Sie sind auf Seite 1von 159

OPTICAL STUDY OF COUPLING MECHANISMS IN QUANTUM DOT QUANTUM WELL HYBRID NANOSTRUCTURE

OPTICAL STUDY OF COUPLING MECHANISMS IN QUANTUM DOT QUANTUM WELL HYBRID NANOSTRUCTURE

A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Microelectronics-Photonics

By

Vitaliy Dorogan Uzhgorod National University Bachelor of Science in Physics, 1999 Uzhgorod National University Master of Science in Physics, 2000

August 2011 University of Arkansas

UMI Number: 3476090

All rights reserved INFORMATION TO ALL USERS The quality of this reproduction is dependent upon the quality of the copy submitted. In the unlikely event that the author did not send a complete manuscript and there are missing pages, these will be noted. Also, if material had to be removed, a note will indicate the deletion.

UMI
Dissertation Publishing

UMI 3476090 Copyright 2011 by ProQuest LLC. All rights reserved. This edition of the work is protected against unauthorized copying under Title 17, United States Code.

ProQuest LLC 789 East Eisenhower Parkway P.O. Box 1346 Ann Arbor, Ml 48106-1346

ABSTRACT The interaction between nanostructures of different dimensionality, twodimensional quantum well (QW) and zero-dimensional quantum dots (QDs), has been studied by photoluminescence (PL) spectroscopy methods in the InAs/GaAs QDs InGaAs/GaAs QW hybrid nanostructure. A strong dependence of the PL spectra on the separation barrier thickness and height was observed. In samples with thick (high) barrier between the QW and QDs essentially no carrier transfer took place and the nanostructures behaved independently. If the separation between the QW and QDs was only several nanometers, the effective carrier transfer from the QW ground state to the 3rd QD excited state occurred. It was shown that the QD state filling at elevated excitation intensities had a strong effect on the tunneling efficiency. The PL data were successfully described by the rate equation model which included state filling effects. The tunneling times extracted from the continuous-wave PL data predicted a subpicosecond tunneling dynamics for the QD-QW samples with spacers < 2 nm, which suggested a microscopic coupling mechanism in this system to be the resonant electron tunneling. Anomalous dependence of the QD PL excitation (PLE) intensity at resonance QW excitation on the separation barrier thickness was observed and explained in terms of intermediate coherent tunnel coupling based on the optical Bloch equation model. