed in the
fluctuations of the quantum-well thickness. The quantum well consisted of
a 3.4 nm thick GaAs layer placed between the Alo.35Gao.65As barriers. The
measurement was carried out using two values of power of the laser beam,
which created excitons: 0.1 /LW and 5 /LW. The area from which the photolu-
minescence signal was collected had diameter about 1 /Lm. One of the original
graphs is shown in Fig. 6.12. In the photoluminescence spectra measured at
a small laser power, a strong wide peak with energy 1671 meV can be iden-
tified, which corresponds to the recombination of two-dimensional excitons,
which move freely in a quantum well. Also, at energies of 1659-1671 meV, a
number of weaker sharp peaks appear. Similarly as in [139, 140], the sharp
peaks with energies below the free-exciton recombination energy (1671 meV)
were identified as the signals of the recombination of excitons localized in
areas of locally increased thickness of the quantum well (for the localization
of the exciton in the dot; see the discussion in Sect. 3.5.1). On the other
hand, in the graphs corresponding to a higher laser power, additional very
weak peaks appear at energies 1653-1655 meV. These energies lie below the
lowest energies of single localized excitons (1659 meV). When a higher ex-
citation (laser) power is applied, the number of created excitons in a given
area increases. Hence, the probability of capturing a subsequent exciton in
the quantum dot before the recombination of the first one occurs also in-
creases. Thus, new peaks were connected with recombination processes from
the systems of two excitons, which were confined in the quantum dot. (A
comment might be added here that the band-structure model of the quan-
tum dot does not allow one to analyze the metastable states of excitons, the
occurrence of which explains the additional luminescence peaks in a natural
way; see Chap. 10.) The difference between the ground-state energies of a
single exciton and of a pair of excitons was read from the photoluminescence
76 6. Interband Optical Transitions

; [ r----G-aA-sl-A-lo.-3s-G-a-o.o-sA-S-O-W---------t

§ (Lz =34A withGI)

.e~ OD-XX oD-X


Pexc '" ,
~ 511W
(J)
Z
W O.1I1W
I-
Z
w
()
Z
W
()
(J)
W Fig. 6.12. Photolumines-
Z cence spectra of a quan-
~ tum-dot array, measured at
:::> three different positions of
....J
o
I-
the detector. Labels indi-
cate peaks corresponding to
o
I
the recombination from the
~ L-~ __~__~__~____~__~__~~ quantum well (2D-X) and
from the quantum dot con-
1650 1655 1660 1665 1670 1675 1680
taining one (OD-X) and two
ENERGY (meV) (OD-XX) excitons [21]

spectrum as the distance between the appropriate peaks. This value allowed
to determine the dot diameter, on the order of a few tens of nanometers.
Figure 6.13 presents the photoluminescence spectra measured by Fafard
et al. [40] for the AlInAs self-assembled quantum dots (SAD) formed on the
AlGaAs substrate. The upper frame (a) shows the spectrum measured in a
sample containing about 106 dots. A single wide Gaussian peak is observed,
which corresponds to the recombination from a great number of dots of differ-
ent sizes. The broadening of the peak position reflects, to a certain degree, the
inhomogeneous broadening of the dot dimensions in the sample (the broad-
ening is also related to the exciton recombination time). The lower frame (b)
presents the spectra that were registered for smaller and smaller numbers of
dots: about 2500, 1400, and 600, which correspond to the sampled areas of
13, 7, and 3 J-Lm 2 , respectively. It is clearly visible that as the number of dots
in the sample decreases, the large Gaussian peak is gradually replaced by
a group of single small peaks, which correspond to the emission of photons
from particular dots. The width of an elementary peaks is ~ 0.4 meV.
Bayer et al. [7] present the results of photoluminescence measurements
in strong magnetic fields. The change of dot diameters (34-41 nm) and the
change of magnetic field in the range 0-12 T enabled the observation of the
influence of both factors on the structure of luminescence peaks. As can be
6.5 Measurements of Photoluminescence 77

3500r--.,........-r---.---.--------.---.---,---,
-2500 dots
3000 (b) T=2K
W
='=2500
§ 1 I Fig. 6.13. Photolumines-
..c:i 2000 Probing -1 31lm2
I cence spectra of self-as-
~1500
(1.1000
I \ sembled quantum dots with
average diameter :::::: 17 nm:
(a) large number of dots
500 (:::::: 10 6 ), (b) small numbers
of dots (600-2500). Zero on
-40 -20 0 20 40 60 the energy axis represents
Energy from the peak (meV) wavelength>. = 660 nm [40]

noted, in addition to the main peak, which dominates in strong fields and for
small dots (see Fig. 6.14), additional peaks appear at weaker fields and for
greater dot sizes. The number of peaks varies between one and three, which
agrees with our theory presented in Chap. 10, which predicts the radiative
recombination from the metastable excited states. Different values of the
excitation (laser) power were used in the experiment (15-15000 W/cm 2 ),
and at the higher values the creation of many-exciton complexes should be
expected.
The matrices of quantum dots were made of a 5 nm thick InGaAs/GaAs
quantum well. Outside the selected areas, the covering layer of GaAs was
etched away, and the following structure was obtained: a thin layer of InGaAs
(quantum well) created on a GaAs layer (barrier) and covered with GaAs
only in the small selected areas. The potential barrier at the GaAs/lnGaAs
interface is much lower than that at the GaAs/vacuum interface (the electron
affinity for GaAs IJ :::::: 4 meV, while the step in the conduction-band edge at
the GaAs/lno.13Gao.87As interface is LlEcB :::::: 0.1 eV). Thus, the carriers are
confined in the areas covered with the GaAs layer. Three different samples
contained dots with diameters 34, 35, and 41 nm.
The strain present in the narrow InGaAs layer created on GaAs (lattice
constants differing by :::::: 1%) causes the splitting of heavy-hole and light-
hole subbands in the valence band by about 60 meV. Hence, the interaction
between the subbands is negligible, and for the description of holes a single-
band approximation can be used. Due to the small sizes of quantum dots, the
78 6. Interband Optical Transitions

(a) (c)
>-
:!::::
en
C
SC
"C
.~

ca~
E
g 34nm 4T
OT
1.43 1.45 1.43 1.45 1.43 1.45
energy (eV)
Fig. 6.14. Photoluminescence spectra of quantum dots in a magnetic field, mea-
sured for different dot sizes: (a) 34 nm, (b) 35 nm, (c) 41 nm [7]

separation between the single-particle energy levels in the quantum dot (up
to 15 meV) is on the same order of magnitude as the energy of the Coulomb
interaction, and the eigenstates of the interacting system are close to those
of the noninteracting system. From the numerical calculations (in the band-
structure model of a quantum dot; see the discussion in Sect. 3.5.1) it was
obtained that in the quantum well of which the dots were made, there is only
one confined subband for electrons and one subband for holes (neglecting
the spin). In a zero magnetic field the quantum dots have three confined
single-particle states (both for electrons and for holes): the s-type state with
angular momentum L = 0 and a pair of degenerate p-type states with angular
momenta L = ±l. In higher magnetic fields the d-type state becomes bound,
having angular momentum L = 2 for an electron and L = - 2 for a hole.
Figure 6.14 presents the photoluminescence spectra measured for three
different dot sizes at different values of the magnetic field. Higher-energy
peaks are observed (overlapping neighboring peaks can be resolved as a
change of the asymmetry of the envelope), as well as the rearrangement of
the spectrum in the increasing field. Figure 6.15 presents the energies of the
photoluminescence peaks, which are taken from the spectra similar to that
in Fig. 6.14, as a function of a magnetic field. Comparing the experimental
data (points) presented in Fig. 6.15 with the curves determined numerically
(diagonalization of the Hamiltonian of an electron-hole pair for the realistic
three-dimensional confining potential in the band-structure model of a quan-
tum dot), Bayer et al. identified the peaks marked in the figures with squares
as recombination from the states with zero angular momenta of both parti-
cles Le = Lh = 0 (s-type states); with diamonds, from the states Le = -1
and Lh = +1; with triangles, from the states Le = +1 and Lh = -1; and
with circles, from the states Le = -2 and Lh = +2.
6.5 Measurements of Photoluminescence 79

34nm DOTS 41nm DOTS


1.46

(a)
1.43 ~O-~2-~4-~6-~a'-!O""---5---1-0---LO
--'2:---4--6---la

magnetic field (T)


Fig. 6.15a-c. Magnetic-field evolution of the photoluminescence energies for dif-
ferent sizes of quantum dots [7]

Obviously, the Coulomb interaction between the electron and the hole that
form an exciton, as well as (neglected in the calculations) the Coulomb inter-
action with other confined electron-hole pairs, makes the above description
even more complicated. The states with defined values of electron and hole
angular momenta are not the eigenstates of the total Hamiltonian, which
conserves only the total angular momentum of the pair. The approximate
agreement of the results of measurements and calculations, and hence the
possibility of a simple interpretation of observed photoluminescence peaks in
terms of transitions between the appropriate single-particle states according
to the appropriate selection rules (zero total spin and angular momentum of
the pair), can be explained as the effect of the small dot dimensions. As men-
tioned above, the quantization of motion in such a small system results, first of
all, from spatial confinement (and a magnetic field, when the magnetic length
is comparable to the dot diameter), whereas the Coulomb interaction does
not substantially change the wave functions. Thus, the spectra in Fig. 6.15
slightly resemble the spectrum of Fock-Darwin single-particle energy levels
in Fig. 3.2 (only the energy of electron excitation fiwo is now replaced by the
sum of energies of electron and hole excitations, and the entire spectrum is
vertically shifted by the energy of electron-hole attraction). The influence of
the Coulomb interaction on the photoluminescence spectrum in such small
dots was discussed in detail in [134], and a summary of those results is given
in Sect. 8.3. The interaction between a number of excitons confined together
in a quantum dot is described in [135, 136] and in Sect. 8.4.
80 6. Interband Optical Transitions

A different interpretation ofthe results ofthis experiment can be obtained


by employing the approximation of a quantum dot as a local perturbation
of the crystal field, within the effective-mass formalism (see Chap. 10 and
[68, 71]). The appearance of additional peaks in the photoluminescence spec-
trum is then related to the existence of excited metastable states of the
exciton in a dot. The occurrence of these states results from the possibility
of blocking the relaxation through the FIR photon emission between states
with the same symmetry, while the interaction of the system with phonons
is negligibly inefficient (due to the low temperature and small dot sizes; see
Sect. 6.2). As shown in the detailed analysis (see Chapt. 10) the form of the
photoluminescence spectrum in a zero magnetic field strongly depends on the
dot size. For the sizes close to the critical value (depending, for example, on
the material forming the dot) an additional peak appears, which corresponds
to a weakly excited metastable state. This peak vanishes for both smaller
and larger dimensions of the dot, and then the ground-state peak dominates
in the spectrum. In the presence of a magnetic field, a pair of metastable
states appear (for not too high fields and medium dot sizes), allowing us to
expect that three peaks would appear. A very strong magnetic field leads to
the disappearance of both metastable states, and then only the main peak
remains. This is convincingly confirmed by the measurements of Bayer et al.
(see Fig. 6.14). A similar effect is also obtained by decreasing the dot diam-
eter.
The structure of luminescence peaks, which occurrs in the model dis-
cussed, results from taking into account the fact that the potential confining
an electron in the quantum dot is modified by the presence of a hole bound
to the electron (the two particles together compose an exciton). As it ap-
pears, this effective electronic potential has the form of a double well. This
leads to the occurrence of an excited state of the exciton, having the same
symmetry as the ground state. This excited state is hence metastable against
the emission of an FIR photon. The relative position of the ground-state and
excited-state energies strongly depends on the geometrical dimensions of the
quantum dot, as well as on the external magnetic field. A strong field re-
places the effective double-well electron potential by a single well. A weak
field leads to a double-well structure also for the effective confining potential
of a hole (magnetic field strengthens the localization of both: the electron
and the hole). This is revealed in the form of two metastable states of an
exciton (for details see Chap. 10). It should also be stressed that the pro-
posed interpteration of observed higher-energy peaks in terms of metastable
exciton states seems to apply to a series of photoluminescence experiments
carried out on dots of different types (measurements of photoluminescence
for varied dot size and magnetic field). This model applies also to the self-
assembled dots, even though the use of the effective-mass method to such
dots is questionable (see Sect. 3.5.1).
6.5 Measurements of Photoluminescence 81
10r---------------------------------------------~

n=3

~8
::i
.!i
Q)
g 6
8III
Q)
c
'E:::I 4
'0
'0
..c
D..
2

1.05 1.10 1.15 1.20 1.25 1.30 1.35


Energy (meV)

Fig. 6.16. Photoluminescence spectra of self-assembled quantum dots for different


values of the excitation power Pexc [114]

Raymond et al. [114, 115] describe the measurements of luminescence from


a system of InGaAs/GaAs self-assembled quantum dots in a strong magnetic
field at a high power of laser excitation. Details concerning the shapes and
sizes of the dots are given in Sect. 8.1 (see Fig. 8.1). Figure 6.16 presents
the photoluminescence spectra measured at different values of the excitation
power Pexc . For Pexc = 0.28 mW (the laser beam focused to the area of
diameter ~ 100J,lm, containing ~ 5.10 5 dots) the spectrum is dominated by a
single peak at energy ~ 1.10 eV, which corresponds to the recombination from
the ground state of a single exciton. With an increase of excitation power,
consecutive peaks appear. For example, at Pexc = 60 m W, four additional
peaks are visible at higher energies: ~ 1.15, 1.20, 1.25, and 1.28 eV. The
fact that the higher-energy peaks appear in the photoluminescence spectrum
only at higher values of excitation power implies that they correspond to
the recombination processes of the following type: N -exciton ground state -+
(N - I)-exciton ground/excited state (with an increase of Pexc the average
number of excitons confined in each dot increases; see Sect. 6.1).
Interestingly, in Fig. 6.16 the rigidity of the measured photoluminescence
spectrum is observed. By rigidity we mean here the lack of shift of individual
peaks toward lower energies (red shift) with an increase of Pexc (i.e., number
of excitons in a dot). In a bulk-semiconductor or quantum-well system such
an effect, the so-called band-gap renormalization, is caused by effe<;tive (elec-
trical) attraction between the charge-neutral excitons, due to the electron-
electron and hole-hole exchange interaction. The magnitude of this renor-
malization can be quite significant. For example, in a quasi-two-dimensional
quantum well that is made of the same materials and has the same thickness
as the described quantum dot and that contains the excitons with concen-
82 6. Interband Optical Transitions

~
'iii
c:
.e
c:
CI)
u
c:
CI)
u
en
CI)
c:
E
::l
'0
'0
..c:
c..

Fig, 6.17. Magnetic-field


evolution of the photolumi-
nescence spectrum of self-
1.10 1.14 1.18 1.22 assembled quantum dots
energy (eV) [115]

tration on the same order of magnitude as in the described experiment at


Pexc = 60 m W, the displacement of the first peak toward lower energy would
be as much as ~ 50 meV. Definitely, no comparable shift is observed in
Fig. 6.16. Detailed analysis of this puzzle is presented in Sect. 8.4.
Raymond et al. [115] describe measurements of photoluminescence in a
magnetic field. The evolution of the photoluminescence spectrum in an in-
creasing field is shown in Fig. 6.17. The splitting of higher peaks, which
signals the magnetic-field-induced removal of single-particle degeneracies, is
visible. The circles represent the results of numerical calculation.
Measurements of time-resolved photoluminescence, i.e., measurements of
the photoluminescence signal (photon emission) as a function of passed time
since the moment of the excitation pulse (photon absorption), are described
in [114] (see also Bockelmann et al. [14, 15]). Such measurements allow for
the determination of the average exciton lifetimes in different initial states
that correspond to different photoluminescence peaks. For example, in [114]
the radiative-recombination lifetime of a single exciton in the ground state
was evaluated as 1 ns, compared to 0.2 ns, the characteristic lifetime that
describes the relaxation between the first optically active excited state and
the ground state.
7. Capacitance Spectroscopy

7.1 Idea of the Capacitance Experiment

The techniques of capacitance spectroscopy are based on the measurement of


the electrical capacity of the system C = dQjdVG (which is related directly
to the density of states dN jdJl) as a function of the gate voltage VG applied
perpendicularly to the quantum well. The quantum well, in which quantum
dots are created, is placed above the doped layer (electrode) and serves as
a reservoir of free carriers. The change of voltage VG leads to a change in
the number of electrons transferred from the well to the dot. Due to the
discrete spectrum of energy levels in the quantum dot, the addition of a
single electron to a dot occurs only at discrete values of VG. The transfer of
the Nth electron from the electrode to the quantum dot occurs when the two
chemical potentials are equal: JlE = JlN. The chemical potential of the dot is
defined as the energy required to add a single electron to the system,
(7.1)
while the chemical potential of the electrode is equal (up to an additive
constant) to its electrical potential:
(7.2)
Thus, the measured consecutive resonance values of VG determine (with the
accuracy of the additive constant) the values of the chemical potential of the
dot JlN for consecutive numbers of electrons. These data give information
about the behavior of the total energy of the system:

E tot = L JlN + const. (7.3)


N

With increasing N, the added electron overcomes the electric repulsion of


more and more electrons, and thus the chemical potential increases with the
number of particles. In the so-called constant-interaction model [4, 5, 80] this
dependence is linear,
(7.4)
where Vint is the constant interaction energy per electron pair, and VI is the
energy of the first unoccupied single-particle level, to which (according to the
84 7. Capacitance Spectroscopy

Pauli exclusion principle) the Nth electron is added. A linear dependence


corresponds, for example, to the situation in which the dot area and the
profile of charge-density distribution do not depend on N (charge density in
each point increases linearly with N). In such a state the interaction energy
per electron pair is fixed and does not depend on N.
However, in a more realistic approach, the system increases in size with
the increasing number of particles N. In the range of densities close to the
maximum density allowed by the Pauli principle, the system size increases
with N even without the interaction, since consecutive electrons occupy the
orbitals with larger and larger radii. Independently, the system size increases
with N, since at a fixed size the total energy of the repulsion between electrons
[ex N(N -1)] grows faster than the potential energy [ex N], and the two have
to compensate for each other. As a result, the interaction is slightly weaker,
and the dependence of the chemical potential on N is sublinear (for large
dots it is I1N ex N 2 / 3 [70, 138]). This corresponds to a decreasing separation
between observed peaks in the capacitance spectrum (resonance values Ve).
This effect, however, is fairly weak for short intervals of N.
In the systems of electrons (charged particles) with dimensions typical
for quantum dots, the interaction between particles is the main factor that
determines the eigenstates and their energies. However, as mentioned before,
the direct measurement of absorption in the far infrared range allows only
for the measurement of the center-of-mass excitations. Hence, capacitance
spectroscopy is a very important tool in the investigation of such systems.
In some experiments [51, 52, 125]' the measured quantity is not the ca-
pacitance C, but its derivative dCjdVe. Due to the small changes of C in
the resonance values of Ve, the signal received from a dot is hardly distin-
guishable from a strong noise coming from the remaining part of the sample.
The measurement of dCjdVe allows for the separation of the signal coming
from the quantum dot, from the noise, and for the observation of the pecu-
liar behavior of C. As a result, sharp peaks in resonance values Ve can be
observed.

7.2 ~easurenaents
by ~eans of Capacitance Spectroscopy

The first capacitance measurements of quantum dots were reported by Smith


et al. [125]. Their experiment was carried out on dots with diameters 100-
400 nm, created in a GaAsj AlGaAs modulation-doped heterost,ructure. The
measurements were made at temperatures 0.4-4.2 K. In the measured depen-
dencies of dCjdVe on Ve, presented in Fig. 7.1, one can clearly resolve the
range of voltage Ve where the dot is empty, a strong peak that corresponds
to the transfer of the first electron to the dot, and a series of weaker peaks
that correspond to capturing consecutive electrons in the dot.
7.2 Measurements by Means of Capacitance Spectroscopy 85

300nm DOTS T=O.5K

200nm DOTS

400nm
DOTS

Fig. 7.1. Capacitance spectra of


-0.9 -0.7 -0.5 -0.3 -0.1 0.1 0.3 0.5 0.7 0.9 quantum dots with different diam-
GATE VOLTAGE (V) eters [125J

As expected, the total energy of the system read from the spectrum in-
creases when the dot size decreases, and so does the distance between subse-
quent peaks (i.e., the chemical potential). It is, however, interesting that the
resolution of peaks becomes worse for smaller dots. The clearest spectra were
obtained for a diameter of 300 nm. As the authors write, this may be caused
by a competition between two opposite effects: (i) decreasing the system size
enlarges the distances between the energy levels and leads to a clearer sep-
aration of individual peaks, and (ii) brings about the important increase in
the effects related to edge defects, which broaden the peaks. It was therefore
impossible to register the peaks for dots with diameter 100 nm. Hansen et al.
[51] for the first time observed the rearrangement of the energy levels of a dot
under the influence of a magnetic field. The dots investigated were prepared
by growing on the AIGaAs surface, slightly above the interface between the
modulation-doped GaAs layer and the AIGaAs layer, the regular lattice of
about 10 5 squares of GaAs, 30 nm thick and with 200-400 nm side length.
Covering this structure with a metal electrode and applying a voltage re-
sulted in the creation of an attractive potential under the GaAs squares, due
to the locally larger distance between the quantum well and the electrode,
and hence in the confinement of two-dimensional electrons from the hetero-
junction. The measurements were taken at temperatures 0.7-4.2 K, while the
magnetic field was varied in the range 0-2 T.
Figures 7.2-4 present three diagrams from the paper cited, measured for
an array of dots with side length 300 nm. Figure 7.2 presents the dependence
86 7. Capacitance Spectroscopy

300nm Quantum Dots


T=0.7K f=10kHz

Fig. 7.2. Capacitance spectra of


quantum dots in different mag-
-0.10 0.00 0.10 0.20 0.30 netic fields; the transition from
spatial to Landau quantization
VG (V) [51]

250 orP •.L "


0·0
ip' 000 . .- 0 o· 0 0
200 •• •
o 000 • 0° 0

...0. 0·0
00 •• 0.

°
",000 _ ••• 0 •
150 .':: 0ni 0 0 o. 0
o
D 00. e • 0 0 ·0
100 ~.- ?~. 0 0 •• 0 0 0
. ; 0 • 0 0
50 _ • ••• cp .....i • 0 0 •.
:>
_E 0
"0 -
00
j~~- [ •
~,,~. _, '!lP[ 0 0
0
o. •
••
0

(!)

._.
DON i~
--'YO
f •

0
00

:> -SO ~•••• ~ ~ •• 0 0 0 0 ••

......... 0.···
• ~ooo~.

...
-100 • .........------, • • • 0

-150
_. ___ •••
o
Fig. 7.3. Magnetic-field evolution

...--,_ ............. .
~_. 0 0 " 0 0 0 0 •••
of potentials VG corresponding to
-2oo~-"•• •
•• • 0 0
••• the maxima in Fig. 7.2 (chemi-
cal potentials of the dot with dif-
-250L----:.---L....--.L...._ _ _ _...J ferent numbers of electrons). Full
0.0 0.4 0.8 1.2 1.6 2.0 (empty) circles - strong (weak)
B (T) peaks in Fig. 7.2 [51]
7.2 Measurements by Means of Capacitance Spectroscopy 87

Fig. 7.4. Capacitance spectra of


quantum dots in a magnetic field
(varied with a step of 0.01 T). The
range of magnetic fields and elec-
-0.10 -0.05 0.00 0.05 0.10
trode potential are marked with a
VG (V) rectangle in Fig. 7.3 [51J

of the capacitance derivative dC jdVG on the gate voltage VG for several


consecutive values of a magnetic field. Visible are the energy shift and change
of intensity of particular peaks, following the rearrangement of the energy
spectrum. Weaker peaks (e.g., the one marked with the empty circle) will
be captured by stronger ones (e.g., the one with the full black circle) in
the strong magnetic field. In Fig. 7.3 are shown the positions of maxima
read from the dependence of dCjdVG on VG. According to Hansen et al.
they correspond to the energies of single-particle excitations, on the other
hand, Silsbee and Ashoori [123] attributed them to filling of energy levels by
consecutive electrons. Full and empty circles denote strong and weak maxima,
respectively, as in Fig. 7.2. In the range of weaker fields (:S: 1 T) the behavior
of peaks is quite complicated, particularly of those with higher energies (VG ;::::
-150 m V), whereas in the fields of 1-2 T a regular structure can be observed
that shows the formation of the Landau levels. The curves dC jdVG in the
range of voltages VG and magnetic fields B, which are marked in Fig. 7.3
with a dotted rectangle, are shown in Fig. 7.4.
Another interesting result obtained by Hansen et al. for quantum dots
containing about 30 electrons [51, 52] in very strong magnetic fields (;::::25 T)
is the peculiar behavior of the capacitance C in the values of VG, which
correspond to fractional filling factors: 1/ = ~ and ~. Figure 7.5 presents the
spectra measured at different temperatures, with indicated peculiar values of
VG . At very low temperatures (T:S: 0.5 K), for the fractional filling factors 1/ =
~ and ~, similarly as for the integer values 1/ = 0 and 1, the downward peaks
appear on the curves, corresponding to the stable electron configurations.
88 7. Capacitance Spectroscopy

300nm dots
f=10kHz
8=30T

ic:
:l -

-
.ci
-
...
CU

> - 0.5K
~"C - 1K
2K
3K
Fig. 7.5. Capacitance spectra of the
quantum dots in a magnetic field of
30 T; minima occurring at low tem-
-O.l -0.1 0.1 D.l 0.5 0.7 peratures correspond to fractional fill-
VG (V) ing factors v = ~ and % [52]

This seems to be closely related to the fractional quantum Hall effect, which
occurs in two-dimensional systems. According to the authors, this is also
confirmed by the observed values of the magnetic field and temperature at
which this effect occurs. It may be also mentioned that in the magnetic field
30 T the magnetic length is as small as 4.5 nm, and the respective diameter of
the 30th orbit in the lowest Landau level is 25 nm. Hence, the dots used for the
measurements (diameter 300 nm) could not produce a spatial confinement
comparable with that due to the magnetic field, and the quantization of
energy in this regime had a purely magnetic (Landau) character, as in a
two-dimensional system.
Ashoori et al. [3, 4, 5] describe the spectroscopic measurements of single
quantum dots by means of a new method, called by the authors single electron
capacitance spectroscopy (SEeS). In [3] the dot diameter was on the order of
1 /-Lm, and the number of electrons was varied over a very wide range, from
zero to several thousand electrons. The measured separation between the low-
energy levels was several meV. This was much more than was expected as a
result of the spatial confinement in a dot with diameter 1 /-Lm. Therefore, the
authors concluded that the electrons were localized in smaller areas inside
the dot, in the local potential fluctuations.
In [4] Ashoori et al. report the investigations that were carried out on a
smaller dot, with diameter about 350 nm, containing 0-50 electrons. The con-
finement of electrons was obtained in the same way as reported by Hansen
et al. [51], i.e., by growing a circular GaAs disk over the GaAs/ AIGaAs
quantum well, and covering it with a slightly wider chrome electrode. The
7.2 Measurements by Means of Capacitance Spectroscopy 89

Fig. 7.6a,b. Quantum dot with


diameter about 350 urn and
number of electrons varying
in the range 0-50. (a) Cross-
section of the sample with
marked configuration of layers;
the arrow indicates electron tun-
neling between the dot and the
electrode. (b) Capacitance spec-
trum obtained in the SECS
method; upper curve - signal in
phase with the voltage oscilla-
tion, lower curve - signal de-
layed by 90° [4J

cross-section of the sample, with marked configuration of layers, is shown in


Fig. 7.6a. In the experiments described, the application of voltage drop VG
to the electrodes indicated in Fig. 7.6a with the value at which the chemical
potentials of the electrode and the dot are equal, leads to the transfer (tun-
neling) of a single electron from the lower electrode to the dot. The electron
transfer induced a small electric charge on the surface of the upper electrode.
In order to accurately measure this charge, and hence also the value of the po-
tential VG , at which the electron transfer occurred, a small voltage oscillating
in time was applied in addition to the constant voltage VG (with the oscilla-
tion frequency 210 kHz). When the two chemical potentials were equal, as a
result of the oscillations of the additional voltage, instead of a single transfer,
the electron traveled back and forth between the lower electrode and the dot,
tunneling through the narrow (~ 10 nrn) Alo.3Gao.7As barrier, as indicated
with the arrow in Fig. 7.6a. The electron motion induced a charge on the
upper electrode, measured with an electronic system using a sensitive lligh
electron mobility transistor (HEMT).
The measurements were made at temperature 0.35 K, while the inten-
sity of the applied magnetic field was varied in the range 0-10 T. A sample
dependence of the measured signal on voltage VG is shown in Fig. 7.6b (up-
per curve). A peak which corresponds to the transfer of the first electron
appears at VG = -373 m V. Consecutive peaks, which correspond to the ad-
dition of subsequent electrons to the dot, are roughly equidistant and have
similar heights, up to the 25th peak. The possibility of resolving so many dis-
tinct, sharp peaks proves the excellent precision and sensitivity of the SECS
90 7. Capacitance Spectroscopy

Fig. 7.7a,b. Chemical


potential of a quantum
dot as a function of the
magnetic field, obtained
with the SECS meth-
od; Numbers of electrons
are indicated. Length of
the white segments in
both diagrams is 5 meV;
dashed line - energy of the
lowest Landau level ~ nwc .
Circles in diagram ~a) -
transition to the l/ = 1
state; triangles in diagram
(b) - transition to the
Magnetic Field (Tesla) l/ = 2 state [4]

method. The weakening of peaks beginning from a certain number of elec-


trons is caused by the extension of the characteristic tunneling time above
the period of oscillations of the forcing voltage. This is confirmed by the lower
curve in Fig. 7.6b, which represents the signal recorded in the detector, which
was delayed in phase by 90° with respect to the voltage VG. Here the high
peaks appear beginning with the 25th.
The main result of Ashoori et al. [4] is measured with an accuracy
unattainable before the dependence of the total (ground-state) energy of
the quantum dot on magnetic field, presented in Fig. 7.7. Figures 7.7a and b
present the evolution of a chemical potential of systems with the number of
electrons 1-35, with an increase in the intensity of the magnetic field from 0
to 10 T. The ground-state energy of an N-electron system is given by (7.3).
The lowest curve, which describes a single electron confined in a dot, follows
the lowest Fock-Darwin level in Fig. 3.2. Based on the course of this curve in
the range of weak fields, and assuming the parabolic shape of the confining
potential V = ~m*w5r2, its parameters were determined: excitation energy
nwo = 5.4 meV and the dot diameter d = 40.8 nm. The upper curves de-
scribe the energy of interacting N-electron systems, and the single-particle
Fock-Darwin approximation is no longer sufficient.
An intriguing result is the occurrence (in all curves except for the lowest
one) of characteristic small peaks, or rather inflections (e.g., in the curve
for two electrons, at the field ~ 1.5 T), which signal the crossings of energy
levels that correspond to different system configurations, and hence to the
7.2 Measurements by Means of Capacitance Spectroscopy 91

transitions in the ground state induced by a magnetic field. In Fig. 7.7 they
are marked with circles and triangles. As shown in Fig. 3.2, the ground-
state energy of two noninteracting electrons (both particles on the 100) Fock-
Darwin orbital with opposite spins) increases monotonously with the increase
in the magnetic field. Thus, the occurrence of peaks must be related to the
effects not taken into account within the simple approach that led to Fig. 3.2,
for example with the spin-orbit interaction and/or interaction between the
neighboring dots. It seems that the spin-orbit interaction [70] (see Chap. 9)
gives a proper explanation of these small inflections on the curves expressing
the dependence of the chemical potential on the magnetic field, which are
measured by means of capacitance spectroscopy.
In numerous works attempts have been made to explain the occurrence
of the characteristic irregularities in the dependence of the energy (or the
chemical potential) on the magnetic field without taking into account the
spin-orbit interaction, but based rather on an analysis of the crossings of
the single-particle Fock-Darwin levels. In such an approach the transitions
in the electron system under an increasing magnetic field can be classified by
a division into three stages, with respect to the value of the filling factor, (i)
v > 2, (ii) 1 < v < 2, and (iii) v < 1, and the mechanism of the transition
that dominates in each stage.
(i) Let us define the average numbers of single-particle excitations with
frequencies w± (see (3.16)) in N-electron quantum state cp:
N N
(n+)¢ = (cpl Laiai Icp), (n_)¢ = (cpl L bib i Icp), (7.5)
i=l i=l

where the pairs of operators (a,a+) and (b,b+) are defined by (3.11). Ac-
cording to (3.19), the asymmetry between the numbers of excitations of both
types is determined by the projection of the angular momentum of the sys-
tem:
(7.6)
The total energy of state cp contains single-particle terms: nw+ (n+) and
nw_ (n_). The first of these quickly increases, while the second decreases
to zero with an increase in the magnetic field. In stronger fields the higher-
energy excitations (nw+ ~ nwc ) correspond to the excitations between Lan-
dau levels, whereas the lower-energy excitations (nw_ « nw c ) correspond to
the excitations within a Landau level.
Due to the exact degeneracy of frequencies w+ and w_ in a zero magnetic
field in the ground state of the system cp in weak fields the projection of the
total angular momentum vanishes: (l z) ¢ = o. The energy of state cp increases
quickly with an increase of the magnetic field and at the critical value crosses
the energy of another state 'Ij; that has a nonzero projection of the angular
momentum (lz),p > 0 (due to the asymmetry of excitation energies, nw_ <
nw+). The state 'Ij; becomes a new ground state of the system. Since Ll (n+) =
92 7. Capacitance Spectroscopy

(n+) 'Ij; - (n+) q, > 0, the change of the ground state can be interpreted as
a transfer of a fraction of electrons (charge) from higher to lower Landau
levels (possibly with a simultaneous change of the spin configuration of the
system). Thus, the resulting peak in the magnetic-field dependence of energy
is a consequence of the crossing of the single-particle Fock-Darwin levels. A
further increase of the field is accompanied by subsequent transitions, until
all the electrons drop to the lowest Landau level: (n+) = 0 in the state
with filling factor v ~ 2. Transformations of the system of 1-6 electrons that
correspond to this stage are described in [133J (see Sect. 8.2).
(ii) A further increase of the magnetic field leads to a compression of the
system of electrons in a constantly diminishing area, since in strong mag-
netic fields the characteristic dimensions of electron orbits are scaled with
the magnetic length lB ex 1/~ (see Sect. 4.3). Thus, with an increase of
the field, the Coulomb energy of the system E ex 111B quickly increases. On
the other hand, the increase of the field reduces the energy of single-particle
excitations to states with greater orbital diameter (in the limit B ----* 00 the
Fock-Darwin spectrum rearranges into the system of degenerate Landau lev-
els; see Sect. 3.3). Moreover, in the state with v ~ 2 the total spin of the
system is close to zero, and the (attractive) exchange interactions are weak.
At the critical value of B, a fraction of charge is transferred from the inside of
the dot to the outside (at the same time reversing the spin) in order to mini-
mize the total energy of the system. The spin direction 1 (along the field B) is
preferred by the Zeeman interaction. However, the energy of this interaction
is considerably smaller than the energy of the exchange interaction.
Hence, the reason for transferring a fraction of charge from the inside of
the dot to the outside is the repulsive direct interaction, whereas the rea-
son for the change of spin configuration (increase of the total spin of the
system) is the attractive exchange interaction between electrons. As a result
of subsequent transformations, the system goes through a sequence of states
with an increasing number of tightly packed spin-l electrons and a decreasing
r
number of spin- electrons, until complete spin polarization is reached. The
completely polarized state closing the sequence has filling factor v = 1 and
is called the compact droplet (see Sect. 4.7).
(iii) A further increase of the magnetic field leads to the reconstruction
of the compact droplet, which occurs by introducing the holes to its inte-
rior. This corresponds to a further reduction of the filling factor v. These
transitions are described in Sects. 4.4 and 4.7.
The numerical calculations including the Coulomb interaction between
electrons were carried out, for example, by Hawrylak [55J and Young et al.
[138J (exact diagonalization of the Hamiltonian for a small n~mber of elec-
trons) , and also by Palacios et al. [107J (exact diagonalization and the
Hartree-Fock approximation). It should, however, be stressed t.hat. in this
approach realistic values of the measured parameters can be obt.ained only
for 2-3 elect.rons. The chemical potent.ial of the N-elect.ron dot, equal t.o
7.2 Measurements by Means of Capacitance Spectroscopy 93

B c
;;-
Ql
J D
E

1
~

..-
~

LY -0.1

N
W I

M
i[
>-
OJ

~ -0.2'
Ql
OJ
C
.~
Fig. 7.8. Chemical potential of
(ll three electrons in a quantum dot
.s:::. as a function of the magnetic
()
field; full circles ~ experiment [4J;
empty circles ~ average measure-
-O.3~~~....,.---.,...--...;..,.--"","",
0.0 2.0 4.0 6.0 8.0 ment error>::: 0.03 meV; solid line
Magnetic Field (Tesla) ~ numerical calculation [55J

the difference between the ground-state energies of Nand N 1 electrons,


/-LN = EN - EN~l, has peaks at the values of the magnetic field at which the
change of the ground state of N or N - 1 electrons occurs (a peak appears
in the ground-state energy EN or EN~l)' As a result, the curve /-LN(B) has a
complicated course, very different from that expected within the simple model
that neglects interactions. In particular, the peak appears also for N = 2.
Figure 7.8 presents a comparison of the chemical potential /-L3(B), which was
calculated numerically for an ideal, parabolic quantum dot, with the values
measured by Ashoori et al. (for 2~3 electrons the agreement is good). Figure
7.9 shows in detail the dependencies /-LN(B) for N =2~6, also calculated nu-
merically for a parabolic quantum dot (n.wo = 2 meV). Analogous curves
calculated within the Hartree-Fock approximation for N =1 ~ 15 electrons
and the quantum dot corresponding to the experiment of Ashoori et al., are
presented in Fig. 7.10. Here, as mentioned before, the divergence between the
calculations and the experiment is significant.
Hence, the above explanations of Ashoori's experiment need refinement.
Using models as in [55, 107, 138] it is not possible to obtain a correct de-
scription for more than several electrons. This was pointed out by Palacios
[107], who showed that the assumption of the characteristic frequency of the
confining potential that is in accordance with the course of the curve /-Ll (B)
measured for the first added electron (wo = 5.4 meV) leads to a consider-
able discrepancy with the experiment for the curves calculated for the higher
number of electrons. This effect is evident in the comparison of the results
94 7. Capacitance Spectroscopy

15.6 -
~
E 15.4 -
........
,g B [TJ
;C 3 4 5 6 7
..-IN ............ N=6 15
I - -_______
::l 15.2 .. -....... N=5
-----------
N=4
---_ _____ 10
.......... ~- N=3
15.0
N=2
.-------------___ 5

2 3 4 5 6 7
B [T]
Fig. 7.9. Dependence of the chemical potential of six electrons in a quantum dot
on the magnetic field; pairs of numbers L,28 denote the states (L - angular mo-
mentum, 8 - spin projection). Inset: chemical potential for two to six electrons
[138]

of Palacios in Fig. 7.10 with the experimental data in Fig. 7.7. Clearly the
distances between the consecutive curves f..LN(B) do not agree (e.g., the cal-
culations give: f..L2(0) - f..Ll(O) = 7.6 meV instead of 4.2 meV measured by
Ashoori), nor do the values of the magnetic field at which the inflections on
the curves appear (e.g., for N = 15 the theoretical value of the field at which
the irregularity of a chemical potential occurs is 8 T, while the experimental
value is on the order of 2 T). Although the first discrepancy could be elimi-
nated by the effective reduction (~ 2 x) of the interaction between electrons
(e.g., by assuming a larger broadening of the electron wave functions across
the dot), the elimination of the second discrepancy requires (on the contrary)
the interaction to be enhanced. This proves that another effect plays a cru-
cial role in the Ashoori's experiment, which has been neglected so far in the
calculations.
Hallam et al. [50] analyzed the influence of the screening of the inter-
actions between the electrons confined in the quantum dot by the electric
charge present in the electrodes. They showed that for the parameters that
correspond to the Ashoori's experiment (the distance between the dot and
each of the two electrodes equal to 80 nm) the screening leads to a consider-
able lowering of the curves f..LN(B) and to a decrease in their separation. For
instance, the separation f..L3(0) - f..L2(0) is reduced due to screening from 2.1 to
1.4 meV (~30%), and f..L4(0) - f..L3(0) decreases from 2.7 to 1.5 meV (~45%).
7.2 Measurements by Means of Capacitance Spectroscopy 95

v>2 .v=2 2<v<1 .v=l


~
.
.
70 ~

.
--- •-----
~

60

~
• 0
.N - 10

- •

D

• --
..
0
..-
0----, •

..
.. O~



N-S_
-
D

--

----
~
~

------
20

to

o- Fig. 7.10. Dependence of the chemical po-


123 4 S 6 789
tential of a quantum dot on the magnetic
B (T) field [107]

According to [107J, the effectively weaker interaction between the electrons


leads to an even stronger discrepancy between the calculated and measured
positions of the inflection points on the curves J-LN(B). Hence, another mecha-
nism explaining the observed effect must be found. A good description of the
experiment can be obtained when in the quantum dot (the artificial atom)
the spin-orbit interaction is taken into account in an analogous way as in
natural atoms, i.e., by the single-particle term DC leT in the Hamiltonian, as
was demonstrated by Jacak et al. [70J (see also Chap. 9). The inclusion of this
interaction leads to a subtle structure of the electron energy spectrum, and
further in the occurrence of inflections on the energy and chemical-potential
curves, in agreement with the experiment of Ashoori et al. Of crucial impor-
tance is the fact that at the same value of the coupling constant ({3 = 0.3,
i.e., on the same order of magnitude as the value for the atoms forming the
quantum dot) this model allows for a description of another, independent ef-
fect: the splitting of the FIR absorption spectrum and the forbidden crossing
of the pair of modes observed by Demel et al. [34J (see the description of their
experiment on page 54).
Ashoori et al. also observed a strong reduction of the intensity of electron
tunneling between the electrode and the quantum dot at certain values of the
magnetic field. In Fig. 7.7 this is visible as the chemical-potential curves van-
ish above the critical fields. The oscillator strength of the transitions between
96 7. Capacitance Spectroscopy

1.0 r - - - - - - - - - - - - -

1.0

6----
<:I 1.0
Fig. 7.11. Oscillator
strength ("'tunneling
intensity) of the transi-
1.0
tions between the elec-
trode and the quantum
dot, determined for two
to five electrons, as a
function of the magnetic
B (T) field [107]

the corresponding ground states of Nand N - 1 electrons (proportional to


the tunneling intensity) was determined by Palacios et al. [107] and qualita-
tively reflects such behavior (see Fig. 7.11). In particular, the intensity of a
transition can drop to zero, due to the spin selection rules, if the difference
between the total spin projection onto the magnetic field in the initial and
final systems (systems of Nand N - 1 electrons) exceeds ~.
8. Description of the Properties
of Self-Assembled Quantum Dots
Within the Band-Structure Model

8.1 Electronic Structure

A method for creating self-assembled dots (SAD's) is presented in Sect. 2.6.


In this chapter we shall concentrate on self-assembled dots in the shape of
spherical lenses shown in Fig. 8.1 (for the description of properties of pyra-
midal SAD's see [47]). Such a dot is formed on a narrow wetting layer of
thickness t w , and modeled as a part of a sphere of height h and radius at the
base s. The conduction-band edge in the material of the wetting layer and
the dot lies below that in the surrounding material [114, 115J. Thus, the elec-
trons are confined in the narrow wetting-layer quantum well due to the step
in the conduction-band edge at the interface, and they are further localized
in the area of the dot due to the locally increased thickness of the layer. The
effective lateral potential V(r, z) that acts on the electrons confined in the
wetting layer is shown in the corner of Fig. 8.l.
One should, however, bear in mind that the self-assembled quantum dot
is not an infinite crystal structure, but rather a small object, with a diame-
ter on the order of 50 lattice constants. However considerably it may simplify
matters, the assumption that the electron (and hole) motion can be described
within the band-structure model, with the band parameters defined by the in-
finite semiconductor lattice (the bulk material of which the dot was made), is
a disputable approximation. We shall use this approximation for the descrip-
tion of both electrons and holes further in this chapter. Another approach,
which is based on the effective-mass method, will be presented in Chap. 10.
The parameters of the pair of materials that make up the quantum dot
and the substrate appear in the model through the effective units of energy
and length: the effective Rydberg Ry* = e 2 /2wB == 1, and the effective Bohr
radius aB = En2 /m;e 2 == 1, where m; is the electron effective mass and E
is the dielectric constant. The effective parameters Ry* and a B take into
account all the effects due to stress, discontinuity of the effective mass and
dielectric constant at the interface, etc.
The Schrodinger equation that describes the motion of a single electron
can be written in cylindrical coordinates in the form
2
1 ( r 8r
[ - r2 8 r 8r
8 8 )
+ 8e 8 2 + v.,(r, z) ] 'IjJ(r, e, z) = E 'IjJ(r, e, z), (8.1)
- 8z
2 2
98 8. Properties of Self-Assembled Quantum Dots

InGaAs/GaAs SAD

s=180A

Fig. 8.1. Schematic drawing of a lens-shaped self-assembled quantum dot; top-


left corner - effective lateral potential that confines electrons due to the locally
increased thickness of the quantum well [134]

where Ve = 0 inside the wetting layer and the dot, and Ve = Vo within the
barrier (Vo = ..:1VCB is the difference between the conduction-band energies
in the two materials, including the effect of strain).
In order to solve this equation two methods were used: (a) The full diag-
onalization of the Hamiltonian in the space of three-dimensional wave func-
tions. In order to construct a convenient set of discrete basis states, the
system was embedded in a large cylinder with an infinite potential barrier
at the edges; (b) The adiabatic approximation. Since the full diagonalization
showed that the adiabatic approximation gives very accurate results, we shall
discuss here only the second method, which is more physically intuitive.
In the adiabatic approximation we can take advantage of the fact that the
electron wave functions are strongly confined to the lowest subband in the
narrow wetting-layer quantum well. Thus, we can write the wave function as
a product

'IjJ(r, e, z) = ~eime 9r(z)fm(r), (8.2)


v21l"
where 9r(Z) is a slowly varying function of r. In the subspace of functions with
integer angular momentum m, the functions 9r and f m satisfy the system
of equations:
8.1 Electronic Structure 99

(8.3)

(8.4)

First, the energy Eo(r), which corresponds to the electron motion in direction
z, is found for a given layer thickness, i.e., at a given distance r from the center
of the dot (Eo (r) is the effective lateral confining potential V, shown in the
corner of Fig. 8.1). Next, the equation for the radial motion of an electron
in the effective potential Eo (r) is solved numerically, separately within each
subspace with a fixed value of angular momentum. In order to solve this
equation, the potential Eo(r) is approximated by the n-step function
(8.5)
where 0 ::; i ::; n, So = 0, Sn+1 = 00, Si < SHI, and Vi < VHI. The wave
function that corresponds to energy E and angular momentum m can be
written in a compact form
(8.6)
where k; = IE - Vii and (F, G) is the appropriate pair of cylindrical Bessel
functions: (Jm, Y m ) for E > Vi and (Km, 1m) for E < Vi. The wave functions
must be continuous and smooth at each interface:
Ai-IFF + Bi-1Gf AiFiR + BiGf (8.7)
Ai - I V' FF + B i - I V'Gf AiV'FiR + BiV'Gf· (8.8)
In the above equations we use the following notation:
FF == F(ki-ISi ), V'FF == ki-lF'(ki-ISi),
Ff' == F(kisi ), V'FiR == kiF'(kis i ), F = F, G. (8.9)
The coefficients (A, B) are related to the transfer matrix Tj by

[BAj] = Tj [AB j-
J J- I
I ]. (8.10)

Based on (8.7-8.8):

(8.11)

Solving the system of equations (8.7-8.8) is equivalent to finding the total


transfer matrix T, defined as

(8.12)

From the condition that the wave functions are finite in the center of the dot
we obtain
100 8. Properties of Self-Assembled Quantum Dots

Bo = o. (8.13)
The energies of bound states Eb are obtained from the condition that the wave
functions vanish in the limit of the infinite radius. This additional boundary
condition,
Bn=O, (8.14)
is equivalent to the vanishing of the corresponding element of the total trans-
fer matrix,
(8.15)
for the discrete set of energies E b .
Electrons with the energies E from the interval that corresponds to the
two-dimensional continuum of states in the wetting-layer quantum well move
in a scattering potential of the quantum dot. This scattering modifies the
electron density of states. The phase shift 'f]m (E), which corresponds to energy
E and angular momentum m, can be expressed in terms of the transfer matrix
as
'f]m = arg(Tu - iT21 ), (8.16)
whereas the change in density of states due to the presence of the scattering
potential reads

oD(E) = L ~ ~~ == LODm(E). (8.17)


m m

Sharp maxima in the density of states correspond to scattering resonant


states, through which it might be possible to populate the bound states in
quantum dots efficiently, for example in op}ical experiments with high exci-
tation power (power of the laser beam). '
Figure 8.2 shows the energy spectrum of an electron in a dot made of
Ino.5Gan.5As and GaAs, with dimensions s = 18 nm, h = 4.4 nm, and tw =
1.6 nm [114, 115]. The effects due to strain were included by calculating
the energy shift caused by the homogeneous hydrostatic pressure and the
splitting of the valence band under an uniaxial stress. Assuming that the
discontinuity of the conduction-band energy on the interface between the
two materials equals 67% of the overall discontinuity in the band gap, one
obtains the effective potential wall that confines electrons in the InGaAs
material: Vo = 350 meV. The effective mass in the strained Ino.5Gan.5As
layer is roughly equal to that in GaAs: m; = 0.067m e , and so is the dielectric
constant: E = 12.5. In the inset in Fig. 8.2 a change in the density of states
oDm (E) due to the presence ofthe scattering potential of the dot is shown, in
the energy range that corresponds to the continuum of states of the wetting-
layer quantum well. For the sake of clarity, only several curves that correspond
to the strongest resonances are shown, for the values of angular momentum
m = 1, 3, 6, 7 (the state with m = 5 is very weakly bound).
8.1 Electronic Structure 101
350

300

250
~
g rr
>- 200 a
.... I -- 1 c:

--
C>
Q)
c
II -- 3
::J
0.
Q)
150 II ~ ---- 6 en
6D m ,I ------ 7 Q)
I ,I
II CD
I ,I en
100 I " "
o Tc.: "
t!_~\_~_#

260 E (meV) 350


50
0 2 3 4 5 6 7
angular momentum
Fig. 8.2. Energy spectrum of an electron in a self-assembled quantum dot. The
horizontal axis gives the angular momentum. The shaded area shows the energy
range corresponding to the continuum of states in the wetting~layer quantum well.
Inset: change in the density of states due to the presence of the quantum dot [134]

The energies of discrete bound states and wetting-layer resonances form


a spectrum composed of almost degenerate and almost equally spaced levels.
They can therefore be very accurately approximated by the truncated spec-
trum of a two-dimensional parabolic well, with the characteristic frequency
fY.ve = 30 meV. The number of bound shells is five, and the total number of
bound states (including the degeneracy due to spin ±u and angular momen-
tum ±m) is thirty.
The average size of the dot in a real structure depends on the growth
conditions. On the other hand, the height of the potential barrier Vo in the
InGaAs/GaAs samples changes with indium concentration. Below we present
the dependence of the energy spectrum on these parameters, i.e., the number
of bound states and the average separation between the shells We' The results
are given in Fig. 8.3, where the electron energy is measured from the bottom
of the continuum of states in the bulk material of the barrier. The upper
frame presents the dependence of the spectrum on the size, given by the dot
radius s at the fixed height-to-radius ratio h/ s = 0.24 and the potential depth
Vo = 350 meV. Beginning with a single, weakly bound state with the orbital
angular momentum m = 0 (s-type orbital) for the radii below 8 nm, it is
possible to follow binding of consecutive shells, coming down to the dot from
the continuum of states in the wetting-layer quantum well with an increase
of radius s. The intershell separation decreases with an increase of the radius,
102 8. Properties of Self-Assembled Quantum Dots

-100

~
.s
'"
<{
'"
Cl -200
>I
W

A.V cb =350meV
-300 h/s=O.24
a 100 200 300
s (A)
a

~ -100
.s
'"
<{
'"
Cl
>I -200
W

-3001-~~~'-~~~~~~~~~~~
o 100 200 300 400
AV cb (meV)
Fig. 8.3. Dependence of the energy spectrum of an electron confined in a self-
assembled dot on the dot radius (upper frame) and on the height of the potential
barrier at the InGaAs/GaAs interface (lower frame). The arrows indicate the spec-
trum in Fig. 8.2 [134]

and the spectrum, which corresponds to an almost perfectly parabolic well


for small dot dimensions, evolves toward a more complicated system of levels
for larger dots (the broadening of higher shells becomes comparable with the
distance between shells).
The lower frame shows the dependence of the energy spectrum on the
potential depth Vo for the fixed size of the dot s = 18 nm and h = 4.4 nm.
Similarly to the upper frame, an increase of the number of states can be
observed with an increase of Vo. Here, however, the separation between the
shells We increases with an increase of Vo, and thus with an increase of the
8.2 Electron System in a Magnetic Field 103

number of bound states. For the dot size chosen for this frame (based on
[41, 114, 115]) clearly separated shells occur in the whole range of Vo.
Connecting both dependencies in Fig. 8.3 it is possible (at least in princi-
ple) to design structures containing the self-assembled dots with an arbitrary
number of bound shells, which are separated by an arbitrary energy We'
Another piece of useful information that can be derived from Fig. 8.3 is a
possibility of connecting a measured distribution of dot sizes in a macroscopic
sample with an inhomogeneous broadening of the density of states. It is
obtained that the homogeneity of dot sizes in the described sample [41, 114,
115] is sufficient to preserve the discrete density of states characteristic of the
zero-dimensional system.
As shown previously, the energy spectrum of an electron confined in a
self-assembled dot can be to a good approximation described as a truncated
spectrum of a two-dimensional parabolic well. Let us now proceed to the
states of a valence-band hole. Due to the strain in the InGaAs layer, the sub-
bands of light and heavy holes in the valence band are strongly split, and the
mixing of states from these two subbands is considerably reduced. Thus, the
states of a hole in a dot, within the band-structure model, can be described
in a way similar to that of the electron states, only with different effective
mass (which for a hole is no longer isotropic) and potential discontinuity at
the interface. Assuming the appropriate parameters the energy spectrum of
a heavy hole in a dot was determined. It is very similar to the electron spec-
trum. The number of the bound, almost degenerate shells is also five, while
the distance between the shells nwh = 15 meV.
The Hamiltonians of the electron He and hole Hh confined in a two-
dimensional parabolic well in a magnetic field B along the z-axis can be
diagonalized analytically. The single-particle eigenstates are the Fock-Darwin
states described in Sect. 3.3. We have checked (carrying out exact numerical
calculations, similar to those described for a zero-field) that the magnetic-
field evolution of the electron and hole energy levels for the realistic confining
potentials V(r, z) is almost identical to that of the Fock-Darwin spectrum,
truncated to the appropriate number of the lowest levels.

8.2 Electron System in a Magnetic Field


In the calculations presented in this section we assumed the parameters that
correspond to a small Ino.5Gao.5As/GaAs self-assembled quantum dot, with
dimensions as described by Drexler et al. [37]: radius s = 10 nm and height
h = 24 nm. As mentioned in Sect. 8.1, the single-particle states in 'small self-
assembled dots are very well approximated by the respective Fock-Darwin
states (described in Sect. 3.3). Figure 8.4 presents a comparison between the
magnetic-field evolution of energy levels of a bound electron (i) calculated
numerically using the method described in Sect. 8.1, i.e., within the band-
structure model (circles), and (ii) the Fock-Darwin energy levels for the best
104 8. Properties of Self-Assembled Quantum Dots

150

~
g
>-
!?
(J) 100
c:
(J)

dots - exact eigen-energies


lines - Fock-Darwin levels
R=O
50

o 6 10 16 20 25 30
magnetic field (T)
Fig. 8.4. Exact energy levels of a self-assembled dot (circles) and the Fock-Darwin
levels (lines) in a magnetic field [133]

fitted characteristic energy of the confining potential fu.,;o = 50 meV (lines).


The angular momenta of orbitals are indicated. The Zeeman energy splitting
is very small (:::::: 1 meV for B = 30 T), so the exact states and the Fock-
Darwin orbitals are approximately doubly degenerate. Due to the very small
dimensions of the dot, for fields B =0-30 T it has only 3-4 bound electron
states (disregarding spin degeneracy). In strong fields the consecutive states
are bound, forming the lowest Landau level in the B ----7 00 limit. However,
in the case of such a strong spatial confinement of the system, the magnetic
fields at which the Landau quantization dominates need to be as strong as a
few tens or more T (we = Wo above B = 30 T). The electron bound levels that
are included in further calculations are indicated in Fig. 8.4 with continuous
lines. The dashed lines correspond to the Fock-Darwin levels that are not
actually bound, due to the finite depth of the confining potential. The trace
of each almost bound Fock-Darwin state is a resonance (peak in the density
of states) in the continuum of states in the wetting-layer quantum well (see
page 100).
In order to calculate the eigenstates of many electrons, the single-particle
states will be approximated by the Fock-Darwin states for fu.,;o = 50 meV.
The Hamiltonian of the system can be expressed in the second quantization
as follows:
8.2 Electron System in a Magnetic Field 105

(8.18)

where c;t (co<) is the operator creating (annihilating) an electron in the single-
particle (Fock-Darwin) state a == In+n_a"), and Vee is the matrix element
of the Coulomb interaction, given explicitly in [56] (see 4.32 for the lowest
Landau level).
The Hamiltonian (8.18) conserves the total angular momentum L, as well
as length S and projection Sz of the total spin of the system S. The diago-
nalization will be carried out separately in the subspaces of states with fixed
values of (R, Sz)' The basis of N-electron states is constructed in the form
of the properly antisymmetrized products of the Fock-Darwin single-particle
states:
(8.19)
For each value of the magnetic field only the four lowest Fock-Darwin states
are used (marked with continuous lines in Fig. 8.4), which correspond to the
bound single-particle states and (in weak fields) to the resonance with angular
momentum L = 2. The presence of higher-energy states, i.e., the interaction
of the Fock-Darwin bound states with the continuum of scattering states is
neglected. However, the presence of the continuum is included indirectly, as
it has an effect on the calculated single-particle states and the characteristic
frequency Wo (see Sect. 3.3). In Fig. 8.5 we present the ground-state energies
of the system of 1-6 interacting electrons in a magnetic field, calculated by
the numerical diagonalization of the Hamiltonian (8.18). The curves that
correspond to different numbers of particles N are vertically shifted, so that
they can be shown in one frame. The energy values in the absence of a
magnetic field are given on the vertical axis, and the arrow in the top-left
corner shows the energy scale common to all curves. The pairs of numbers
in parentheses (total angular momentum, total spin projection) = (L, Sz)
define the system configuration. Due to the small dimensions of the quantum
dot, the characteristic interaction energy (00; OOIVeeIOO; 00) ~ 30 meV (100) is
the lowest Fock-Darwin level) is considerably smaller than the characteristic
confinement energy fiJ.,Jo = 50 meV. As a result, the N-electron ground states
are close to the configurations with the lowest single-particle energy (kinetic
energy plus the confining potential of the dot). Subsequent electrons occupy
the lowest available single-particle states (according to the Pauli exclusion
principle), and the coupling between the states with different single-particle
energies is weak. Thus, the main effect of the interaction is the displacement
of the energy spectrum of the noninteracting system toward higher values
by the interaction energy, weakly dependent on the magnetic field and the
system configuration (but strongly dependent on the number of particles).
The Coulomb interaction and the spin configuration of the system play
a crucial role when in the absence of interactions the ground state is de-
generate (partially filled shells). The ground states of the system of one to
106 8. Properties of Self-Assembled Quantum Dots

~
g (0,0)
6; 804 5
a;c: 606
Q)
Q)
+-'
ctI 4
+-'
C/)
I
"0
c:
::J
o
L-
0)

1301------
50~-------------

o 5 10 15 20 25 30
magnetic field (T)
Fig. 8.5. Dependence of the ground-state energy of two to six electrons on the
magnetic field. The curves are vertically shifted, and the vertical axis gives the
energies in the absence of a magnetic field. The arrow in the top-left corner gives
the energy scale that is common to all curves. Pairs of numbers on the curves are
(L, Sz) [133]

three confined electrons in Fig. 8.5 are to an excellent approximation the


noninteracting states with the lowest kinetic energies: 1001), 1001; 00 i), and
1001;00 I; 011). The spin polarization, and at even higher fields the transi-
tions between the magic states (see Sect. 4.4 and the description of Ashoori's
experiment on page 88), for two and three electrons take place in the magnetic
fields considerably above 30 T.
The first change of the ground state induced by the magnetic field and
the Coulomb interaction takes place for four electrons, i.e., for two electrons
filling the (almost) four-fould degenerate p-type shell (states with angular
momenta L = ±1 in Fig. 8.4). Without a magnetic field, the pair of electrons
in a partially filled, degenerate shell assumes the maximum length of spin,
Sz = 1 (according to the first Hund rule; see Sect. 4.2), which leads to the
reduction of the total interaction energy by the large exchange term. The
nonzero magnetic field removes the degeneracy of the p-shell, and the single-
particle energy of the maximum-spin configuration increases with an increase
of the field. At B ~ 2.8 T the exchange-interaction energy term becomes equal
to the single-particle energy of the excitation of one of the electrons to the
Fock-Darwin state with angular momentum L = -1, and the system changes
8.2 Electron System in a Magnetic Field 107

its ground state. In the new configuration the pair of electrons, primarily spin
polarized, occupies the orbital with L = +1, and the total spin of the system
Sz = O. This is the configuration with the lowest single-particle energy.
The change of the ground state in the system of five electrons at the
magnetic field of ~ 15 T corresponds to the transfer of an electron from the
L = -1 orbital to the L = +2 orbital, in the presence of a rigid core composed
oHour electrons in the two lowest orbitals L = 0, +1. The transfer takes place
earlier than the crossing of the corresponding single-particle levels, because
the Coulomb energy of the new ground state is lower.
In the case of six electrons, the consecutive two electrons jump from the
L = -1 orbital to the L = +2 orbital. Between the transitions, (when the
pair of jumping electrons occupy different orbitals, the system asSumes the
largest possible length of spin, Sz = 1, in order to decrease the total energy
by the exchange term (analogously as for four electrons).
Figure 8.6 presents the chemical potential of one to six electrons confined
in a self-assembled dot and in the magnetic field, which was calculated within
the band-structure model of a quantum dot. The chemical potential J-lN is
defined as the difference of ground-state energies of Nand N - 1 electrons
(drawn in figure 8.5), i.e., as the energy required to add the Nth electron to
the quantum dot containing N -1 electrons (see Sect. 7.1). The main graph
presents the difference between the chemical potentials of the interacting and
noninteracting systems, J-l - J-lni, i.e., the energy of the Coulomb interaction
between the dot containing N - 1 electrons and the added Nth electron
(the charging energy). The chemical potential of the noninteracting system is
equal to the energy of the Fock-Darwin level to which the electron is added,
i.e., the first unoccupied level (see Fig. 8.4). The transitions in the ground
states of four, five, and six electrons in the magnetic field are accompanied
by the changes of slope of the curves J-l- J-ln;, marked in Fig. 8.6 with circles.
The change of slope for five electrons in the magnetic field of 20.5 T is due
to the crossing of the pair of Fock-Darwin levels (hence this is a change of
ground state in the noninteracting system).
The exact N-electron states are close to the noninteracting, single-particle
configurations, which are only slightly modified by the Coulomb interaction.
Therefore, it can be approximately assumed that the Nth electron is added
to a given, unoccupied single-particle Fock-Darwin state, in the presence of
a repulsive electrical potential, which is produced by the charge of N - 1
initially confined electrons. On the curves in Fig. 8.6 we have marked the
angular momentum and spin of the state to which the Nth electron is added.
The smaller graph in the inset to Fig. 8.6 shows the total chemical po-
tential of the quantum dot J-l, which is the directly measured quantity in the
capacitance-spectroscopy experiments (see Chap. 7). The dashe'd line repre-
sents the energy of the lower edge of the continuum of scattering states in the
quantum well. As can be seen in the figure, only three to four electrons can
be bound in the dot with energy below the continuum. However, a couple of
108 8. Properties of Self-Assembled Quantum Dots

100 - -It ....,


2t -It 2t
2t 2-
-
80 IZ -It
2t
2t 5

~ -It 1t 4

---
60
.5 V
'c
::i.
I
::i. 6
40 1t 3 200
~
5
-
ot 2
_-------4-
--- --- ---
150
3
20 J..L
100
2

o ot N=1
50
1

0 5 10 15 20 25 30

o 5 10 15 20 25 30
magnetic field (T)
Fig. 8.6. Addition energy (chemical potential) of one to six electrons in a quan-
tum dot, as a function of the magnetic field. Angular momenta and spins of the
single-particle states to which electrons are added are indicated. Main frame: differ-
ence between the chemical potentials of the interacting and noninteracting systems.
Inset: chemical potential of the interacting system [133]

additional electrons can be captured in the resonance states above the con-
tinuum edge, which may be stabilized by the potential barrier that appears
at the edge of the dot due to the drop of the confining potential.
The number of electrons that can be bound by a dot depends on its size.
Assuming that the dot radius is 25 nm (instead of 20 nm), we have obtained
the distance between the single-particle energy levels actually measured by
Drexler et al. [37] (41 meV). As a result of the calculations cafl'ied out for
s = 25 nm we have obtained that the maximum number of bound interacting
electrons in a dot is five.
It should also be noticed that the real strength of the Coulomb interaction
between electrons is slightly weaker than that included here. This is due to
8.2 Electron System in a Magnetic Field 109

the finite broadening of the electron wave function across the dot [56] and the
screening by the charged electrode [50] (see also Sect. 4.6). We have checked
that the assumption of the effective broadening of the electron charge across
a dot over a length of 5 nm leads to a reduction of the values of Coulomb
matrix elements by ~ 10%.
Figure 8.6 describes only those processes in which the dot passes from
one ground state (N - 1 electrons) to the other (N electrons). However, in
the case of reduced probability (intensity) of a transfer of an electron from
the electrode to the final ground state of the dot, the transfer of an electron
to an excited state of the system can be observed [107]. This can happen,
for instance, in strong magnetic fields, when the electrons in the electrode
from which they are injected to the dot are spin polarized (transfer of an
electron between the electrode and the dot takes place without a change of
spin). Hence, the capacitance spectrum can additionally reflect the effects
connected with the changes of the state of the reservoir of electrons, i.e., the
metal electrode, induced by a strong magnetic field.
In the first order of perturbation, the probability of transferring an elec-
tron from the electrode to the dot is given by the golden Fermi rule (written
here separately for each spin projection of the transferred electron):

pt <X I (fl L eta. If) 1


2, (8.20)
k
where i is the initial state of the dot (N - 1 electrons), f is the final state (N
electrons), and the summation runs over all bound Fock-Darwin orbitals k
(without spin). In Fig. 8.7 we present the dependence of the addition spec-
trum (energy and probability of transitions) on the magnetic field. The en-
ergies of transitions are given on the vertical axis, and the probabilities P
are proportional to the areas of the circles. In the range of the magnetic field
studied the ground-state configuration of the system of one, two, and three
electrons does not change, and the spectra for these numbers of electrons fol-
low the respective Fock-Darwin levels. Hence, in the figure only the spectra
for N 2: 4 are presented. The continuous lines show the chemical potential of
a quantum dot, as in Fig. 8.6. The difference between the spectra for the two
projections of the added electron's spin (1 on the left, and i on the right)
reflects a large single-particle excitation energy. It is clearly visible for the
values of the magnetic field for which the system of N electrons is not spin-
degenerate (the total system spin S = 0). In such a case the spin-1 spectrum
runs above the chemical potential (e.g., for N = 4 and fields B > 3 T). It
can also be noted that for certain magnetic fields the transition energy does
not depend on spin (disregarding the Zeeman splitting), unlike the' transition
intensity (substantially reduced for spin i, for example for N = 4 and fields
B < 3 T).
Drexler et al. [37] also describe measurements of the excitation spectrum
of quantum dots in the far infrared range (FIR; see Chap. 5). The Hamiltonian
that describes the interaction of electrons confined in a quantum dot with
110 8. Properties of Self-Assembled Quantum Dots

210
200
190
180
170
160
150
240
230 .. ' o ~. ..'

>a;
0 ••••

220
~
.-z 210
LlJ 200
CI)
(!J
I 190
z
LlJ 180
170
250
240
230
220
210
200
190 N=6
o 5 10 15 20 25 o 5 10 15 20 25 30
magnetic field (T)
Fig. 8.7. Evolution of the electron addition spectrum of a quantum dot in a mag-
netic field. Areas of circles are proportional to the intensities of individual transi-
tions; continuous lines give the chemical potential [133]

the FIR electromagnetic radiation can be written in the dipole approximation


(see (4.22)). For the confining potential in the form of an infinite parabolic
well, the FIR radiation couples only with the center-of-mass motion (for small
number of electrons the spin-orbit interaction is weak), and the excitation
spectrum in a magnetic field consists of a pair of modes that correspond
to a pair of characteristic energies of single-particle excitations (generalized
Kohn theorem; see Sect. 4.1). However, as a result of the deviation from the
parabolic form of the confining potential, additional lines in the FIR spectrum
are expected to appear [112]. As shown below, additional lines in the FIR
8.2 Electron System in a Magnetic Field 111

90
80
70
60
50
40
30
20
90
~ 80
..s 70
z 60
UJ
CfJ 50
(.!)
'z 40
UJ 30
20
90
80
70
60
50
40
30
20
o 5 10 15 20 25 o 5 10 15 20 25 30
magnetic field (T)
Fig. 8.8. Magnetic-field evolution of the FIR absorption spectrum of a self-
assembled dot. Areas of circles are proportional to the intensities of individual
transitions; continuous lines give the energies of single-particle excitations [133]

spectrum appear in small self-assembled dots due to a limited number of


bound single-particle levels (confining potential is approximately parabolic,
but with a finite depth). The Kohn theorem is thus violated, and the structure
of the FIR spectrum reflects the changes of the ground state of the dot that
are induced by the magnetic field and interactions.
In the first order of perturbation, the probability of absorption of an FIR
photon is given by the golden Fermi rule and can be expressed by the creation
and annihilation operators of the single-particle excitations: a and b, defined
by (3.11) [56, 112]:
112 8. Properties of Self-Assembled Quantum Dots

N
A(w) ex :L I (fl :L(aj + at + bj + bj) Ii) 2 8(EJ
1 - Ei - w), (8.21 )
J j=l

where Ii) is the initial (ground) state ofthe quantum dot, and the summations
run over all bound final states If), and over all electrons. The operator of the
interaction with the FIR radiation connects the electron states Ii) and If)
with equal total spins and orbital angular momenta differing by ±1.
In Fig. 8.8 we present the evolution of the FIR absorption spectrum of
one to six electrons in an increasing magnetic field. The areas of circles are
proportional to the intensities of individual transitions A(w). The continuous
lines give the energies of resonance single-electron (center-of-mass) excita-
tions nw±, defined by (3.16). The visible deviations from the single-particle
spectrum (an effect of coupling between the center-of-mass and relative mo-
tions of the system) increase when the number of particles increases. For
larger N the small number of the bound single-particle states becomes par-
ticularly important. For instance, at the magnetic field B = 30 T it is not
possible to construct a state of six electrons with total angular momentum
L = 7, and there is no line that corresponds to the energy>::: w_ in the FIR
spectrum. The deviations disappear with an increase in the number of bound
Fock-Darwin states, i.e., when the dimensions of the dot increase.

8.3 Exciton in a Magnetic Field

We can take advantage of the fact that the single-particle states of self-
assembled quantum dots are almost exactly the states of a two-dimensional
harmonic well, and write the approximated Hamiltonian of an exciton (an
electron-hole pair) in a magnetic field in the form

(8.22)
where (the index e refers to the electron and h to the hole) m* are the
effective masses (in the case of the hole it is the mass in the plane of the
quantum well), r are the positions, p are the momenta, A = ~B x r is
the vector potential of a magnetic field B (in the symmetric gauge), and
ware the characteristic frequencies of the effective confining potentials V,
determined by the diagonalization of the single-particle electron and hole
Hamiltonians, as described in Sect. 8.1. The last term of the Hamiltonian
(8.22), Veh, describes the electron-hole Coulomb attraction. For simplicity,
we omit here the Zeeman spin splitting (negligible for GaAs).
8.3 Exciton in a Magnetic Field 113

Let us also reiterate here that in the band-structure model of a quantum


dot we assume that electrons and holes are independently bound in the dot
due to a local inflection of the edges of the valence and conduction bands, as
in the contravariant quantum well (see the discussion in Sect. 3.5.1).
For not too weak confining potentials Ve (Vh), the localization of the
electron (the hole) is most of all due to the potential Ve (Vh), and not due to
the Coulomb attraction between the particles. The separation between the
discrete single-particle levels nwe (nwh) exceeds the characteristic interaction
energy within the pair (Veh), and in this case the interaction can be treated as
a perturbation. Due to very small dimensions, such a situation can be realized,
for example, in the self-assembled dots. In such a case it is suitable to carry
out the numerical diagonalization of (8.22) in the basis of noninteracting
configurations (eigenstates of the noninteracting system), i.e., the basis of
products of respective single-particle states:
In+n~; n~n~) = In+n~) In~n~). (8.23)
In such a basis the matrix elements of the Coulomb interaction can be ex-
pressed in the analytical form [132]

min(N.'t-,n,+) min(N'::.,n'::.) min(Nt,n~) min(N~,n".)

X
L
Pe=O
Pel L
qe=O
qe! L
Ph=O
Ph! L
qh=O
qh!

(8.24)

where l = le + lh (L = Le + Lh) is the total angular momentum of the


pair, Eo = J27fRy* aB/lo,eh is the effective energy scale, Ry* = e 2 /2wB is
the effective Rydberg, aB = E!i2/m*e 2 is the effective Bohr radius, effective
length lO,eh is expressed by the lengths defined for electrons and holes by
(3.12): l5,eh = ~(l5,e + l5,h)' A2 = 1 + 4w 2 /w~ (respectively for electrons and
holes), 2A;h1 = l/A;l + l/Ah"\ P = Nt. +n~ + N~ + n~ - Pe - qe - Ph - qh,
and s = n+ + n_ (8 = N+ + N_) (compare (4.32)).
In the first order of perturbation, the probability of creation or recom-
bination of an exciton in the radiative process X ...... 'Y h denotes a photon
with energy nw), proportional to the intensity of the peak in the spectrum
of absorption or emission of light A(w), is given by the Fermi golden rule
A(w) = L I (fIPli) 2 8(Er - Ei - nw).
1 (8.25)
f
114 8. Properties of Self-Assembled Quantum Dots

Here (in the case of absorption) Ii) denotes the initial state of the system (an
empty dot), and If) runs over all possible final states (a single exciton in the
dot). P is the dipole transition operator, which, neglecting spin and using
the composite index i = [n+, n_], can be written in the form

Pex L ctht. (ielih), (8.26)


ie,ih

where c+ and h+ are the operators creating an electron and a hole in a given
single-particle state. In terms of wave functions, the dipole matrix element
(fIPli) can be written in the form

(fIPli) ex J JdTe dTh ¢/ (Te, Th)8(Te - Th), (8.27)

i.e., the amplitude of the probability of creating an exciton is proportional


to the amplitude of probability of finding an electron and a hole at the same
place in the final state.
The spins of generated/annihilated particles (an electron and a hole) are
governed by the polarization of absorbed/emitted light. In the case of ab-
sorption, the polarization of light determines the spin configuration of the
created pair. In the case of emission, the polarization of the emitted photon
reflects the spin configuration of the pair at the moment of recombination. If
the coupling between the light- and heavy-hole subbands of the valence band
is neglected, the circularly polarized light (]'± creates electrons and holes with
definite spins (within each subband). This is illustrated in Fig. 8.9. For ex-
ample, in the excitation from the subband of heavy holes, the (]' + photon
with angular momentum +1 ((]' _ with angular momentum -1) can excite an
electron with spin i (1) to the conduction band, leaving a hole with spin 1
(r). Hence, when the spin is included, the dipole transition operator for the
circularly polarized light (for the heavy-hole subband) has the form

P+ IX L
io,ih
ht.r (ielih) ,
<1

p- IX L <T ht.1 (ielih)'


ie,ih
(8.28)

The linearly polarized photon (L) excites an electron with the spin function
in the form of the symmetric combination of states 1 and i:

(8.29)

The antisymmetric combination (P + - p _) corresponds to the so-called dark


exciton, which is not optically active (cannot be created/annihilated by ab-
sorption/emission of a photon).
In the case of absorption of a single photon by an empty quantum dot
l -+ X, the absorption spectrum is not sensitive to the polarization of light,
since the probabilities of absorption of a photon with different polarizations
8.3 Exciton in a Magnetic Field 115

j =m+(j
+112 t @
± 112 electron
-112 ~ @

band gap 0'- 0'+

heavy
±3/2
hole
-3/2 ~
~
+3/2 t
t
light +112 ~ t
hole
±112
-112 t !
Fig. 8.9. Interband optical transitions between the heavy-hole (j = ±~) and light-
hole (j = ±~) subbands of the valence band (m = ±1), and the conduction band
(m = 0, j = ±~), for the two circular polarizations of light: 0"+ and 0"_

(creating electrons and holes with different spins) are equal. Denoting the
expansion coefficients of the exciton eigenstates in the basis (8.23) by

Ulie, ih) = eLih (8.30)


(they are determined through the numerical diagonalization of the Hamilto-
nian) one obtains
2

A(w) =L L eLiI, (ielih) 8(Ef - Ei - fu.u). (8.31 )


f ie ,ilt

On the other hand, the polarization of light (spin) plays an important role in
the investigation of emission, when it is possible to control the spin config-
uration in the initial state by the application of an external magnetic field,
and thus to change the distribution of the emitted light into the pair of com-
ponents (J + and (J _ •
As shown in [132]' the effect of polarization appears in the absorption
spectrum when a dot in the initial state is not empty, but contains a confined
electron (a hole). The presence of a particle in the initial state, with a spin
that is controlled, for instance, by the magnetic field, partly blocks the space
of available final states (due to the Pauli principle). For instance, if there
was an electron with spin 1 in the initial state, only the absorption of a (J_
photon can lead to the creation of a bound complex X- (negatively charged
exciton) with a pair of electrons in the spin-singlet configuration 1i - i 1
(zero total spin of the pair). Thus, if we compare a pair of absorption spectra
116 8. Properties of Self-Assembled Quantum Dots

that were registered for the two polarizations of light (J" _ and (J" +, then for the
process "y -+ X (initially an empty dot) they will be identical, whereas for the
process e- + "y -+ X- (initially a bound electron) they will be different: In
the spectrum (J" _ a new peak with the lowest energy will appear (lower than
the energy of the first peak in the spectrum of an empty dot by the electron-
exciton binding energy). The experimental identification of a bound complex
X-, based on a comparison of the two spectra (J"+ and (J" _ , is reported in
[82,83].
Below we present some selected results concerning the calculations of
the absorption spectra within the band-structure model, applied to the self-
assembled dots described in [114, 115]. Figure 8.10 presents the evolution
of the absorption spectrum in an increasing magnetic field B, determined
for the same parameters as in Fig. 8.2. The upper frame corresponds to
the noninteracting electron-hole pair, and the lower frame to the interacting
magneto-exciton. The areas of black circles are proportional to the intensities
of individual optical transitions (peaks in the absorption spectrum), while the
energies of these transitions (measured from the band gap) are given on the
vertical axis.
In the absence of the electron-hole interaction all allowed transitions have
the same intensity, and their evolution in a magnetic field is similar to that
of the Fock-Darwin spectrum of each of the particles. For chosen transitions
we have indicated the orbitals In+n-), in which the electron-hole pair is
excited. Figure 8.10 demonstrates the magnetic-field-induced breakup and
restoration of the dynamical symmetries of the Fock-Darwin spectrum. At
certain magnetic fields (B = 0, 12, 19, and 25 T) the levels cross, and the
intensities of neighboring peaks add up. For instance, at B = 0 the observed
overall intensity increases linearly with energy.
The Coulomb interaction leads to a certain rearrangement of the spec-
trum, as shown in the lower frame in Fig. 8.10. However, the spectrum of
a magneto-exciton slightly resembles the spectrum of a noninteracting pair;
in particular, at B ~ 13 T we observe a reconstruction of almost degen-
erate groups of strong transitions, separated by wide gaps. The reason for
this resemblance are the small dimensions of typical self-assembled dots. The
characteristic energy of a confining potential: fiJ..J e + fiJ..Jh (~ 45 me V) exceeds
the characteristic energy of the Coulomb interaction (~ 25 meV). The main
effect of the interaction is thus a parallel shift of the entire spectrum toward
lower energies (red shift), by the (approximately constant) binding energy of
the pair.
Figure 8.11 presents the absorption spectra in the absence of a magnetic
field, with the inclusion of the inhomogeneous broadening of individual peaks
due to the distribution of dot sizes in the sample. In both diagrams (the non-
interacting electron-hole pair and the interacting exciton) the broadening of
peaks does not exceed fiJ..J e + fiJ..J h , and there are five strong, clearly resolved
maxima, which correspond to five shells in the electron and hole energy spec-
8.3 Exciton in a Magnetic Field 117

I: .. ··i·· ... ,., ...•......


250~----~~~----~--------------~~
x5
·1 : .I·ii " :.: :: .•
• '1'":,, ••• ,., •••••••
· ..'
·.....
'.. !.....
....
.......•..
200 " •• , ". '.:"
~ll: " I •• ••••••• •••
" '
150
.•.. '~.':"'..
3 •I•••
'" .' " I • • • •••

..
x .' • ••••••••
'"~
o>
100 x2 [!Q]... ••• •••••••••
••••• , •• • •••••• ~
~

,::: ........•.•.....•.•..• ~
~ ~
~
•••••••••••••••••• • • • • • • • • • • !QQlo
50 x1

I without interaction
O~~~~~~~~~rr~~~~~~~~
250.-------------------r------------,
· .• ' .... i· .jwith interaction
,:• ...••
• • • :.
• ·a· ... · .~. e.. ..•.
.- :., .._, •
• • •• • ••• io-.......-c..--....---.---=---!
.~~.
200 • •- •• •• z.
••'. :·i.···
.. ~~. .::",i'~~
... .... ,..
:: •.••.. ' - •••
~ ~·l.;·~!·
: ..•••..••
:~.

150
.. ;
'

•••..
... .
....!!~.;:.' •••.
.'
':'.i;~~: :::.
. : •••• !t , . . . . . . . .
. ... , .:'.'.Ii
~~;.'i·· •• : •
' ·~········i
. ' ..
...... ,!
.1 .. '
• ··'~.....

..
•••••• I . "~,.::: • • • i ·' •• '
.
100
::::: :;;
.. .Iil' •••••••••••..
. .....
, ••
•••••••• ....... .
• • • • • • •• ••••••••••••••••••••••
50

••••••••••••••••••••• ••••••••••
o 5 10 15 20 25 30
magnetic field (T)
Fig. B.10. Dependence of the absorption spectrum of an empty self-assembled dot
on the magnetic field. Areas of circles are proportional to the transition intensities.
Upper frame: noninteracting electron-hole pair (degeneracies in the absence of a
field and the Fock-Darwin states for chosen transitions are indicated). Lower frame:
interacting exciton [134]

tra. The inclusion of the interaction results in the transfer of absorption


intensity to the lower-energy peaks. As a result, instead of a linear increase
of the overall intensity with energy (as is the case without interactions), we
observe the suppression of higher-energy transitions and the enhancement of
the lowest-energy transition, which corresponds to the creation of an exciton
in the ground state.
Using a high-power laser, Raymond et al. [114, 115] also observed five
peaks in the photoluminescence spectrum measured for the described sample
(see the description on page 81 and Fig. 6.16). Also, the measured separation
118 8. Properties of Self-Assembled Quantum Dots

5 without interaction
=i 4
.3- 8=0 , CT s =10A
§ 3
'';:::;

o
0.
(/)
2
.0
co

04-~~~~~~~~~~~~~~~

5 with interaction
::j Fig. 8.11. Absorption
.3 4 spectrum of a self assem-
c bled dot (in the absence
o 3 of a magnetic field). Up-
'';:::;
per frame: noninteracting
o
0.
(/)
2 electron-hole pair. Lower
.0 frame: interacting exciton .
co Broadening of peaks re-
flects the distribution of
O~~~~~~~~~~~~~~~~ dot sizes in the sample
o 50 100 150 200 250 ((8) = 18 nm, (Js = 1 nm)
energy (meV) [134]

between the peaks agrees very well with the results of calculations presented
here.

8.4 Condensation of Excitons


In Sects. 8.1-3 we have shown (within the band-structure approximation)
that the single-particle states in the self-assembled dots form degenerate
shells, similar to the electron shells in atoms, and also how this property
is reflected in the absorption spectrum of a single magneto-exciton (joint
electron-hole density of states). Let us now proceed to the description of the
system of a number of interacting excitons that are bound in a quantum
dot [135, 136J. Using a composite index i = [n+,n-J the Hamiltonian of an
interacting electron-hole system can be written in the compact form (in the
representation of the second quantization)
H LEfctCia + LEfhthia
ia ia
- L (ijIVehl kl ) cthja,hka'Cza
ij klaa'

+~ L (ijIVeel kl ) ctcja,Cka'Cla
ijk/aa'
8.4 Condensation of Excitons 119

The operators ct (Cia) and ht (h ia ) create (annihilate) an electron/hole with


spin (J, in orbital i, with the single-particle energy E i . The two-particle matrix
element (ijlvlkl) describes the Coulomb interaction (index ee for electron-
electron, hh for hole-hole, and eh for electron-hole scattering) [132J (see
(4.32) and (8.24)). The space of N-exciton states is spanned by the properly
antisymmetrized products of single-particle states
(8.33)
where Ivac) stands for vacuum. This basis is further used for the numeri-
cal calculation of the exact eigenstates of the system. The combination of
the numerical diagonalization (within partially filled valence shells) with the
Hartree-Fock approximation (for the description of the interaction of the par-
tially filled valence shell with the completely filled lower shells) allowed the
calculations to be carried out for as many as 20 excitons. As will be shown
below, the numerically determined, interesting properties of many-exciton
states can be explained in terms of the hidden symmetries associated with
the Hamiltonian (8.32).
The excitation of an electron from the valence band to the conduction
band can be viewed as creation of the polarization of the system. The many-
exciton dot can therefore be described in terms of the algebra of a three-
dimensional polarization operator P. Three components of this operator have
a simple physical interpretation:
(8.34)

creates/annihilates an electron-hole pair coherently over all orbitals i, and

(8.35)

counts the number of created pairs. The commutation relations that connect
p± and Pz are those of the three-dimensional angular-momentum algebra
[135J

(8.36)

Further, the operator of the squared polarization can be introduced,


1
p2 = _(p+ P- + P- p+) + P; == pcp + 1), (8.37)
2
together with component Pz making a pair of independent commutating vari-
ables.
Since the electron that is excited from the valence band can carry either
of the spins, we are dealing with two kinds of excitons, which are created
120 8. Properties of Self-Assembled Quantum Dots

optically by means of two different circular polarizations of light [132] (see


page 114 and Fig. 8.9). Therefore, in order to describe the spin-unpolarized
electron-hole system, it is necessary. to introduce independent algebras of
polarization for each of the spins (J =1, i.
While the the Hamiltonian (8.32) obviously conserves Pz , the conservation
of p 2 depends on the commutator of the Hamiltonian with p±:

[H,P:]
ijk

+2
1
L ((ijlv"el kl ) - (iklVehlJ l ))
ijkla'
+ h+kaCia'
x (Cja + C/fI' + CjfI'
+ h+ + fI )
kfI' CiaCl

+2
1
L ((ijlVhhl kl ) - (iklv"hljl))
ij klfI'

X (CjfIhthi:r,hifI, + cja' htfI' hi:rhia)' (8.38)


If the electron and hole orbitals overlap, (ielih) = Dic,ih' the interactions that
appear in the Hamiltonian (8.32) are symmetric,
(8.39)
and the last two sums in (8.38) vanish. This applies almost exactly to the
dots described by Raymond et al. [114, 115], where the respective ratios are
Vee/v"h = v"h/Vhh= 1.04.
Let us now consider an isolated, N-fold degenerate shell. We must hence
exclude from (8.32) all intershell scattering processes. For almost-symmetric
interactions the commutator (8.38) can be written in the form
(8.40)
where

(8.41 )

is the energy of electron-hole attraction (the exciton binding energy) in that


shell. Since the exact proportionality [H, P+] ex: p+ implies the commutation
between Hand p2, the eigenstates of the pair of operators (P 2, Pz ) are the
approximate eigenstates of the Hamiltonian H. The N -exciton eigenstates are
obtained by the subsequent application of the operator p+ on the vacuum:
(p+)N Ivac) , (N = 0,1, ... ,N). (8.42)
The (eigen) states obtained in this way will be called the coherent many-
exciton states (CMES). They correspond to the following states in terms of
(P 2 , P z ):
1
P=~N Pz =N - -N. (8.43)
2 ' 2
8.4 Condensation of Excitons 121

Using a convenient analogy between P and the total spin N of a set of spin-~
particles (the mapping c± f--+ at and h± f--+ af), the addition (subtraction)
of excitons to (from) the shell can be viewed as rotation of the polarization
vector P without changing its length P. An empty shell corresponds to P
pointing down (Pz = -~N), whereas a completely filled shell corresponds to
P pointing up (Pz = +~N). The ground-state energy of the shell is linear in
N, and the exciton addition/subtraction energy JL(N) = E(N) - E(N -1) is
a constant. Thus, despite the occurrence of strong correlations in the system,
the consecutive excitons are added to a shell with the same energy, forming
a condensate. If the excitons of both types are added (a = 1, i), the joint
CMES is simply a product of two spin polarized CMES's, and the exciton
addition/subtraction energy does not depend on a.
Dynamical symmetries that lead to the Bose-Einstein condensation of
excitons in a two-dimensional electron-hole system in very strong magnetic
fields are described in [24, 96, 99, 108].
As an illustration, in Fig. 8.12a we have drawn the energy of addi-
tion/subtraction of an exciton JL(N), calculated for the dot that was de-
scribed by Raymond et al. [ll4, ll5]. We have subtracted here the kinetic
energy (constant within each shell), so that the quantity on the vertical axis
contains only the energy of the electron-hole attraction within the added
exciton and the energy of the interaction of this exciton with otherexcitons
in the dot. The deviations from a perfectly flat dependence within each shell
« 0.25 meV) are small compared to the sharp jumps Ex at the transitions
between neighboring shells (1-3 meV), and negligible compared to the char-
acteristic energy of electron-hole attraction (16-23 meV) , or the jump of
kinetic energy between the shells (45 me V).
In the next step we shall include the scattering of electrons and holes to
higher (empty) shells. The main effect of adding the respective terms to the
Hamiltonian (8.32) is the formation of bound bi-excitons, composed of a pair
of electrons and a pair of holes, both in the spin-singlet configuration. Such
spin-singlet bi-excitons, within an isolated shell created with the following
operator,

Q+ -- '12 " + cn
0 ( cil + + Cj1
+ CiT
+)(h+h+ h+ h+)
iT jl + n il ' (8.44)
ij

interact weakly with each another and with excess spin-polarized excitons.
Denoting the bi-exciton binding energy in a given shell by Llxx, we can
write the approximate energy of the ground state (without the kinetic-energy
terms, for Nl > NT):
(8.45)
The consecutive excitons are added to a shell alternately, L 1, L ... (so as
to maximize the number of 1i pairs), and the addition/subtraction energy
splits into a pair of lines, Ex and Ex - Ll xx .
122 8. Properties of Self-Assembled Quantum Dots

-15~-----------------------,
(a) f
-17
d ~
-19
~~
p ,/ P
-21 . ~Ex

....
I

S :
-23 ~
,--....
>Q) 2 5 -'-_ _~. laolated ehelle
*
-

E with sGBtterine
- 15 (b) to hleher shells 1-----,
>-.
..... ..- ........
,.....--- . -....
0)
~
-17 , ~
Q)
~ .~
cQ) -19
/~ Af
c -21
r-..... ·~ ~
o :<!~ ~
~
u
-23 ....,'K"r':,p
,' !:lxx
]
§ - 27
-25
~~
-'-----1* without lower ehelle;
~ - 15 - r - - - - - I . with lower e;hells
~ (C) aomplt:tely flllM

-20 ~~
¢
-25 .,,1 ~~
.. ~B
1 ";tro~ f
IDI:>..
. F~ ~~~-~~-I
-30 , Si""tS
\ g

-35
~,, Fig. 8.12. Exciton addition en-
ergy of a self-assembled dot
,, o 50 100 150 200
(without the single-particle en-
,, energy (meV)
ergy): (a) squares - isolated
shells; (b) stars - with scatter-
-40 ~,, ing to higher shells; (c) circles
,, - with interaction with com-
, pletely filled shells. Inset: en-
-45
~ ergy of addition of a single ex-
citon to different shells (upper
frame), and energies of addition
o 5 10 15 20 of consecutive electrons to the
number of excitons ground state (low'er frame)
8.4 Condensation of Excitons 123

The results of the calculations for a real system are given in Fig. 8.12b,
where for each shell we have included the scattering of one electron and/or
one hole to the higher shell. Such a choice of the Hilbert space is sufficient to
obtain the effect of the bi-exciton binding. It also allows for carrying out the
calculations as far as for the four lowest shells, i.e., up to eight excitons on
shell f. The inclusion of higher-order scattering processes, which limits the
possibility of performing the calculations to lower shells, does not qualita-
tively change the curve shown in Fig. 8.12b. The main effect is the decrease
of the slope of the J.L( N) line within each shell, which then turns out to be
almost perfectly fiat (as in Fig. 8.12a). This means that the bi-exciton bind-
ing energy considerably exceeds the binding energies of larger complexes, and
the system consists of a gas of weakly interacting bi-excitons.
In the last step we shall fill all the shells below the valence shell with
excitons, so that the entire system can be divided into a compact (rigid) core
and a partially filled valence shell. The electrons and holes feel the presence of
the charge-neutral core mainly through the exchange interaction. The inter-
action between the core and the valence shell can be very well approximated
by the single-particle potentials that acts independently on electrons and
holes. Within a given shell these potentials are almost perfectly fiat, so that
the main effect of the presence of the core is a parallel shift of the exciton
addition/subtraction to/from the shell toward higher energies (blue shift).
The size of this shift (approximately constant within the shell) is the energy
of exchange attraction between the valence shell and the core. In particular,
the inclusion of the presence of the core has no influence on the occurrence
of the oscillations of J.L(N) within the shell (bi-excitons binding), or on the
slope of J.L(N) (the lack of interaction between the bi-excitons).
As shown in Fig. 8.12c, the effective renormalization of the energy of
exciton addition/subtraction (i.e., the band-gap renormalization) is a strong
effect, on a scale exceeding the oscillations within the shells (L\xx) by an order
of magnitude. In the inset to this figure we give the addition/subtraction
energies of an exciton with the inclusion of the so far omitted kinetic energy
(constant within each shell). The positions of bars correspond to the energies
of peaks that were measured in the photoluminescence spectrum.
The upper diagram describes the situation where a single exciton is added
to an empty dot (or a single exciton recombines leaving an empty dot). An
exciton can be added to any shell; the four columns correspond to the data for
N = 1, 3, 7, and 13 in Fig. 8.12a. The lower diagram summarizes the results
concerning the filling of the consecutive shells with subsequent excitons. The
following features can be seen: (i) the energy J.L(N) is practically constant
within each shell (the bi-exciton binding energy is not visible in this scale),
(ii) the bars that correspond to consecutive shells are almost equidistant, and
(iii) the distance between them is effectively reduced by ;::::: 25%, compared to
the upper diagram.
124 8. Properties of Self-Assembled Quantum Dots

s g

(f)
c
Q)
+-
c
c
o
+-
0..
\...
o
(f)
..0
o

o 50 100 150 200


energy (meV)
Fig. 8.13. Absorption spectra of one to six excitons in a self-assembled dot; full
peaks correspond to the addition of an exciton to a partially filled shell [136]

The basic feature of the calculated many-exciton spectrum is a constant


value of the exciton addition/subtraction energy within a shell. It can explain
the recently observed rigidity of the many-exciton photoluminescence spec-
tra (the lack of shift of the photoluminescence peaks at increasing excitation
power, i.e., increasing density of excitons) [13,41, 114]. Moreover, the simi-
larity between the two diagrams in the inset proves that it is not possible to
determine, for example, the strength of the confining potential or the size of
a dot based solely on the photoluminescence spectrum, i.e., on the distance
between the peaks. For this, knowledge of the number of excitons created per
dot, the efficiency of relaxation processes in dots, etc., is required.
A striking illustration of hidden symmetries of the Hamiltonian, forbid-
ding the interaction of excitons within the same shell while excitons from
different shells clearly attract each other, is shown in Fig. 8.13. We present
here the full absorption spectra of one to six excitons, subsequently filling the
two lowest shells. We assume that before creation of the next, Nth exciton,
the system relaxes to the ground state. As shown with the dash~d lines, the
energy of adding an exciton is almost constant during the filling of the valence
shell, whereas it decreases with an increase of N if the consecutive excitons
are added to a higher, completely empty shell. Based on the slope of these
lines it is, for example, possible to estimate the average energy of attraction
8.4 Condensation of Excitons 125

between an exciton on shell d and a single exciton from the core (from shells
sand p) as 5.7 meV, visibly exceeding the bi-exciton binding energy on each
of these shells (e.g., Ll~x = 0.3 meV).
9. Description of a Many-Electron
Quantum Dot with the Inclusion
of the Spin-Orbit Interaction

In this chapter we shall present an analytical theory given by Jacak et al. [70],
which is based on the Hartree-Fock approximation and describes a quantum
dot that contains many electrons. The limitation of a large number of elec-
trons (a few tens) results from the perturbational character of the calculation.
However, it allows for the inclusion of the Hartree-Fock interaction with a
controlled accuracy. In addition, the spin-orbit interaction is taken into ac-
count, which seems to explain many subtle characteristics that are measured
for various dots. It should be added that the effects due to the spin-orbit inter-
action become more significant for larger numbers of electrons in a quantum
dot. Its role in this artificial atom, which is squeezed in the plane, is even
more important than in the case of ordinary, three-dimensional atoms.
Let us include the spin-orbit interaction in our model of a quasi-two-
dimensional quantum dot. This can be done in an analogous way as in a
many-electron atom, i.e., through the single-particle potential
VLS = ala, (9.1)
where l and a are the orbital and spin angular momenta of an electron. The
coupling constant a is connected with the average self-consistent field (8U),
that acts on the electron through the following relation [91]:
a = (3 (8U). (9.2)
Let us also reiterate that in an N-electron atom the dimensionless parameter
(3 equals

(9.3)

In a quantum dot the parameter (3 will be treated as a fitting parameter,


and the magnitude of the field (8U) will be calculated based on the electron
structure of the system.
Let us write the total Hamiltonian of the system in the effective-mass ap-
proximation. This Hamiltonian contains the terms that describe the kinetic
energy in a perpendicular magnetic field B, the parabolic confining poten-
tial with characteristic frequency wo, the spin-orbit coupling, the Pauli term
with effective giromagnetic Pauli factor g, and the energy of the Coulomb
interaction between electrons with dielectric constant f:
128 9. Description of a Many-Electron Quantum Dot

n
In the above equation is the Planck constant; m* is the effective mass; rare
the positions; A is the vector potential of the magnetic field B (B = rotA, in
the symmetric gauge: A = ~B[y, -x, 0]); land 0" are the components of the
orbital angular momentum l and of the spin ( j across the plane of motion.
In the Hartree-Fock (HF) approximation the equation for the HF wave
functions 'Ij.; has the form

[HB + Vi]'Ij.;i(rO") + L: j dr' LJ.i(rO",r'O"')'Ij.;i(r'O"') = ci'lj.;i(rO"), (9.5)


,,'
where HB is the Hamiltonian of a single (noninteracting) electron in field B,
given by (9.4), Vi denotes the Hartree potential

Vi = e
2
jdr' ni(r') (9.6)
E Ir' - rl'
where
ni(r) = L: L:'I'Ij.;j(rO")1 2 , (9.7)
" j

and LJ.i is the Fock correction,


e 2 s: L:''Ij.;j(r'O"')'Ij.;j(rO")
Ir' - r I .
A ( ") _
Lli rO", r 0" - --u"',, (9.8)
E
j

Introducing the operator of exchange interaction G as

Gi'lj.;i(rO") = L: j dr' LJ.i(rO",r'O"')'Ij.;i(r'O"'), (9.9)


,,'
the Hartree-Fock equation (9.5) can be written in a compact form
[HB + Vi + G i ]'lj.;i(rO") = ci'lj.;i(rO"). (9.10)
In accordance with Shikin et al. [120] here we will make use of the approx-
imate expression for the charge density in a parabolic quantum dot (with
characteristic frequency of binding potential wo). It was obtained in a classi-
cal regime and is hence applicable particularly for a large number of electrons
N:
n(r)={ n(0)J1-(r/R)2 ifr:::;R, (9.11)
o if r > R,
where n(O) = 3N/27rR2 is the charge density in the center of the dot. For
a classical system, the dot radius R is given explicitly in [120], but here it
9.1 Quantum Dot in the Absence of a Magnetic Field 129

will be determined from the minimum-energy condition, in order to take into


account the quantum corrections (the radius plays the role of a variational
parameter). The Hartree potential determined using (9.11) reads

31fNe ( r2 ) 2
VR(r)=-- 1-- (9.12)
4ER 2R2
(this form is also assumed for r > R). Neglecting for the moment the exchange
term, which later will be taken into account perturbatively, we obtain the
following Hartree equations:
(9.13)
parametrized by a variational variable - the dot radius R.

9.1 Quantum Dot in the Absence of a Magnetic Field

Let us first consider the case of a zero magnetic field. Equations (9.13) then
assume the form

(- 2~* Ll + ~m* [2b2 r2 + ala) 'ljJi(ra-) = (Ci - 3:~;2) 'ljJi(ra-), (9.14)

where the effective frequency appears (the frequency renormalized by the


Hartree term):

[2'2 _ w2 _ 31fNe 2 (9.15)


o - 0 4Em*R3·
Equation (9.14) can be solved analytically, and the eigenstates are
(9.16)
where the spin functions are denoted by XO" (with the eigenvalues a = ±~),

n = 0,1, ... and m = -n, ... ,n, are given in Sect. 3.3.
and the explicit forms of the angular and radial wave functions, labeled by

The eigenenergies connected with functions 'ljJnmO" read


31fNe 2
+ 1) + ama + --R-.
1
CnmO" = fi[2o(n (9.17)
410
Without the spin-orbit interaction (a = 0) these energy levels form degener-
ate shells, labeled by n. The nonzero a splits these shells to doubly degenerate
sublevels.
In the Fock-Darwin representation (see Sect. 3.3) an equivalent notation
is used:
'ljJi(ra) = 'ljJn+n_O" (ra) = rPn+n_ (r)xO", (9.18)
with n± = 0,1, .... The eigenenergies labeled by n± (n
m = n_ - n+) and a are
130 9. Description of a Many-Electron Quantum Dot

(9.19)

where C± = S?b ± QO".


The Hartree ground-state energy of the system can be expressed by the
lowest N Hartree eigenenergies Ci, with a composite index i == [n, m, 0"]:

t: = LN
o
. _ e22
c, E
Jd Jd' n(r)n(r')'I '
T T l T-T
(9.20)
i=l

where the subtracted integral is the energy of the direct Coulomb interaction
in the system, counted twice in the summation of Hartree energies Ci. By
introducing the Fermi energy CF, which separates the occupied and unoccu-
pied Hartree energy levels in the ground state, and calculating the interaction
integral, the following equation can be obtained:
37rN2e2
Eo = L 8(cF - c,) c, - 10ER (9.21)
i

In the above 8 stands for the Heaviside function. The Fermi energy is deter-
mined from the condition for the fixed electron number N:
(9.22)

Let us now evaluate the magnitude of the spin-orbit coupling constant


Q. According to (9.2) one needs to estimate the average self-consistent field
(8U) that acts on an electron. The energy of a classical particle moving in a
two-dimensional potential !m* S?b2r2 on the orbit with radius r is

8U = ~m* S?b2r2 + ~S?bl. (9.23)


2 2
By replacing the classical variables with appropriate operators and averaging
8U first in a single quantum state [n, m, 0"],

(nmO"I8UlnmO") = ~nS?b(n + m + 1), (9.24)


and then over all the occupied states,

(8U) = 2~nS?b L 8(cF - cnma) (n + m + 1), (9.25)

the following equation [70] is obtained:

Q = ~fJnS?b2ffi
3 . (9.26)

The details of calculation of the Hartree energy Eo defined by (9.21) are given
in [70], and here we present only the final result:

t: = 97rN 2e2 ~N3/2nS?' J1- fJ2N. (9.27)


o 20ER +3 0 36
9.1 Quantum Dot in the Absence of a Magnetic Field 131

The dot radius R will be determined from the minimum-energy condition


8£oj8R = 0, which leads to the equation
2 _ 31rNe 2
Wo - 4fR3 m *
[1 _100aB
271rR
(1 _ 3N)]6 '
(32 (9.28)

where aB = d'i,2 j(m*e 2 ) is the effective Bohr radius. Under the constraint of
a large number of electrons, the above equation can be solved perturbatively
with respect to the small parameter aBj R « 1. The zeroth-order approxi-
mation, which can be written as
R3 _ 31rNe 2 (9.29)
0 - 4fm*w2'
o
corresponds to the classical result of Shikin [120j. Assuming that the first-
order solution is of the form R = Ro(l + 8), we can obtain the expression for
the correction 8:
8= 100aB
811l"Ro
(1- (32 N ).
36
(9.30)

Neglecting the higher-order corrections (nonlinear in 8), it is possible to find


the effective frequency of the confining potential [2b based on the definition
(9.15):

[2'2
o
= [22
0
(1- (32N)
36'
(9.31 )

where we use the notation


[22 _ 100aB w2
o - 271rRo o· (9.32)

Eventually, we obtain the expression for the ground-state energy in the


Hartree approximation:

£ = 91rN 2e2 ~h[2


o 20fRo + 3 0
(1 _(32N) N 3/ 2.
36
(9.33)

The first term in this equation, given in [120], is the classical energy of N
interacting electrons, confined in a parabolic well:

91rN 2e2 -_ Jd r n ()
r ~2 m * wor
2 2 2
+ e2f Jd r Jd'
r n(r)n(r')
Ir - r' I . (9.34)
20f R 0
The second term, a quantum correction, is composed of the energy of oscil-
lation with frequency [20,

(9.35)

and the energy of the spin-orbit interaction.


132 9. Description of a Many-Electron Quantum Dot
____ 80
><1>
E
'-"

~ 70
to
a..
...
~ 60
>-
...<1>
0)

C
<1> 50
.:£
.8
(J)

"§ 40
Fig. 9.1. Dependence of
the average ground-state
o
'-
0) energy per one electron on
the number of electrons in
15 20 25 30 35 a quantum dot [70]
number of electrons

Now we shall include perturbatively the exchange interaction, which has


so far been neglected in the Hartree approximation. As the first-order correc-
tion with respect to LlE let us calculate the average value of the exchange-
interaction operator G (defined by (9.9)) in the Hartree ground state deter-
mined in the absence of the spin-orbit interaction:

LlE = L 6>(cF - Ci) (iIGili) I (9.36)


, /3=0
As can be shown through a tedious analytical calculation, which is presented
in [70], this correction reads

LlE = -
4V5 (
9v3 1-
3)
4VN
N7/4e 2
fRo (1 - 80 ), (9.37)

where 80 = 8({3 = 0). It has been verified numerically that the higher-order
corrections to the Fock term are negligibly small (decreasing by about an
order of magnitude with the order of perturbation). Thus we have obtained
the total ground-state energy E of the system of N electrons confined in
a parabolic well. The energy E contains the kinetic energy, the direct and
exchange Coulomb interaction energies, and the spin-orbit interaction energy
E = Eo + LlE. (9.38)
Figure 9.1 shows the average ground-state energy per one electron C = E/ N,
drawn as a function of the number of electrons N. The very good agreement
with the results of measurements is clearly visible. Two curves that corre-
spond to the values of parameter {3 = 0.3 and 0.5 lie between the classical
9.2 Quantum Dot in a Magnetic Field 133

result of Shikin et al. [120] and the experimental data reported by Ashoori
et al. [4].
Let us now consider the selection rules for the optical transitions of a
system under far infrared (FIR) radiation. The light from the far infrared
range excites the electron system, so the FIR absorption is an important tool
in experimental studies of quantum dots (see Chap. 5).
As we have already mentioned, the wavelength of light from this range
considerably exceeds the dot dimensions, and in the description of the cou-
pling of the FIR light with a quantum dot the dipole approximation can
be used (see Sect. 3.4). In the first order of perturbation, the probability of
an optical transition between the initial state Ii) and the final state If) is
proportional to the squared matrix element of the interaction,
(9.39)

where E is the electrical field, uniform over the area of the dot. The dipole
matrix element dfi does not vanish if and only if it is possible to choose the
pair of occupied Hartree-Fock states, one in the initial state Ii) and one in the
final state If), with equal spins (T and each of the orbital quantum numbers
[n, m] differing by unity,
(Tf = (Ti, Inf - nil = 1, Imf - mil = 1, (9.40)
and with the remaining occupied Hartree-Fock states being equal in pairs
in the states Ii) and If). In other words, the absorption of the FIR photon
leads to the excitation of a single electron from its (initial) Hartree-Fock
state to another (final) Hartree-Fock state with the same spin and both
orbital quantum numbers changed according to (9.40). In the Fock-Darwin
representation above selection rules take the form
(Tf = (Ti, n~ = n~ ± 1, n~ = n~,
f .
or (Tf = (Ti, n~ = n~ ± 1, n+ = n+, (9.41 )
i.e., the excited electron changes one of its orbital quantum numbers [n+, n_]
by unity. The above selection rules lead to the splitting of resonance energy:

[f _ [i = c± = hDb ± ~a. (9.42)

The magnitude of this splitting a (the splitting is due to the spin-orbit


interaction) depends on the number of electrons, according to (9.26).

9.2 Quantum Dot in a Magnetic Field

The procedure of the minimization of the Hartree energy of a system with


respect to the dot radius R, and further, of the perturbational inclusion of
the exchange interaction, will now be repeated for the case of a quantum dot
134 9. Description of a Many-Electron Quantum Dot

in a magnetic field. The form of the Hartree equations with the inclusion of
a perpendicular magnetic field (9.13) is the following:

ti2 + -m
( ---Ll
2m-
1 * ( n B 1'2 + -we
2
1
4 2
2) 2
1 . - gj..tBa B + a I)
r - -nwel a 'Pi (
nl. ra )

31fNe2 )
= ( ci - 4cR{B) 'lfJi(ra), (9.43)

where We = eBI(m*c) denotes the effective cyclotron frequency, and the


radius in the absence of a magnetic field R, which appears in the definition of
nk (9.15), is replaced by R(B). The effective frequency of the total confining
w;
potential will be now denoted by n12 = n'J + 14, and the corresponding
characteristic length as I = y!til(m*n'). The eigenfunctions of (9.43) have
the same form as given in (9.16) and (9.18), only the characteristic length
changes, in the presence of the field equal to I. The eigenenergies are
, 1 31fNe 2
Cnma = tin (n + 1) - 2nwem - gj..tBaB + ama + 4cR(B) ' (9.44)
or in the Fock-Darwin representation,

cn+n_a = c+ (n+ +~) + (n_ +~) - gj..tBaB +3:~;2,


E:- (9.45)

where C± = nb ± (nw e /2 + aa). Since the Zeeman spin splitting is negligi-


bly small in GaAs (g ~ 0.5 gives gj..tB ~ 0.05 meV IT), which is the most
frequently used material for creating quantum dots, it will be neglected in
further calculations (all comparisons between the theory and experiment that
are made in this chapter are appropriate for dots grown on GaAs substrates).
The inclusion of a magnetic field leads to possible crossings of different
energy levels Ci' Whether a crossing of a pair of close energy levels C1 and
C2 is allowed or forbidden depends on the off-diagonal matrix element of the
operator, which describes the change of the Hamiltonian under an infinitesi-
mal change of the field [91]. The condition for the allowed crossing of levels
is the vanishing of the off-diagonal matrix element:

(9.46)

The operator BHI BB commutates with the spin ((j') and inversion (r -7 -r)
operators. The commutation with the angular-momentum operator requires
the (assumed here) circular symmetry of the confining potential. Thus, (9.46)
is always satisfied for a pair of states with different quantum numbers n and a.
However, it is generally violated for a pair of states that differ only in the
quantum number m. This leads to the forbidden crossings of levels, which
can be taken into account by changing the equation for eigenenergies (by
introducing a modulus in (9.44)):

= tin ,(n + 1) + m \12nwe + aa \+ 4cR(B)'


31fNe 2
Cnma (9.47)
9.2 Quantum Dot in a Magnetic Field 135

A similar change should be made in the definition of a pair of energies c+


and c, which occur in (9.45): c± = nh ± Iliwc /2 + aO'I. It should be noted
that the above modification of the single-particle energy spectrum, which is
a result of the analysis of the possible crossings of the energy levels, is of a
perturbational character, but beyond the Hartree-Fock approximation.
Following the procedure described above for the zero magnetic field, it
is possible to estimate the magnitude of the spin-orbit coupling constant,
which is now a function of the magnetic field [70],

a(B) = ~j3iB'LstkVN, (9.48)

with the renormalizing function

iB= J1+Z2/N(1- ~)I


1+z z=wcl2n~
(9.49)

In weak magnetic fields the function f B approaches unity and decreases for
an increasing field. With good accuracy it is possible to replace nk in the
definition of fB with no, which is defined by (9.32). The Hartree energy of
the system is expressed by equation [70]

£ (B) - 9nN 2e 2 ~N3/2li 2 _ 3nNe2 . / _ j32f~N (9.50)


o - 20€R o + 3 WOUB V
4€m* R(B)3 1 36'

where we use the following notation:

UB = 1 + (wc/2nO)2 _1_ (~ _ 4_Z_) I . (9.51 )


1- z N 1- z) z=f32f~N/36
Similarly as in the absence of the field, the dot radius R(B) is found from
the minimum-energy condition
B£o(B)
BR(B) = 0, (9.52)

which leads to the equation

2 3nNe 2 [ 100aB ( j32f~N)] (9.53)


WoUB = 4€m*R(B)3 1 + 27nR(B) 1- 36 .
In the zeroth-order approximation we obtain

Ro(B)3 = 3nNe 2 = ~R~, (9.54)


4€m*w5uB UB
where Ro is defined by (9.29). The first-order correction 8(B), defined as
R(B) = Ro(B)(l + 8(B)), reads

8(B) = 100aB
8In Ro(B)
(1 _13 f~N)
2
36'
(9.55)
136 9. Description of a Many-Electron Quantum Dot

69
>-
Q)
N=32 >115
., N=32
E E
'--' ~110
Q) 67 N=31 .2
(J

t 15 105
0
0. N=30 0
0..
N=27
!o...
Q)
65 "0 100
0
0.

>-
N=29 ·E.,
OJ
!o...
.<: 95
0
Q)
63 0 2 3 4
c N=28
Q) magnetic field (T)
2
.E
(I) N=27
61
"U
c Fig. 9.2. Dependence of
:::J
0
the average ground-state
!o...
OJ ~=O.O energy per one electron on
59
the magnetic field for f3 =
0 2 3 4 0.0. Inset: chemical poten-
magnetic field (T) tial [70]

As a result, the total energy in the Hartree approximation can be written in


the form
E (B) = 91fN 2e 2 ~ 2/3!ifl ( _ (32 f~N) N3/2 (9.56)
o 20fRo(B) + 3uB 0 1 36 .
In the presence of a magnetic field, the calculation of the correction to the
ground-state energy due to the exchange interaction turns out to be much
more difficult than in the absence of the field. We shall hence make use of a
hypothesis according to which the magnetic field has no effect on the form
of dependence of the exchange energy on the number of particles [130]. The
expression for the exchange energy obtained from this assumption can be
written in the following way:

L'lE(B)
4V5 (
= - 9v'3 1 - 4VN 3) N 7 / 4 e2
fRo (B) (1 - c5 o(B)), (9.57)

where c5o(B) = c5(B; (3 = 0). The total ground-state energy of the system
reads
E(B) = Eo(B) + L1E(B). (9.58)
Figures 9.2-4 present the evolution of the average ground-state energy of a
quantum dot per one electron c(B) = E(B)/N, in a magnetic field (for GaAs,
nwo = 5.4 meV). The three diagrams correspond to the parameter (3 equal to
0.0 (without the spin-orbit interaction), 0.3, and 0.5. The curves that have
been calculated here are very similar to the experimental curves reported by
Ashoori et al. [4], obtained by means of the capacitance spectroscopy (see
the description on page 88 and Fig. 7.7). A comparison of the curves from
9.2 Quantum Dot in a Magnetic Field 137
69
;;:- N=32
~115
> N=32
"
Q)

E
........ E
......., 110
~ 67
u N=31
t
0
C1.
N=30
'-
Q)
65 -0 100
C1. .~
E
>- N=29
Ol
'- "
.r:: 95
Q)
c 63
0
a
Q) magnetic
N=28
.!
0
Vi 61 Fig. 9.3. Dependence of
N=27
u
c the average ground-state
::J
0 energy per one electron on
'-
Ol
59
13=0.3 the magnetic field for f3 =
0.3. Inset: chemical poten-
a 2 3 4 tial [70]
magnetic field (T)

69
;;:-
Q)
~115
>Ql
E E
N=32 ......., 110 N=32
~ 67 .Q
u
t N=31 ] 105
0 0
C1. 0.
'-
Q)
65 N=30 -0 100
C1. .~

>-
E
Ol
'- N=29 .r::"
0
95
a
Q)
63 2 3 4
c
Q) magnetic field (T)
.! N=28
2CIl
61
u
c N=27 Fig. 9.4. Dependence of
::J
0 the average ground-state
'-
Ol
59
13=0.5 energy per one electron on
the magnetic field for f3 =
0 2 3 4 0.5. Inset: chemical pot en-
magnetic field (T) tial [70]
138 9. Description of a Many-Electron Quantum Dot

8
GaAs, N=25
>-
Ol (3=0.3
~
E
>-
~ 6
Ol
C
Ol

C
o
~c 4
o
.;::
co
§ 2 Fig. 9.5. Dependence of the
CIl FIR absorption spectrum of
~
a quantum dot containing
25 electrons on the magnetic
O-r---.---r--.---.---~--.---.---~ field; squares - the experi-
a 2 3 4
ment [34); lines - the model
magnetic field (T) [70]

Figs. 9.2-4, which were calculated for different values of {3, shows that the
inclusion of the spin--orbit interaction allows for a visible improvement of the
model curves toward the experimental data (the agreement is particularly
good for (3 = 0.3).
Let us now proceed to the selection rules for the FIR optical transitions
in a magnetic field. Since the magnetic field does not affect the form of
the Hartree-Fock wave functions, the selection rules are the same as in the
absence of the field (9.40). Hence, the resonance energies have the form

[f(B) - [i(B) = c± = nn' ± I~nwc ± ~a(B)1 (9.59)

and describe four different transitions.


Figure 9.5 presents a comparison of the magnetic-field dependence of the
FIR resonance energies that have been obtained within our model (for the
parameters appropriate for GaAs with nwo
= 7.5 meV) with the experimen-
tal results of Demel et al. [34] (see the description of the experiment on page
54). Assuming that the splitting of modes obtained in the experiment is due
to the spin-orbit interaction, a good agreement of the model curves with the
experimental data is obtained again for {3 = 0.3. In particular, it is possible
to obtain the correct values of (i) the energy of the lower transition in the
absence of the field (~ 2.8 meV) , (ii) the critical magnetic field at which
the anticrossing of levels (~ 1 T) takes place, and (iii) the resonance energy
at the point of this anticrossing (~ 3.9 meV). It is absolutely crucial that
a good agreement between a number of calculated and measured quantities
is obtained using only one fitting parameter {3. The small repulsion of anti-
crossing levels observed in the experiment is most probably caused by a slight
anisotropy of the confining potential.
9.2 Quantum Dot in a Magnetic Field 139

It should be stressed that the inclusion of the spin-orbit interaction ex-


plains various measurable details of the experimental data, all leading to the
same value of the fitting parameter /3, where for GaAs dots /3 = 0.3. This is a
realistic value, as for ordinary atoms with atomic numbers ~ 40 this parame-
ter is /3 ~ 0.1. It should be mentioned that attempts to include the spin-orbit
interaction within the band-structure model have already been reported, but
only for up to two confined electrons [31]. The numerical results that were
described in [31] show a similar tendency and predict the splitting of the FIR
spectruni and the forbidden crossing of levels, which confirms the validity of
the theory presented here.
10. Description of an Exciton
in a Quantum Dot
Within the Effective-Mass Approximation

In the following description a quantum dot is modeled as a local perturba-


tion of the periodic crystal field of a semiconducting structure that surrounds
the dot. In the description of such a system, the effective-mass approximation
can be applied if the perturbation potential varies slowly over the interatomic
distance. In the case of quantum dots with sizes a few tens of nanometers,
this assumption is expected to be satisfied. This also seems doubtless the case
for the dots created by means of interlayer diffusion or by the application of
a modulated electric field. It is probably a good approach for lens-shaped
self-assembled dots. However, the cases of dots in the form of compact bub-
bles embedded in another semiconductor are disputable, since their sharply
defined geometrical shapes may lead to a more rapid jump of the· potential
at the interface.
The interaction of the quantum dot modeled as a local perturbation of
the periodic crystal structure with charged carriers is always an electric in-
teraction, and hence the bare potentials of the dot-electron and dot-hole
interaction must differ in sign due to opposite electric charges of the two
types of carriers. Therefore, an empty dot cannot at the same time bind an
electron and a hole, as is assumed in the band-structure model of a quan-
tum dot for the appropriate pair of materials, as discussed in Chap. 8. This
contradiction might be explained in the possible binding of a hole by the
quantum dot in which an electron is already bound (despite the fact that an
empty dot will repel a hole). This leads to the possible binding of an exciton
(electron-hole pair) in the dot [68, 71], which can be obtained from first prin-
ciples. Binding of an electron and a hole in a quantum dot turns out to be
a very subtle effect, which is highly sensitive to the geometrical sizes of the
structure and magnetic field. This effect seems fully to explain the whole se-
quence of experimental results obtained in the photoluminescence studies of
quantum dots. They could not be described within the band-structure model
(on the other hand, the band-structure model describes better the Coulomb
interactions, due to the numerical character of the calculation). Moreover, it
should be stressed that the model of the quantum dot as a perturbation of
a crystal field described in the effective-mass approximation has a character
independent of the type of structure. It is therefore expected to lead to the
common features of quantum dots of different types. The photoluminescence
142 10. Description of an Exciton in a Quantum Dot

properties of quantum dots described below have a universal character and


seem to be confirmed well by the experimental data obtained for different
types of dots.
Without defining the type of quantum dot (i.e., deciding on the simplified
but universal model) we assume that the potential of an empty quantum dot
can be modeled in the form of a Gaussian well:

V(r) = ±Vo exp [- ~:] . (10.1)

We shall assume that the sign "-" (attraction) refers to an electron and
the sign "+" (repulsion) to a hole. It can certainly be reversed, but with no
effect on the following considerations. The model potential (10.1) is smooth
and has a finite depth (Vo) and a characteristic finite radius L. Denoting the
positions T, momenta p, projections of angular momenta onto direction of the
field I = -i(x8y - y8x ), effective masses m*, and the cyclotron frequencies
We = eB / (em *) of the pair of particles with indices e (electron) and h (hole),
the Hamiltonian of the system can be written in the form
p; 1
H 2m; + v,,(re) + '2tiweele
p~ 1 e2
+-2
* + Vh(rh)
mh
- -2tiwchlh - I
E Te - Th
I + Eg , (10.2)

where Eg is the band gap of the semiconductor in which the dot is created.
v" and Vh are the bare single-particle potentials of an electron and a hole in
a magnetic field (the sign "-" refers to the electron and "+" to the hole):

i = e,h. (10.3)

The last term in the above expression describes the confining potential due
to the magnetic field. It should be observed that independently of the charge
of a carrier, the magnetic field enhances its binding in the dot.
The eigenfunctions of the Hamiltonian (10.2) are postulated in the Hartree
form
(10.4)
In such an approximation the equation for the eigenstates of (10.2) is equiv-
alent to the following system of equations (sign "-" refers to the electron,
and "+" to the hole):
p; + Ui(ri) ± 1)
( 2m* '2 tiwci 1i (MTi) = Ei(/Ji(Ti) , i = e, h, (10.5)

where the self-consistent effective field that acts on the electron Ue is given
by the formula

Ue(re) = v,,(re) - -
e2
E
J
dTh
I(Ph(Th)1 2
ITe - Th
I' (10.6)
10.1 Exciton in the Absence of a Magnetic Field 143

and the self-consistent effective field that acts on the hole Uh is obtained by
changing of indices in the above formula: e into h and vice versa. The total
energy of the pair E is given by

E-Eg = Ee+jdrh<Pi;(rh) ( p~* +Vh(rh)-~nwchlh)<ph(rh)


2mh 2

+Eh + j dre<p:(re) (2~~ + v,,(re) + ~nwcele) <Pe(re).


(10.7)
The system of equations (10.5) will be solved using the method of consecutive
approximations in an analytical way. It is worth considering separately the
two cases, both with and without the magnetic field.

10.1 Exciton in the Absence of a Magnetic Field


In the zeroth-order approximation we neglect in (10.5) the term that de-
scribes the electron-hole interaction. The potential of the quantum dot is
approximated by a harmonic well (the first two terms of the power-series
expansion of V). The self-consistent effective field that acts on the electron
assumes then the form
U(O) ( ) -
e r - _ V;
0 + ~2 mewOe
* 2 r2, (10.8)

where the frequency WOe is given by the equation


2 L2
V;
0= "21 me*WOe . (10.9)

It is useful to introduce the following length scale,

AOe =
~
---,
m~WOe

and the dimensionless parameter


(10.10)

nwOe
K= Vo . (10.11)

The ground-state energy and wave function of the electron in the zeroth-order
approximation can be written in the form
E~O) nwoe(1- K- I ) (10.12)

<p~O)(r) _1_ exp [-~] (10.13)


-j1T AOe 2A5e .
For fixed curvature of the potential WOe, the length scale AOe and the ratio
Vol L2 are constant. The parameter K determines the depth of the potential
well Vo and, related to it, the dot radius L. By changing K it is possible to
144 10. Description of an Exciton in a Quantum Dot

8
o'"
>E
..r::.
~

r lADe

Fig. 10.1. Hole effective self-consistent potential as a function of parameter K


(liwo e = 7.5meV). Characteristic structure of a double well in variable r is visible

model the changes of the quantum dot dimensions: an increase of K corre-


sponds to a decrease of the dot dimensions. However, in the zeroth approxi-
mation the electron is not bound in the dot if E£O) > O. We should therefore
confine our consideration to the interval
(10.14)
The state of the hole in the zeroth-order approximation can be calculated by
solving (10.5) with the distribution of the electron charge given by (10.13).
After integration of the interaction element we obtain

p~
( 2m* + u.(0)) ¢(O)(r) = E(O)¢(O\r) (10.15)
h h h h '
h
where

U~O)(r) = Vh(r) - 2
v;e exp [-
f Oe
2~ Oe
] [2:: ],
10
Oe
(10.16)

and 10 is the Bessel function. The characteristic profile of the effective self-
consistent potential of the hole is presented in Fig. 10.1. This potential has
two minima: at r = 0 and r = r3h (see Fig. 10.2). Their positions and the
extreme values of the potential depend on the value of the parameter K. With
an increase of K (i.e., with a decrease of dot diameter) the second minimum
moves toward the dot center, and both minimum values of the potential attain
equal values. As a result, for K ~ 0.8 we can observe only one minimum of the
potential. Thus, in the zeroth approximation for K < 0.8, the self-consistent
effective field that acts on the hole is a double potential well.
When the potential structure is of the double-well type, the usual pro-
cedure is to search for the states of a particle in both wells simultaneously,
and then a superposition of these states is made. Near the dot center the
potential curvature is described by the frequency Wlh:
10.1 Exciton in the Absence of a Magnetic Field 145

0.4

2h
0.2

0.0 -+-+--.,.-\-----,---,---r------, Fig. 10.2. Hole effec-


tive self-consistent po-
tential for Ii 0.5
-0.2 (nwo e = 7.5meV); two
minima (lh) and (3h) are
3h separated by a maximum
-0.4 (2h)

mhW~h = d;~h I . (10.17)


r=O
The potential here is approximately parabolic:

Uh(r) ~ U1h + ~mhW~hr2, (10.18)

and hence, by solving the equation

(2~h + Uh(r) ) ~lh ( r) = Elh ~lh (r) (10.19)

we can obtain the ground-state energy and wave function of the hole in the
"left" well in the form
(10.20)

~lh Alh exp [- ~r ;;J ' (10.21)

where O"r = mhw~h/(m;w5e)' and Alh = O"I/(y'7fAOe)'


The analogous procedure is used for the "right" potential well. Its curva-
ture is denoted by W3h,

* 2 d2Uh
mh w3h = dr2
I (10.22)
r=r3h

and the potential can be approximated by a parabola:

Uh(r) ~ U3h + ~mhw~h(r - r3h)2. (10.23)

Thus, the ground-state energy and wave function of the hole in this well are
obtained by solving the equation:

(2~h + Uh(r)) ~3h(r) = E3h~3h(r). (10.24)

As a result we obtain
146 10. Description of an Exciton in a Quantum Dot

(10.25)

(10.26)

where a~ = m;;w~h/(m;w5e)' The function ih introduced in (10.26) allows


for the smooth connection of functions 'l/Jlh and 'l/J3h. The analytical form of
function ih and of the normalization constant A3h were determined in [71]
by substituting function 'l/J3h into (10.24).
The state of the hole in a double well is postulated in the form
(10.27)
The coefficients Cl and C3 can be found using the condition of the minimum
hole ground-state energy E h :

Eh = J dr¢;;(r) [2~;; + Uh(r)] (fJh(r). (10.28)

Hence, we calculate the variation of (10.28) with respect to the coefficients


Cl and C3, and then vanish it. It is useful to introduce here a new pair of
coefficients CL and CR:
1
Cl -1--2 (CL - CRPh) , (10.29)
- Ph
1
C3 -1--2 (CR - CLPh), (10.30)
- Ph
where

Ph = J
dr¢lh(r)¢3h(r) (10.31 )

is the overlap integral of the pair of functions in the two potential wells. The
following system of equations is obtained for the new coefficients:

{ CL(ELh - Eh) + cRhR =0 (10.32)


cLIRL + cR(ERh - E h ) =0
where
E p~bLh
ELh ~ Ih - -1--2 '
- Ph
E p~bRh
ERh ~
3h - -1--2 '
- Ph
PhbLh
hR ~
-1--2 '
- Ph
PhbRh
hL (10.33)
12'
~
- Ph
and
10.1 Exciton in the Absence of a Magnetic Field 147

U 0"32 ) U * w1h
m h 2 ( r3h 0"32 ) 2
h ( r3h 2 2 - 1h - * 2 - 2 2
0"1 + 0"3 2mewOe AOe 0"1 + 0"3
U ( 0"32 ) U * w3h
m h 2 ( r3h 0"32 ) 2
(10.34)
h r3h 2 2 - 3h - * 2 - 2 2
0"1 + 0"3 2mewOe AOe 0"1 + 0"3
The condition for the existence of nonzero solutions of the system of equations
(10.32) is the vanishing of its determinant,

(10.35)

In further considerations we can use the relation ELh > E Rh . This assumption
is justified by the fact that the left well Uh (r) of the potential is always more
shallow than the right one [71]. Therefore, the solutions of (10.35) have the
form
Eh1 = ELh + Ih, Eh2 = ERh -Ih, (10.36)
where

Ih = ELh - ERh ( 1+ 4p~bLhbRh _ 1) . (10.37)


2 (1 - p~)(ELh - ERh)2
Using the energies calculated above, we can obtain two sets of solutions of
the system (10.32). The problem of the motion of the hole in the double-well
potential U~O) has a pair of the following solutions, the first one with energy
and wave function
ELI. + Ih, (10.38)
cll1/'J1h(r) + C311/'J3h(r) , (10.39)
where
1 1 + BLhP~
Cll
y'1 - p~ y'1 + 2BLhP~ + BlhP~ ,
1 Ph(1 + B Lh )
y'1 - p~ y'1 + 2BLhP~ + BlhP~ ,
Ih(1-p~)
(10.40)
p~bLh
and the second one
ERh -Ih, . (10.41)
c121/'J1h(r) + c321/'J3h(r), (10.42)
where
148 10. Description of an Exciton in a Quantum Dot

4 (a) 4 (b)

-
~3 ~3
.r:::.
~2 W
2
<:I
1 1

0 0
0 0.4 0.8 0 0.4 0.8
K K
Fig. 10.3. Dependence of (a) energies Ehl and Eh2, (b) difference between them
L1E, on parameter '"

1 1 + BLhPf,
VI - Pf, VI + 2BRh pf, + B~hPf, ,
!'h(1 - Pf,)
(10.43)
Pf,bRh
An important property of the pair of states (10.39) and (10.42) is that they are
of the same axial symmetry. The investigation of the dependence of energies
Ehl and Eh2 on the size of the quantum dot, shown in Fig. 10.3, turns out
to be very important. It appears that in the whole range of the parameter
'" energy Ehl is positive, while energy Eh2 is negative. This means that in
the state described by the wave function ¢hl the hole is not localized in the
quantum dot. On the other hand, from the analysis of coefficients C12 and C32
it follows that for any value of Ii the relation IC121 < IC321 is fulfilled. Hence,
in the state ¢h2 the hole resides near the second minimum of the potential
U~O) .
Let us now calculate the next correction to the electron state. Taking
into account the interaction of the electron with the hole, the effective self-
consistent potential that acts on the electron can be written as

Ur (i)( )
e
= -Vr
0 exp
[_~]
L2
_ e2
E
Jd r I l¢hi(r')jZ
Ir _ r'l ' (10.44)

where index i = I, 2 refers to the two states of the hole (in a zero magnetic
field only the case of the bound state of the hole is significant, i.e., i = 2).
The explicit form of the above equation is the following:
[K,r
2 e 2 y'7f
11;.;.;0 { - -I exp - - - ] - --:--:---
e K, 2A5e EAoe11;.;.;oe
10.1 Exciton in the Absence of a Magnetic Field 149

Fig. 10.4. Electron effective self-consistent potential as a function of parameter '"


(nwo e = 7.5 meV)

J
+ 2Cil Ci3Ph err ; cr§ exp [- cr~~6~§ (r - r3h err ~cr§ ) 2]
xIo [cr~~6~§ (r - r3h err ~ cr§ r]) }. (10.45)

The potential U~2) for i = 2 will be denoted simply by Ue . The profile of this
potential, in dependence on the dot size, is shown in Fig. 10.4. As can be
seen, the double-well structure in the variable r is more distinct here than
in the case of the hole. The existence of the first minimum (in the center of
the dot) is due to the minimum in the confining potential V. The existence
of the second minimum is due to the attractive potential of the hole. The
relative depths of the pair of minima of potential Ue depend on the dot size,
parametrized by /'£. Both situations are possible: the two minima can have
comparable depths, or one minimum can be clearly deeper than the other.
The first case is illustrated in Fig. 10.5, where we present a characteristic
profile of the potential Ue for /'£ = 0.5.
The analysis of the electron motion in the obtained potential is carried
out in an analogous way as for the hole. In the vicinity of the minima rle = 0
and r3e, the double-well structure of the potential Ue is approximated by a
pair of parabolas,

(10.46)

(10.47)
150 10. Description of an Exciton in a Quantum Dot

r/AOe
o -r----r----r----r----r---,----,---~--~

3e

Fig. 10.5. Electron effective self-consistent potential for Ii. = 0.5 (nwo e = 7.5meV)
two minima (Ie) and (3e) are separated by a maximum (2e)

The frequencies that determine the curvatures of the parabolas are given by
the equations

(10.48)

For each of the approximated potentials we can now solve the equation for
the ground-state energy and wave function

(10.49)

For the potential (10.46) the solution has the form


E le Uel + nwl e1 (10.50)

~le(r) = Aleexp [-~ (~~:)] , (10.51 )

where 8r = Wle/WOe and Ale = 81/ ,fifAOe. For the potential (10.47) we obtain
(10.52)

(10.53)

where 8§ = W3e/WOe. The function Ie in the above expression plays the same
role as the function Ih for the hole. It ensures a smooth connection between
~le and ~3e. Its analytical form is given in [71]. This leads to the following
form of coefficient A3e:

83e exp
.,fiKAoe
[Ie (s:UAOe
T3e)]
3-

X {"21 exp [- (83 A6e ) 2] + 83 AOe""'2r:;;: [ 1 + p ( 83 AOe )] } - !


T3e T3e y 7f T3e

(10.54)
10.1 Exciton in the Absence of a Magnetic Field 151

The function P in the above stands for the usual error integral:

p(x) =
2
,;:rr r e-
io Y
2
dy. (10.55)

Now it is possible to find the ground-state wave function of the electron in


the double-well potential. The following form of this function is assumed:
(10.56)
The procedure of finding the coefficients d l and d3 from the condition of the
minimum electron energy is the same as described previously in the case of
the hole. We will therefore confine ourselves to the presentation of the final
results. The two solutions have the form
ELe - "(e, (10.57)
dn(/>!e(r) + d3l ¢3e(r), (10.58)
where
1 1 + BLeP~
dn
\11 - }1 + BleP~ + 2BLeP~ ,
P~
1 Pe (1 + B Le )
}1 - P~ }1 + BleP~ + 2BLeP~ ,
"(e (1 - p~)
(10.59)
p~bLe
and
ERe +"(e, (10.60)
d l2 ¢le(r) + d 32¢3e(r), (10.61 )
where
1 Pe (1 + BRe)
}1 - p~ }1 + B~eP~ + 2BReP~ ,
1 1 + BReP~
}1 - p~ }1 + B~eP~ + 2BReP~ ,
"(e (1 - p~)
(10.62)
p~bRe
All symbols used in the above equations have an analogous meaning corre-
sponding to the symbols that refer to the hole.
A crucial problem here is the dependence of energies Eel and Ee2 on the
dot size (Fig. 10.6). With a decrease of the dot diameter (i.e., with an increase
of 1£), Eel increases and Ee2 decreases. However, for 1£ ;:::; 0.7 this tendency
reverses and the difference between the two energies begins to increase again.
Additionally, from analysis of the coefficients d it follows that for any value of
1£ the relations Idnl > Id3l 1 and Idd < Id32 are fulfilled [71]. This means that
1
152 10. Description of an Exciton in a Quantum Dot

o (a) (b)
Q)
~-1
8 ~-
.r:= .r:=
-r.,-2 LiJ·2
W -<:l
-3

4+---.----.---.---.
o 0.4 0.8 o 0.4 0.8
K K.
Fig. 10.6. Dependence of (a) energies Eel and E e2 ) (b) difference between them
L1E) on parameter K,

the electron that occupies its ground state described by (10.58) is localized
closer to the dot center, and the electron in the excited state (10.61) resides
closer to the second minimum of the potential Ue . These states have the same
axial symmetry, which results from the fact that with respect to the radius
the potential Ue has the double-well structure.
When we know the electron and hole states it is possible to construct
the states of an electron-hole pair that is confined in a quantum dot. They
assume the form
Pi(rh, re) = ¢h2(rh)¢ei(r e ), i = 1,2. (10.63)
These states correspond to the energies

Ei Jdre JdrhP:(rh,Te) [-2~;Lle+v;,(re)


2
!i,2
--2
mh
* Llh + Vh(rh) - ITe e-
E Th
I] Pi(Th, Te)
Eei + Eh2 + LlEi ) (10.64)
where [71]

+ 2Cl2C32 exp [- -21 (:3h 2


"'-Oe (}l
()~+ (}3
2) 2]

X [ (r3h
I O 1
-- 22 ) ()~ 2] } . (10.65)
2 AOe (}l (}3 +
Due to the fact that the hole in the state with energy Eh2 is localized near
the dot and the localized electron can assume a pair of energies Eel and E e2 )
the energies of the ground and excited states of the exciton have the form
10.1 Exciton in the Absence of a Magnetic Field 153

Eel + Eh2 + f1E2 ,


Ee2 + Eh2 + f1E2 . (10.66)
The difference between the energies
f1E12 = IEe2 - Eel I (10.67)
can be interpreted as the distance between the peaks in the photolumines-
cence spectrum of a quantum dot. The following table presents the ener-
gies of the ground and excited states, and the difference between them, for
three chosen values of the parameter /'i, (the energies are given in units of
nwOe = 7.5 meV).

/'i, = 0.55 /'i, = 0.68 /'i,= 0.7

El -0.897 -0.218 -0.112


E2 -0.461 -0.111 -0.246
E2 -El 0.436 0.107 0.124

The fundamental property of the pair of exciton states, the ground and first
excited state, is that they have identical axial symmetry, which leads to the
vanishing of the dipole matrix element between them:

d 12 = J J dTe dTh c]);(Te, Th)( -Te + Th)c])l(Te , Th) = O. (10.68)

This is because in both states both the electron and the hole have zero orbital
angular momentum. This means that the radiative dipole transition between
this pair of states (by means of absorption or emission of an FIR photon)
is forbidden. The blocking of this transition should have a striking effect
on the photoluminescence spectrum of a quantum dot. At low temperatures
and in small dots the relaxation of a bound exciton through the interaction
with phonons is considerably weakened (see Sect. 6.2). Hence, the additional
blocking of the relaxation through the emission of an FIR photon leads to the
occurrence of a metastable excited state in the exciton spectrum. It has to
manifest itself by the splitting of the photoluminescence peak for a quantum
dot. A strong additional peak, which appears in the photoluminescence spec-
trum, corresponds to the radiative recombination from the metastable state.
As illustrated in Fig. 10.6, the difference of energies f1E12 (i.e., the distance
between the ground peak and the additional peak in the photoluminescence
spectrum) is strongly dependent on the parameter /'i, and reaches its mini-
mum at /'i, ~ 0.7, when the two minima of the potential Ue have a similar
depth. An increase or decrease of the parameter /'i, (i.e., the introduction of a
difference between the depths of the two potential minima Ue ) leads to the
increase of f1E12 (see Fig. 10.7 and the above table).
154 10. Description of an Exciton in a Quantum Dot

11: small, L large

E, I:::=,L-..;'------

(a)

11: middle, L middle


I.

E,I-----,----!---
Etf-::;,.~=>-,,=,,'-=--

(b)

11: large, L small Fig. 10.7. Electron effective


Us self-consistent potentials UP)
E,I---":~---!-- (left), and approximated pho-
toluminescence spectra (right),
E,1-----"cc:7-=-- for three values of K (i.e., three
dot sizes): (a) large dot, K =
0.4, (b) medium dot, K = 0.7,
(e) and (c) small dot, K = 0.9

Some imagination about the ratio of intensities of the recombination pro-


cesses (recombination rates) from the states <PI and <P2 at the temperature
T can be found from the following evaluation:
12
-~exp
[- E2 - EI] . (10.69)
h kBT
Using the above formula for T = 5K and nwOe = 7.5 meV we obtain
5 x 10- 4 if /'i, = 0.55,
h {
h ~ 0.155 ~f /'i, = 0.68, (10.70)
0.025 If /'i, = 0.8.
Thus, we observe a strong enhancement of the second peak for /'i, ~ 0.68. This
effect is clearly visible in Fig. 10.8a. Similarly, the ratio of widths of the pair
of peaks TIl and T;l can be roughly evaluated as
78.15 if = 0.55,
/'i,

T;~ ~
TI
:¢e2
¢el t3h~ :: =
T3h
{ 2.413 if /'i, = 0.68, (10.71)
10.625 if /'i, = 0.8.
Hence, an increase in the intensity of the second peak is accompanied by a
decrease in its broadening. A more rigorous analysis, using the electron-hole
10.1 Exciton in the Absence of a Magnetic Field 155

0.20 (a) 8 (b)

V
0.15 e
"';" ....
N
::::::: 0.10 .,t4
-: ";"pN

0.05 2

0.00 0
0 0.4 0.8 0 0.4 0.8
K K
Fig. 10.S. Dependence of (a) ratio of recombination intensities for the recombina-
tion from the ground and metastable states of an exciton, at a temperature of 5 K
and nwOe = 7.5 meV, and (b) ratio of widths of corresponding photoluminescence
peaks, on parameter K

;: 1.76
Q)
...., 1.75 10 ~
> E
(!) 1.74
Q::!
w 5(!)
z 1.73 Z
w
l=
~ 1.72 O~
(!) en
o 1.71 ::.::
Z -0(
Fig. 10.9. Splitting of
:i 1.70
w
a..
the photoluminescence
1.69 peaks and the main-
0 200 400 600 800 1000 peak energy as a func-
DOT SIZE w (nm) tion of the dot size [20]

overlap formula I ~ I J ¢:(r)ch(r)drI 2 (see Fig. 10.8) allows us to state that


for Ii ~ 0.7 we observe a minimal broadening of the second peak and the
maximal intensity compared to the main peak.
Since the parameter Ii is directly related to the dot radius, the charac-
teristic evolution of the photoluminescence spectrum in the function of this
parameter, i.e., the occurrence of a pair of peaks for Ii ~ 0.7, and the decay of
the second peak for both K> 0.7 and Ii < 0.7, seems to be a general property
of quantum dots with dimensions close to a certain critical size (depend-
ing on the curvature of the confining potential, effective masses, etc.). Many
experimental works have already been published where this effect could be
clearly observed. Brunner et al. [20] and Bockelmann et al. [ll] observed the
maximum splitting of the photoluminescence peak for dots with a diameter
on the order of 450 nm (see Figs. 6.3, 10.9, and 10.10). The effect was clearly
156 10. Description of an Exciton in a Quantum Dot

. . Dot Barrier
S . ..... ,. .
Interdlffused Riljeg.,on
..
Interdiffusion
"§ 1000A Sh N4 ." " " " I

.e..!!.
10M GaAs
2~g~ ~-:s
200 A AIGaAs
.
>-
1-
(x=0.35) Quantum Dot VB

t..
E ~
~ ~B
W w=
I-
~
W
U 400nm
Z
W
U
CI)
W
z
:E
~
...J 450nm
~
o Fig. 10.10. Photoluminescence spec-
J:
D..
500nm tra of quantum dots with different di-
ameters; the strongest splitting is ob-
1690 1700 1710 1720 1730 served for a dot with diameter 450 nm
ENERGY (meV) [11]

~
I::
::J
.ci
p Exc .==
~
~
iii
z
W
I-
~
W
U
Z 1. 25 1lW
W
U
If)
w
z
:iE
~ 125nW
...J

~
J: 125nW
D.. Fig. 10.11. Photoluminescence spec-
tra of a quantum dot with diameter
1.68 1.70 1.72 1.7.<1 1.76 450 nm for various excitation powers
ENERGY (eV) [20]
10.2 Exciton in a Magnetic Field 157

9 8 7
+ + +

PLE

InGaAs OD's

~
Lz confinement only: 5ML

850 875 900 925 950 975 1000 1025


Wavelength (nm)
Fig. 10.12. Photoluminescence spectra of self-assembled quantum dots of different
diameters (with the ~ 10% distribution around the mean value) obtained in the
selective-excitation experiment [39]

visible, independently of the laser excitation power (see Fig. 10.11). The fact
that the splitting remained unchanged even at small excitation powers proves
the occurrence of the metastable exciton states in a dot. Moreover, it is the
experimental confirmation of the result obtained in a theoretical way. A sim-
ilar structure of the photoluminescence spectrum of self-assembled dots was
by observed Fafard et al. [39] (see Fig. 10.12).

10.2 Exciton in a Magnetic Field

The application of a magnetic field leads to the enhancement of the localiza-


tion of both the electron and the hole in a quantum dot, since the cyclotron
motions of both carriers differ only in helicity. In particular, the magnetic
field modifies the potential that binds the hole in the dot occupied by the
electron. At not too high fields, the double-well structure of the potential
binding the hole turns out to be distinct enough for the occurrence of a pair
of bound hole states with the same axial symmetry. This enriches the number
of possible metastable exciton states. At higher fields the magnetic localiza-
tion is so strong that the double-well structure for the hole is removed, and
the respective metastable exciton state disappears.
The consecutive steps of the perturbational calculation, and the corre-
sponding subsequent approximations of the states <Pe and (Ph, are determined
in an analogous way as without a magnetic field. The potential of an elec-
tron depends now also on the field intensity, which can be described by a
dimensionless parameter
158 10. Description of an Exciton in a Quantum Dot

(J= (10.72)

where WOe is determined by (10.9). The inclusion of a magnetic field leads to


a redefinition of the parameters of parabolic potential U~O), i.e., the charac-
teristic frequency
WOe -+ we(B) = (JWOe, (10.73)
the characteristic length
AOe
AOe -+ Ae (B) = V1J' (10.74)

and the dimensionless parameter that determines the dot size (and hence the
number of bound electron states)
r;, -+ r;,(B) = (Jr;,. (10.75)
The electron energy and wave function in the zeroth-order approximation
can be calculated by solving the equation

p~ - Vr0
[ 2m~ + ~m*w2r2
2 e e
~(O)(r e ) = E(O)~(O)(r
+ mv2cel ] '+'e Z e '+'e e
). (10.76)

As a result, we obtain
E~O) mve((J - r;,-l), (10.77)

¢iO)(r) J~~ exp [-~~;] . (10.78)

In the zeroth-order approximation, the self-consistent potential of the hole


reads (see (10.16))

(10.79)

where Io is the Bessel function. Due to the occurrence of the Landau term
~mhw~r2 (see (10.3)), the profile of potential Uh strongly depends on the
magnetic field. Figure 10.13 presents a sample curve that is determined for
r;, = 0.5, mvOe = 7.5 meV, and B = 2 T.
As described previously for the zero field, at weak magnetic fields the
potential Uh has two minima: one for r3h and the other in the center of the
dot. For the intermediate values of the magnetic field (2-4 T) both minima
have similar depths. At a strong field, however, the Landau term dominates,
and the potential Uh has only a single minimum, in the center' of the dot.
Since at weak and medium fields we are dealing with a double-well potential,
the search for the energy and wave functions of the hole in the zeroth-order
approximation is identical to the case of a zero field. It shows, however, (see
Fig. 10.14) that both energies Ehl and Eh2 are positive. But due to the fact
10.2 Exciton in a Magnetic Field 159

1.0
______ ~:L _______ _

0.5

0.0 -+--I--~.,.__=t'~-_r_-_, Fig. 10.13. Hole effective self-


consistent potential for magnetic
2 4 6 8
field B = 2 T and K, = 0.5 (liw oe =
-0.5 7.5 meV)

2.0 (a) 2.0 (b)

-
CD 1.5 CD 1.5
8
.c. 1.0
8
.c. 1.0
;:a= W

~
0.5 <1 0.5

0.0 0.0
0 0.4 0.8 0 0.4 0.8
K K
Fig. 10.14. Dependence of (a) energies Ehl and Eh2, (b) difference between them
,dE, on parameter K,; magnetic field B = 2 T

that in a magnetic field Uh - t 00 for r - t 00, the hole can be localized near
the dot also for Eh > O. In addition, the analysis of the coefficients c shows
that the hole with energy Ehl resides closer to the dot center and the hole
with energy Eh2 closer to the second minimum of the potential.
The increase of a magnetic field drives the hole spectrum through the
three distinct regimes. In weak fields (B =0-4 T) it is possible to bind the
hole in the second minimum of the potential and in the dot center. However,
the probability that the second case will occur is limited (the energy Ehl is
much higher than Eh2; see Fig. 10.14). For intermediate values of the field
(B =4-6 T) a pair of states with similar energies occur in the spectrum that
correspond to the hole localized in the center of the dot and in the second
minimum of the potential. This pair of states have the same zero orbital
angular momentum (s-type states). Finally, at strong fields (B >6 T) the
hole is bound in the dot center. It should be stressed that the transitions
between consecutive stages are continuous and consist in the transformation
of one potential minimum into two potential minima, or in the meeting of
the two minima in the dot center.
If the hole is localized in only one potential minimum Uh , then the first-
order approximation of the electron state is identical, as in the case of a zero
field. This is out of the question as far as the weak fields are concerned. For
intermediate fields, however, we are dealing with two bound hole states and
160 10. Description of an Exciton in a Quantum Dot

r/AOe
2~----~--~----~--~--~----~--~--~

2 4 6 8

-6
Fig. 10.15. Electron effective self-consistent potentials U~l) and U~2), for magnetic
field B = 2 T and K. = 0.4 (nwo e = 7.5 meV)

o (a) 1.50 (b)

1.25
8 -2

r
CI) CI)

~
8 1.00
~

~-4 W
-<3 0.75

-6 0.50
0 0.4 0.8 0 0.4 0.8
K K
Fig. 10.16. Dependence of (a) energies E~~) and E~i), (b) difference between them
iJ.E, on parameter K.; magnetic field B = 2 T

thus with two effective self-consistent electron potentials - in (10.45) the index
i can assume the values 1 or 2. The profiles of electron potentials U~l) and
U~2) for K. = 0.4 and B = 2 T are presented in Fig. 10.15. A rigorous analysis
of the model shows that for intermediate fields the potential U~l) has always
only one minimum - in the dot center. Similar behavior is observed for the
potential U~2), but for some values of the magnetic field and of parameter K.
(e.g., B = 4 T and K. = 0.4, or B = 6 T and K. = 0.33), the second minimum
appears at T = T3e and disappears with a decrease or increase of K.. The
transition of the potential UP) from a single well into a double well and vice
versa is continuous.
Let us first consider the case, where each of the potentials (hole and
electron) has a single minimum - in the dot center. The electron state <p~i)
corresponds to the hole state <PhI, and the electron state <p~;) corresponds
10.2 Exciton in a Magnetic Field 161

0 (a) 4 (b)

;;
e(2)

~
•2
8-
.3

..c:
8
..c:2
i:Lr-4 e1 W
-<l
-6 0
0 0.4 0.8 0 0.4 0.8
1C K
Fig. 10.17. Dependence of (a) energies E~~) and E~;), (b) difference between them
dE, on parameter K; magnetic field B = 4 T

Fig. 10.18. Dependence of the hole effective self-consistent potential Uh on param-


eter K; magnetic field (a) 2 T, (b) 4 T, (c) 6 T, and (d) 8 T

to the hole state ¢h2. The energies of these states (see Fig. 10.16) and the
explicit forms of corresponding wave functions can be calculated through the
procedure described in Sect. 10.1, but the calculation is even easier here due
to a simpler structure of the potential. When the potential U~2) has two
minima, two electron states ¢~i) and ¢i;) correspond to the hole state ¢h2.
The energies that correspond to these states are presented in Fig. 10.17. In
that case we have to deal with three electron states (see Figs. 10.18 and
10.19). When the magnetic field becomes too strong (B ~ 6 T), the double-
well structure is washed out, and there are no excited states with the same
axial symmetry as the ground state.
162 10. Description of an Exciton in a Quantum Dot

.... 1

B=8T

no corresponding
.... 1 hole state (h2) for
O.2<K<O.7
c c·
1(
c c ·
(e) (d) ""Y

Fig. 10.19. Dependence of the electron effective self-consistent potential uF) on


parameter I'i, magnetic fields (a) 2 T, (b) 4 T, (c) 6 T, and (d) 8 T

Hence, depending on the magnetic field we obtain the following types of


exciton states with equal (zero) orbital angular momenta:
- weak field: ¢h2¢ii) and ¢h2¢ii) (behavior similar to the case of a zero field);
- intermediate field, small or large r;,: ¢hl ¢ii) and ¢h2¢ii);
- intermediate field, intermediate r;,: ¢hl ¢~i), ¢h2¢~i), and ¢h2¢~;);
- strong field: ¢hl¢~i) (only the ground state).
As described for the case without a field, at the suppression of relaxation
through the interaction with phonons, the excited states with the same ax-
ial symmetry as the ground state are metastable. So, the rearrangement of
the exciton spectrum in the increasing magnetic field is accompanied by the
appearance and disappearance of peaks in the photoluminescence spectrum.
Beginning with a pair of peaks in a zero field, the photoluminescence spec-
trum develops into a system of three peaks, then returns to a pair of peaks,
and eventually to a single peak that corresponds to the ground-state recombi-
nation. The positions of peaks also change with the variation of the field. All
three exciton states have the same symmetry, so the crossings of their ener-
gies are forbidden (anticrossing). Instead of crossings we then observe closing
and receding of neighboring peaks, which is accompanied by the enhance-
ment and suppression of their intensities. This is similar to the variation of
dot dimensions (parameter r;,) in the absence of a magnetic field. Such behav-
ior of the photoluminescell{:e spectrum in a magnetic field was observed by
Bayer et al. [7] (see the description on page 76, and Figs. 6.14 and 6.15). For
10.2 Exciton in a Magnetic Field 163

a dot with diameter 35 nm in a zero magnetic field they report the existence
of two photoluminescence peaks, and also the appearance of the third one at
B ~ 4 T, and finally its disappearance when the. field was increased to 8 T.
Similar, but more complex photoluminescence spectra were also observed
by Raymond et al. [115] (see Fig. 6.17). However, their experiment was carried
out on highly excited dots, where a larger number of excitons (even up to 20)
were bound in a dot. The characteristic property of highly excited dots is the
the existence of three peaks in the absence of a magnetic field, which further
split in the magnetic field. In order to explain it within the presented above
model it is crucial to consider the problem of three particles in a dot: two
electrons and one hole (or two holes and one electron). As follows from the
detailed analysis [72], in these cases four states with zero angular momentum
of the recombining particles occur in the energy spectrum of the system. They
correspond to the two states of each electron (singlet and triplet states of the
electron pair). For GaAs, in the absence of a magnetic field two of these four
states are almost degenerate, and hence the three peaks are observed. The
inclusion of a magnetic field removes the degeneracy and unravels the fourth
transition.
References

1. ALSMEIER J., BATKE E., KOTTHAUS J. P., Voltage-tunable quantum dots on


silicon, Phys. Rev. B 41, 1699 (1990).
2. ANDREEV A. A., BLATNER Y. M., LOZOVIK Yu. E., Microscopic theory of
excitations of a quantum dot in strong magnetic field, Solid State Commun.
91, 581 (1994).
3. ASHOORI R. C., STORMER H. L., WEINER J. S., PFEIFER L. N., PEARTON
S. J., BALDWIN K. W., WEST K., Single-electron capacitance spectroscopy
of discrete quantum levels, Phys. Rev. Lett. 68, 3088 (1992).
4. ASHOORI R. C., STORMER H. L., WEINER J. S., PFEIFER L. N., BALDWIN K.
W., WEST K., N-electron ground state energies ofa quantum dot in magnetic
field, Phys. Rev. Lett. 71, 613 (1993).
5. ASHOORI R. C., Electrons in artificial atoms, Nature 379, 413 (1996).
6. BAKSHI P., BROIDO D. A., KEMPA K., Electromagnetic response of quantum
dots, Phys. Rev. B 42, 7416 (1990).
7. BAYER M., SCHMIDT A., FORCHEL A., FALLER F., REINECKE T. L., KNIPP
P. A., DREMIN A. A., KULAKOVSKII V. D., Electron-hole transitions between
states with nonzero angular momenta in the magnetoluminescence of quantum
dots, Phys. Rev. Lett. 74, 3439 (1995).
8. BAYER M., TIMOFEEV V. B., GUTBROD T., FORCHEL A., STEFFEN R.,
OSHINOWO J., Enhancement of spin splitting due to spatial confinement in
InxGal-xAs quantum dots, Phys. Rev. B 52, 11623 (1995).
9. BENISTY H., Reduced electron-phonon relaxation in quantum-box systems:
Theoretical analysis, Phys. Rev. B 51, 13281 (1995).
10. BOCKELMANN U., EGELER T., Electron relaxation in quantum dots by means
of Auger processes, Phys. Rev. B 46, 15574 (1992).
11. BOCKELMANN U., BRUNNER K., ABSTREITER G., Relaxation and radiative
decay of excitons in GaAs quantum dots, Solid State Electronics 37, 1109
(1994).
12. BOCKELMANN U., Electronic relaxation in quasi-one- and zero-dimensional
structures, Semicond. Sci. Technol. 9, 865 (1994).
13. BOCKELMANN U., ROUSSIGNOL P., FILORAMO A., HELLER W., ABSTREITER
G., BRUNNER K., BOHM G., WEIMANN G., Degenerate Fermi gas of excitons
in a single quantum dot structure, Phys. Rev. Lett. 76, 3622 (1996).
14. BOCKELMANN U., HELLER W., FILORAMO A., ROUSSIGNOL P., Micropho-
toluminescence studies of single quantum dots: I. Time-resolved experiments,
Phys. Rev. B 55, 4456 (1997). .
15. BOCKELMANN U., HELLER W., ABSTREITER G., Microphotoluminescence
studies of single quantum dots: II. Magnetic-field experiments, Phys. Rev.
B 55, 4469 (1997).
16. BONCH-BRUEVICH V. L., KALASHNIKOV S. G., Physics of semiconductors, in
Russian, Nauka, Moscow (1977).
166 References

17. BREY L., JOHNSON N. F., HALPERIN B. I., Optical and magneto-optical
absorption in parabolic quantum wells, Phys. Rev. B 40, 10647 (1989).
18. BRomo D. A., SHAM L. J., Effective masses of holes at GaAs-AIGaAs het-
erojunctions, Phys. Rev. B 31, 888 (1985). .
19. BRomo D. A., CROS A., ROSSLER U., Theory of holes in quantum dots,
Phys. Rev. B 45, 11395 (1992).
20. BRUNNER K., BOCKELMANN U., ABSTREITER G., WALTHER M., BOHM G.,
TRANKLE G., WEIMANN G., Photoluminescence from a single GaAs/AIGaAs
quantum dot, Phys. Rev. Lett. 69, 3216 (1992).
21. BRUNNER K., ABSTREITER G., BOHM G., TRANKLE G., WEIMANN G.,
Sharp-line photoluminescence and two-photon absorption of zero-dimensional
biexcitons in a GaAs/AIGaAs structure, Phys. Rev. Lett. 73, 1138 (1994).
22. BRUNNER K., ABSTREITER G., BOHM G., TRANKLE G., WEIMANN G.,
Sharp-line photoluminescence of excitons localized at GaAs/AIGaAs quan-
tum well inhomogeneities, Appl. Phys. Lett. 64, 3320 (1994).
23. BRYANT G. W., Electronic structure of ultrasmall quantum-well boxes, Phys.
Rev. Lett. 59, 1140 (1987).
24. BYCHKOV Yu. A., RASHBA E. I., Two-dimensional electron-hole system in
a strong magnetic field, Solid State Commun. 48, 399 (1983).
25. CHAKRABORTY T., HALONEN V., PIETILAINEN P., Magneto-optical transi-
tions and level crossings in a Coulomb-coupled pair of quantum dots, Phys.
Rev. B 43, 14289 (1991).
26. CHAKRABORTY T., Physics of the artificial atoms: Quantum dots in a magnetic
field, Comments Condo Mat. Phys. 16, 35 (1992).
27. DE CHAMON C., WEN X.-G., Sharp and smooth boundaries of quantum Hall
liquids, Phys. Rev. B 49, 8227 (1994).
28. DE CHAMON C., FREED D. E., WEN X.-G., Tunneling and quantum noise in
one-dimensional Luttinger liquids, Phys. Rev. B 51, 2363 (1995).
29. CHANG L. L., ESAKI L., Tsu R., Resonant tunneling in semiconductor double
barriers, Appl. Phys. Lett. 24, 593 (1974).
30. CIBERT J., PETROFF P. M., DOLAN G. J., PEARTON S. J., GOSSARD A.
C., ENGLISH J. H., Optically detected carrier confinement to one and zero
dimension in GaAs quantum well wires and boxes, Appl. Phys. Lett. 49, 1275
(1986).
31. DARNHOFER T., ROSSLER U., Effects of band structure and spin in quantum
dots, Phys. Rev. B 47, 16020 (1993).
32. DARNHOFER T., ROSSLER U., BROmO D. A., Far-infrared response of holes
in quantum dots: Band structure effects and the generalized Kohn's theorem,
Phys. Rev. B 52, 14376 (1995).
33. DARWIN C. G., The diamagnetism of the free electron, Proc. Cambridge Phi-
los. Soc., 27, 86 (1930).
34. DEMEL T., HEITMANN D., GRAMBOW P., PLOOGK., Nonlocal dynamic re-
sponse and level crossings in quantum-dot structures, Phys. Rev. Lett. 64,
788 (1990).
35. DINGLE R., WIEGMANN W., HENRY C. H., Quantum states of confined carri-
ers in very thin AlxGal-xAs-GaAs-AlxGal-xAs heterostructures, Phys. Rev.
Lett. 33, 827 (1974).
36. DINGLE R., GOSSARD A. C., WIEGMANN W., Direct observation ofsuperlat-
tice formation in a semiconductor heterostructure, Phys. Rev. Lett. 34, 1327
(1975).
37. DREXLER H., LEONARD D., HANSEN W., KOTTHAUS J. P., PETROFF P.
M., Spectroscopy of quantum levels in charge-tunable InGaAs quantum dots,
Phys. Rev. Lett. 73, 2252 (1994).
References 167

38. EKIMOV A. I., EFROS A. L., ONUSHCHENKO A. A., Quantum size effect in
semiconductor microcrystals, Solid State Commun. 56, 921 (1985).
39. FAFARD S., LEONARD D., MERZ J. L., PETROFF P. M., Selective excitation
of the photoluminescence and the energy levels of ultrasmall InGaAs/GaAs
quantum dots, App!. Phys. Lett. 65, 1388 (1994).
40. FAFARD S., LEON R., LEONARD D., MERZ J. L., PETROFF P. M., Visi-
ble photoluminescence from N -dot ensembles and the linewidth of ultrasmall
AlyInl-yAs/AlxGal-xAs quantum dots, Phys. Rev. B 50, 8086 (1994).
41. FAFARD S., LEON R., LEONARD D., MERZ J. L., PETROFF P. M., Phonons
and radiative recombination in self-assembled quantum dots, Phys. Rev. B
52, 5752 (1995).
42. FAFARD S., HINZER K., RAYMOND S., DION M., MCCAFFREY J., CHARBON-
NEAU S., Red-emitting semiconductor quantum dot lasers, Science 274, 1350
(1997).
43. FOCK V., Bemerkung zur Quantelung des harmonischen Oszillators im Mag-
netfeld, Zeitschrift fUr Physik 47,446 (1928).
44. FRANCESCHETTI A., ZUNGER A., Direct pseudopotential calculation of exci-
ton Coulomb and exchange energies in semiconductor quantum dots, Phys.
Rev. Lett. 78, 915 (1997).
45. FUKUI T., ANDO S., TOKURA Y., GaAs tetrahedral quantum dot structures
fabricated using selective area metalorganic chemical vapor deposition, App!.
Phys. Lett. 58, 2018 (1991).
46. GIRVIN S. M., JACH T., Interacting electrons in two-dimensional Landau
levels: Results for small clusters, Phys. Rev. B 28, 4506 (1983).
47. GRUNDMANN M., STIER 0., BIMBERG D., InAs/GaAs pyramidal quantum
dots: Strain distribution, optical phonons and electronic structure, Phys. Rev.
B 52, 11969 (1995).
48. G RUNDMANN M., BIMBERG D., Formation of quantum dots in twofold cleaved
edge overgrowth, Phys. Rev. B 55, 4054 (1997).
49. GUDMUNDSSON V., GERHARDS R. R., Self-consistent model of magnetoplas-
mons in quantum dots with nearly parabolic confinement potentials, Phys.
Rev. B 43, 12098 (1991).
50. HALLAM L. D., WEIS J., MAKSYM P. A., Screening of the electron-electron
interaction by gate electrodes in semiconductor quantum dots, Phys. Rev. B
53, 1452 (1996).
51. HANSEN W., SMITH T. P., LEE K. Y., BRUM J. A., KNOEDLER C. M., HONG
J. M., KERN D. P., Zeeman bifurcation of quantum-dot spectra, Phys. Rev.
Lett. 62, 2168 (1989).
52. HANSEN W., SMITH T. P., LEE K. Y., HONG J. M., KNOEDLER C. M.,
Fractional states in few-electron systems, App!. Phys. Lett. 56, 168 (1990).
53. HAWRYLAK P., Optical properties of a two-dimensional electron gas: Evolution
of spectra from excitons to Fermi-edge singularities, Phys. Rev. B 44, 3821
(1991).
54. HAWRYLAK P., PFANNKUCHE D., Magnetoluminescence from correlated elec-
trons in quantum dots, Phys. Rev. Lett. 70, 485 (1993).
55. HAWRYLAK P., Single-electron capacitance spectroscopy of few-electron arti-
ficial atoms in a magnetic field: Theory and experiment, Phys. Rev. Lett. 71,
3347 (1993). '
56. HAWRYLAK P., Far infrared absorption by screened D- states in quantum
wells in a strong magnetic field, Solid State Commun. 88, 475 (1993).
57. HAWRYLAK P., PALACIOS J. J., Correlated few-electron states in vertical
double-quantum-dot systems, Phys. Rev. B 51,1769 (1995).
168 References

58. HAWRYLAK P., WOJS A., LOCKWOOD D. J., WANG P. D., SOTOMAYOR-
TORRES C. M., PINCZUK A., DENNIS B. S., Optical spectroscopies of elec-
tronic excitations in quantum dots, Surface Science 361/362, 774 (1996).
59. HAWRYLAK P., Artificial impurity in interactlng electron droplets in a strong
magnetic field, Phys. Rev. B 51,17708 (1995).
60. HAWRYLAK P., WOJS A., BRUM J. A., Magneto-excitons in droplets of a chi-
ral Luttinger liquid formed in quantum dots in a magnetic field, Solid State
Commun. 98, 847 (1996).
61. HAWRYLAK P., WOJS A., BRUM J. A., Magneto-excitons and correlated elec-
trons in quantum dots in a magnetic field, Phys. Rev. B 54, 11346 (1996).
62. HEITMANN D., KOTTHAUS J. P., The spectroscopy of quantum dot arrays,
Physics Today, January, 56 (1993).
63. HEITMANN D., Far infrared spectroscopy of quantum-dots and anti-dot arrays,
Physica B 212, 201 (1995).
64. HERMAN M. A., Semiconductor heterojunction superlattices, in Polish,
Postepy Fizyki 34, 431 (1983).
65. IMAMURA K., SUGIYAMA Y., NAKATA Y., MUTO S., YOKOYAMA N., New
optical memory structure using self-assembled InAs quantum dots, Jpn. J.
Appl. Phys. 34, L1445 (1995).
66. ITOH T., IWABUCHI Y., KATAOKA M., Study on the size and shape of CuCl
microcrystals embedded in alkali-chloride matrices and their correlation with
exciton confinement, Physica Status Solidi B 145, 567 (1988).
67. JACAK L., WOJS A., NAWROCKA W., Analogy between shape transitions of
a quantum dot in magnetic field and a cranking nucleus, Phys. Low-Dim.
Struct. 3, 59 (1996).
68. JACAK L., KRASNYJ J., HAWRYLAK P., WOJS A., NAWROCKA W., On the
electronic properties of quantum dots, in: From quantum mechanics to tech-
nology, Editors: Z. Petru, et al., Lecture Notes in Physics, vol. 477, p. 39,
Springer Verlag (1996).
69. JACAK L., SITKO P., WIECZOREK K., Anyons and composite fermions, mono-
graph of Tech. Univ. Wroclaw (in Polish), Wroclaw (1995).
70. JACAK L., KRASNYJ J., WOJS A., Spin-orbit interaction in the quantum dot,
Physica B 229, 279 (1997).
71. JACAK L., KRASNYJ J., KORKUSINSKI M., WOJS A., Theory of radiative
recombination from the metastable excited states of quantum dots, preprint
of TU Wroclaw (1996).
72. JACAK L., KRASNYJ J., JACAK D., Photoluminescence of highly excited quan-
tum dots, preprint of TU Wroclaw (1997).
73. JAIN J. K., KAWAMURA T., Composite fermions in quantum dots, Europhys.
Lett. 29, 321 (1995).
74. JOHNSON B. L., KIRCZENOW G., Electrons in quantum dots: A comparison
of interaction energies, Phys. Rev. B 47, 10563 (1993).
75. JOHNSON N. F., PAYNE M. C., Exactly solvable model of interacting particles
in a quantum dot, Phys. Rev. Lett. 67, 1157 (1991).
76. JOHNSON N. F., Quantum-dot excitation spectrum: Laughlin-like states, Phys.
Rev. B 46, 2636 (1992).
77. KAMILLA R. K., JAIN J. K., Composite fermion description of correlated
electrons in quantum dots: Low-Zeeman-energy limit, Phys. Rev. B 52, 2798
(1995).
78. KANE C. L., FISHER M. P. A., POLCHINSKI J., Randomness at the edge:
Theory of quantum Hall transport at filling v = 2/3, Phys. Rev. Lett. 72,
4129 (1994).
References 169

79. KASH K., SCHERER A., WORLOCK J. M., CRAIGHEAD H. G., TAMARGO M.
C., Optical spectroscopy of ultrasmall structures etched from quantum wells,
Appl. Phys. Lett. 49, 1043 (1986).
80. KASTNER M. A., The single-electron transistor, Reviews of Modern Physics
64, 849 (1992).
81. KASTNER M. A., Artificial atoms, Physics Today, January, 24 (1993).
82. KHENG K., Cox R. T., MAMOR K., SAMINADAYAR K., TATARENKO S., Neg-
atively charged exciton X- and donor bound exciton DO X in planar doped
CdTejCdl_xZnx Te quantum well heterostructures, Journal de Physique IV
3, 95 (1993).
83. KHENG K., Cox R. T., MERLE D'AUBIGNONE Y., BASSANI F., SAMINADAYAR
K., TATARENKO S., Observation of negatively charged excitons in semicon-
ductor quantum wells, Phys. Rev. Lett. 71, 1752 (1994).
84. KIRSTAEDTER N., LEDENTSOV N. N., GRUNDMANN M., BIMBERG D., USTI-
NOV V. M., RUVIMOV S. S., MAXIMOV M. V., KOP'EV P. S., ALFEROV ZH.
1., RICHTER U., WERNER P., GOSELE U., HEYDENREICH J., Low threshold,
large To injection laser emission from (InGa)As quantum dots, Electronics
Lett. 30, 1416 (1994).
85. KLEIN 0., DE CHAMON C., TANG D., ABUSCH-MAGDER D. M., MEIRAV U.,
WEN X.-G., KASTNER M. A., WIND S. J., Exchange effects in an artificial
atom at high magnetic fields, Phys. Rev. Lett. 74, 785 (1995).
86. VON KLITZING K., DORDA G., PEPPER M., New method for high-accuracy de-
termination of the fine-structure constant based on quantized Hall resistance,
Phys. Rev. Lett. 45, 494 (1980).
87. KOHN W., Cyclotron resonance and de Haas-van Alphen oscillations of an
interacting electron gas, Phys. Rev. 123, 1242 (1961).
88. KUMAR A., LAUX S. E., STERN F., Electron states in a GaAs quantum dot
in a magnetic field, Phys. Rev. B 42, 5166 (1990).
89. LAHELD U. E. H., PEDERSEN F. B., HEMMER P. C., Excitons in type-II
quantum dots: Binding of spatially separated electron and hole, Phys. Rev. B
48, 4659 (1993).
90. LAHELD U. E. H., PEDERSEN F. B., HEMMER P. C., Excitons in type-II
quantum dots: Finite offsets, Phys. Rev. B 52, 2697 (1995).
91. LANDAU L. D., LIFSHITS E. M., Courseoftheoreticalphysics, vol. 3: Quantum
mechanics: non-relativistic theory, Pergamon Press, Oxford, New York (1977).
92. LAUGHLIN R. B., Quantized Hall conductivity in two dimensions, Phys. Rev.
B 23, 5632 (1981).
93. LAUGHLIN R. B., Quantized motion of three two-dimensional electrons in
a strong magnetic field, Phys. Rev. B 27, 3383 (1983).
94. LAUGHLIN R. B., Anomalous quantum Hall effect: An incompressible quantum
fluid with fractionally charged excitations, Phys. Rev. Lett. 50, 1395 (1983).
95. LEBENS J. A., TSAI C. S., VAHALA K. J., Application of selective epitaxy to
fabrication of nanometer scale wire and dot structures, Appl. Phys. Lett. 56,
2642 (1990).
96. LERNER 1. V., LOZOVIK Yu. E., Two-dimensional electron-hole system in
a strong magnetic field as an almost ideal exciton gas, Zh. Eksp. Teor. Fiz.
80, 1488 (1981); [Sov. Phys. JETP 53, 763 (1981)]. .
97. LORKE A., KOTTHAUS J. P., PLOOG K., Coupling of quantum dots on GaAs,
Phys. Rev. Lett. 64, 2559 (1990).
98. LUTTINGER J. M., Quantum theory of cyclotron resonance in semiconductors:
General theory, Phys. Rev. 102, 1030 (1956).
99. MACDoNALD A. H., REZAYI E. H., Fractional quantum Hall effect in a two-
dimensional electron-hole fluid, Phys. Rev. B 42,3224 (1990).
170 References

100. MACDoNALD A. H., YANG S. R. E., JOHNSON M. D., Quantum dots in


strong magnetic fields: Stability criteria for the maximum density droplet,
Australian J. Phys. 46, 345 (1993).
101. MAKSYM P. A., CHAKRABORTY T.,· Quantum dots in a magnetic field: Role
of electron-electron interactions, Phys. Rev. Lett. 65, 108 (1990).
102. MAKSYM P. A., CHAKRABORTY T., Effect of electron-electron interactions
on the magnetization of quantum dots, Phys. Rev. B 45, 1947 (1992).
103. MARZIN J.-Y., GERARD J.-M., IZRAEL A., BARRIER D., BASTARD G., Pho-
toluminescence of single InAs quantum dots obtained by self-organized growth
on GaAs, Phys. Rev. Lett. 73, 716 (1994).
104. MERKT V., HUSER J., WAGNER M., Energy spectra of two electrons in a har-
monic quantum dot, Phys. Rev. B 43, 7320 (1991).
105. MEURER B., HEITMANN D., PLOOG K., Single-electron charging of quantum-
dot atoms, Phys. Rev. Lett. 68, 1371 (1992).
106. PALACIOS J. J., MARTIN-MoRENO L., TEJEDOR C., Magnetotunneling
through quantum boxes in a strong correlation regime, Europhysics Lett. 23,
495 (1993).
107. PALACIOS J. J., MARTIN-MoRENO L., CHIAPPE G., LOUIS E., TEJEDOR C.,
Capacitance spectroscopy in quantum dots: Addition spectra and decrease of
tunneling rates, Phys. Rev. B 50, 5760 (1994).
108. PAQUET D., RICE T. M., VEDA K., Two-dimensional electron-hole fluid in
a strong perpendicular magnetic field: Excitom Bose condensate or maximum
density two-dimensional droplet, Phys. Rev. B 32, 5208 (1985).
109. PETROFF P. M., GOSSARD A. C., LOGAN R. A., WIEGMANN W., Toward
quantum well wires: Fabrication and optical properties, Appl. Phys. Lett. 41,
635 (1982).
110. PETROFF P. M., DENBAARS S. P., MBE and MOCVD growth and proper-
ties of self-assembling quantum dot arrays in III- V semiconductor structures,
Superlattices and Microstructures 15, 15 (1994).
111. PFANNKUCHE D., GUDMUNDSSON V., MAKSYM P. A., Comparison of
a Hartree, a Hartree-Fock, and an exact treatment of quantum-dot helium,
Phys. Rev. B 47, 2244 (1993).
112. PFANNKUCHE D., GUDMUNDSSON V., HAWRYLAK P., GERHARDS R. R., Far-
infrared response of quantum dots: From electron excitations to magnetoplas-
mons, Solid State Electronics 37, 1221 (1994).
113. PFEIFFER L., WEST K. W., STORMER H. L., BALDWIN K. W., Electron mo-
bilities exceeding 107 cm 2 /Vs in modulation-doped GaAs, Appl. Phys. Lett.
55, 1888 (1989).
114. RAYMOND S., FAFARD S., POOLE P. J., WOJS A., HAWRYLAK P., CHARBON-
NEAU S., LEONARD D., LEON R., PETROFF P. M., MERZ J. L., State filling
and time-resolved photoluminescence of excited states in Inl-xGaxAs/GaAs
self-assembled quantum dots, Phys. Rev. B 54, 11548 (1996).
115. RAYMOND S., HAWRYLAK P., GOULD C., FAFARD S., SACHRAJDA A.,
POTEMSKI M., WOJS A., CHARBONNEAU S., LEONARD D., PETROFF P. M.,
MERZ J. L., Exciton droplets in zero-dimensional systems in a magnetic field,
Solid State Commun. 101,883 (1997).
116. REED M. A., BATE R. T., BRADSHAW K., DUNCAN W. M., FRENSLEY W.
M., LEE J. W., SMITH H. D., Spatial quantization in GaAs-AlGaAs multiple
quantum dots, J. Vacuum Sci. Technol. B, 4, 358 (1986).
117. REED M. A., Quantum dots, Scientific American, March (1993).
118. RICE T. M., PAQUET D., VEDA K., On the ideal Boson ground state of
excitons in two dimensions in a strong magnetic field, Helv. Phys. Acta 58,
410 (1985).
References 171

119. RORISON J. M., Excitons in type-II quantum-dot systems: A comparison of


the GaAs/AIAs and InAs/GaSb systems, Phys. Rev. B 48,4643 (1993).
120. SHIKIN V., NAZIN S., HEITMANN D., DEMEL T., Dynamic response of quan-
tum dots, Phys. Rev. B 43, 11903 (19m). .
121. SHOJI H., MUKAI K., OHTSUKA N., SUGAWARA M., UCHIDA T., ISHIKAWA
H., Lasing at three-dimensionally quantum-confined sublevel of self-organized
Ino.s Gao.sAs quantum dots by current injection, IEEE Photonics Technol.
Lett. 7, 1385 (1995).
122. SIKORSKI C., MERKT U., Spectroscopy of electronic states in InSb quantum
dots, Phys. Rev. Lett. 62, 2164 (1989).
123. SILSBEE R. H., ASHOORI R. C., Comment on "Zeeman bifurcation of
quantum-dot spectra", Phys. Rev. Lett. 64, 1991 (1990).
124. SKOCPOL W. J., JACKEL L. D., Hu E. L., HOWARD R. E., FETTER L. A.,
One-dimensional localization and interaction effects in narrow (0.1 11m) silicon
inversion layers, Phys. Rev. Lett. 49, 951 (1982).
125. SMITH T. P., LEE K. Y., KNOEDLER C. M., HONG J. M., KERN D. P.,
Electronic spectroscopy of zero-dimensional systems, Phys. Rev. B 38, 2172
(1988) .
126. STONE M., WYLD H. W., SCHULT R. L., Edge waves in the quantum Hall
effect and quantum dots, Phys. Rev. B 45, 14156 (1992).
127. STRANSKI I. N., VON KRASTANOW L., Akad. Wiss. Let. Mainz Math. Natur
Kl lIb 146, 797 (1939).
128. TEMKIN H., DOLAN G. J., PANISH M. B., CHU S. N. G., Low-temperature
photoluminescence from InGaAs/lnP quantum wires and boxes, Appl. Phys.
Lett. 50, 413 (1987).
129. TSUI D. C., STORMER H. L., GOSSARD A. C., Two-dimensional magneto-
transport in the extreme quantum limit, Phys. Rev. Lett. 48, 1559 (1982).
130. VONSOVSKIJ S. V., Magnetism, in Russian, Nauka, Moscow (1971).
131. WAGNER M., MERKT U., CHAPLIK A. V., Spin-singlet-spin-triplet oscilla-
tions in quantum dots, Phys. Rev. B 45, 1951 (1992).
132. WOJS A., HAWRYLAK P., Negatively charged magneto-excitons in quantum
dots, Phys. Rev. B 51, 10880 (1995).
133. WOJS A., HAWRYLAK P., Charging and infrared spectroscopy of self-
assembled quantum dots in a magnetic field, Phys. Rev. B 53, 10841 (1996).
134. WOJS A., HAWRYLAK P., FAFARD S., JACAK L., Electronic structure and
magneto-optics of self-assembled quantum dots, Phys. Rev. B 54, 5604 (1996).
135. WOJS A., HAWRYLAK P., Exciton-exciton interactions in highly excited quan-
tum dots in a magnetic field, Solid State Commun. 100, 487 (1996).
136. WOJS A., HAWRYLAK P., JACAK L., Many-exciton complexes in self-
assembled quantum dots, Acta Phys. Pol. A 90, 1108 (1996).
137. WOJS A., HAWRYLAK P., Theory of photoluminescence from modulation-
doped self-assembled quantum dots in a magnetic field, Phys. Rev. B 55,
13066 (1997).
138. YANG S.-R. E., MACDONALD A. H., JOHNSON M. D., Addition spectra of
quantum dots in strong magnetic fields, Phys. Rev. Lett. 71, 3194 (1993).
139. ZRENNER A., BUTOV L. V., HAGN M., ABSTREITER G., BOHM G., WEI-
MANN G., Quantum dots formed by interface fluctuations in AIAs/GaAs cou-
pled quantum well structures, Phys. Rev. Lett. 72, 3382 (1994). .
140. ZRENNER A., Optical spectroscopy of single quantum dots, Proc. XI Con-
ference on Electronic Properties of Two-Dimensional Systems, Nottingham
(1995).
Index

absorption 16,23,31,32,46,51-57, bound states 44,100,101,103-105,


60,61,65,66,68-70,82,84,95, 112,148,157-159
111-118,124,133,138,153 boundary condition 100
absorption spectrum 53-55,57,65,66,
68-70,95,111,112,114-118,124,138 capacitance spectroscopy 52,83,84,
acoustic longitudinal phonons 61 88-91,107, 109, 136
adiabatic approximation 98 carrier scattering 61
angular momentum 17-19,23,29-42, carriers 5,21,23,33,61,70,73,77,83,
44-49,62,63,65,66,70,78,79,91,94, 141,157
98-101,104-107,112-114,119,127, center-of-mass excitation 31,35,36,
128,134,142,153,159,163 84,112
anharmonic terms 56 charge density 45, 84, 128
anisotropy 62,138 chemical potential 56,83-86,89-95,
annihilation operators 17,42,111 107-110, 136, 137
anticrossing of levels 54,56, 138, 162 circular representation 39
arrays of quantum dots 7,52-58,61, circular symmetry 18,19,23,24,38,
75,85 42,51,134
artificial atoms 2,95, 127 classical system 128, 130
Ashoori's experiment 93-95, 106, 133, coherent many-exciton states 120, 121
136 commutation relations 18,28,119
asymmetry 78,91 compact bubbles 141
Auger processes 61 compact droplet 41,42,44-46,65,66,
axial symmetry 148, 152, 153, 157, 161 68,92
azimuthal quantum number 19 compact states 39,40,42,67
composite fermions 2,38,39
conduction band 9,17,21,22,31,54,
band gap 10,11,22,33,36,65,72,81, 59,65,66,72,77,113-115,119
100, 116, 123, 142 conduction-band edge 97
band-gap renormalization 81,123 conduction-band electron 17,21,31,
band-structure model of the dot 24, 72
69,72,75-78,97,103,107,113,116, conduction-band energy 1,21,73,98,
118,139, 141 100
bi-exciton 121, 123, 125 confinement 1,2,32-38,42,56,61,79,
binding of an exciton 23,120 85,88, 104, 105
binding of holes 21,22,141,157 confining potential 8,69,70; 78, 80,
Bohr radius 24,25,42,97, 113, 131 90,93,99,10,17,18,32,33,35,37-39,
Bose-Einstein condensation of excitons 46,48,52,53,56,62,63,103-105,108,
121 110-113,116,124,127,131,134,138,
Bose-Einstein distribution 62 142, 149, 155
bottom-gate voltage 56 consecutive approximations 143,157
bound excitons 60, 153 constant-interaction model 83
174 Index

contravariant heterostructure 21, 22 electron-hole overlap 120,154


core 107,123,125 emISSIon 2,23,31,60-63,70,75,76,
Coulomb interaction 16,22,30,32, 80,8~, 113-115, 153
33,38,40,42,43,52,54,58,78,79,92, empty dot 53,69,84,114,116,117,
105-108,112,113,116,119,127,130, 123,141,142
132,141 etching 5-7,11,12,14
Coulomb repulsion 32,33,58 exact diagonalization 36,38,92
covariant heterostructure 21, 22 excitation energy 17,30,32,33,36,40,
covering layer 77 41,43,45,51,53,54,56,57,59,60,64,
creation operators 17,18,42 81,90,91,109,110
crossings levels 19,90-92,95, 107, 134, excitation power 81,100,124,156
135, 139, 162
crystalline structure 21 far infrared 2,16,30--32,51,63,84,
cyclotron frequency 17,31,33,134, 109,133
142 Fermi edge singularity 66
cyclotron motions 157 Fermi golden rule 61,109,111,113
cyclotron-resonance 31 Fermi level 44,54,57,66,130
filling factor 37,38,41,44,45,65,66,
degeneracy 18,19,32,36,40,41,78, 87,88,91,92
82,91,92,101,103-106,109,116-118, flux 39
120,129,163 Fock correction 128, 132
density of states 15,61,83,100,101, Fock-Darwin energy levels 17, 18,20,
103,104,118 30,51,54,56,79,90-92,103-105,107,
diamagnetic states 49, 52 109,116
dielectric constant 30,97,100,127 forbidden transitions 56, 6~
dipole approximation 20,30, 110, 133 fractional quantum Hall effect 1,2,
direct process 72 37,38,88
donor 52
double-well electron potential 80, 144, Gaussian well 16, 142
147,149,151,157,158,160 giromagnetic Pauli factor 31,127
dynamical symmetries 116
harmonic interaction 40,41
edge effects 8 Hartree approximation 16,131,132,
edge magneto-roton 45 136
edge reconstruction 41 Hartree-Fock approximation 49,92,
effective attraction 81 93,119,127,128,133,135,138
effective length scale 18,97 heat capacity 46,47
effective mass 1, 17,23-25,52,62,70, heavy-hole 23,24,31,77,114,115
97,100,103,112,128,142,155 helicity 157
effective Rydberg 42,97,113 heterojunction 85
effective-mass approximation 17,22, heterostructure 21,22,54,84
23,25,80,97,127,141 hidden symmetries 119,124
eigenstates 17-19,34,38,39,41-43, high electron mobility transistor 89
47,78,79,84,103,104,113,115,119, Hilbert space 38,123
120, 129, 142 hole-hole exchange 75,81
electrodes 5,7-9, 14, 17, 52, 56, 57, 83, Hund principles 32,106
85,87-89,94-96,109 hydrostatic pressure 100
electron-electron exchange 70,75,81
electron-hole attraction 66, 70, 79, incompressible states 36-38,41
112, 120, 121 interband transitions 59,65, 115
electron-hole pair 22,23,43,44, inter diffusion 10
59-61,63,70,78,112,116-119,141, interface 5,21,77,85,97,99, 100, 102,
152 103,141
electron-phonon coupling 62 interlayer diffusion 141
Index 175

intersubband mixing 23-25 photoluminescence 59-61,63,64,


70-82,117,123,124,141,153-155,
Jacobi coordinates 27,28 157,161,163
Jastrow factor 39 photolurriinescence peaks 70,73-76,
78-82,117,118,124,153-155,161,
kinetic energy 32,45,66,105,106, 163
121,123,127,132 photon absorption 65,82, 114, 115,
Kohn theorem 16,25,27,31,32,51, 133
52,110,111 Planck constant 128
point r 72-75
Landau quantization 2,33,86,104 polarization operator 119
laser beam 10,53,59,60,75,81,100 polarized light 66,68,69,114-116,120
laser power 75-77,81,117 potential curvature 143-145, 150, 155
lateral confinement 7,8, 16,51,98 potential fluctuations 88
Laughlin's function 40,41 principal quantum number 19
layer thickness 12, 13, 17,72,97,99 pseudopotential 22
lifetime 23,60,72,73,82
light-hole 23,24,31,77,115 quantization of motion 79
linearly polarized photon 114 quantum well thickness 13,16,62,70,
lithographic techniques 2, 7 72-75,81,98
Luttinger Hamiltonian 24,25 quantum wire 2,5, 12, 15,70,71
quasi-momentum 72
magic states 35-38,41,45,65-67,106
magnetic length 18,33,79,88,92 radial quantum number 19
magnetization 46,48,49 radiative recombination 2,23,60-63,
magneto-exciton 69,116,118 77,82,153
magneto-roton 41, 43-46 Raman scattering 2
many-electron states 35,42 recombination 60-64,72-78,81,113,
many-exciton complexes 77,119, 120 114,154, 155, 162
metal electrode 17,85, 109 recombination rates 60,63,154
metastable state 60,63,64, 75, 77, 80, relative motion 31,40,51,112
153,156,157,161 relaxation 60-63,73,80,82, 124, 153,
microcrystals 10, 11 161
mixing 23-25,33,38,47,103 resonance energy 21,31,51-56,58,
133, 138
nanostructure 5,17 rotational symmetry 54
noninteracting system 39,46,49,51,
60, 78, 105, 107, 108, 113 scattering 100, 120-123
scattering states 16, 100, 105, 107
occupied electronic states 66, 130, 133 Schrodinger equation 2,16,97
off-diagonal element 44, 134 screening 40,94,109
optical longitudinal phonons (LO) 61 second quantization 104,118
optical transitions 20,21,65,66,73, selection rules 20,21,62,63,79,96,
115, 116, 133, 138 133,138
orbitals 20,23,31,39,41,44-46,65, selective growth 11,12
84,104,107,109,116,119,120 self-assembled quantum dots 3,13,
oscillator strength 21,95 14,22,25,32,61,76,80-82,97,98,
overlap of orbitals 120 101-104,107,111-113,116,117,118,
122,124,141,157
parabolic well 2,16, 17,30,31,41,46, self-consistent field 127,130,142-145,
51,101-103,110,131,132 148-150,154,158,159,161,162
paramagnetic states 49, 52 self-organized growth 12
Pauli exclusion principle 32,39,44, single electron capacitance spectroscopy
60,66,67,84,105,115 88-90
176 Index

single quantum dot 7,54,57,61,64, thermodynamic average 47-49


73,88 thin layer 1,5,13,72,77
single-band approximation 25, 76 top-gate voltage 56
singlet 74,75,115,121,163 transfer matrix 99, 100
Slater determinant 38,39 triplet 74,75,163
spatial confinement 61,79,88,104 tunneling 8,89,90,95,96
splitting of photoluminescence peaks
82, 153, 155, 156 uniaxial stress 100
spin-orbit coupling 31,32,51,52,56,
57,127,130,135 vacuum 42,77,119,120
spin-orbit interaction 25,31,32, valence band 9,21-23,25,31,59,65,
49,51,52,56,91,95,110,127,129, 66,72,77,100,103,113-115,119
131-133,135,136,138,139 variational method 43,129
strain 12,13,25,77,98,100,103
Stranski-Krastanow transition 12-14 wave function 19,23,34,35,39,40,70,
strong magnetic field 18,33,36,38, 79,94,98-100,109,114,128,129,138,
40-42,65,69,70,76,87,91,92,104, 143,145,147,148,150,151,158-161
109,121,159,161 wetting layer 13,97,98,100,101, 104
subbands 23-25,31,77,98,103,114,
115 Zeeman splitting 19, 36, 46, 75, 104,
superatoms 2 109,112,134
symmetric gauge 17, 112, 128 zero-dimensional system 15,16,70,
103

Vous aimerez peut-être aussi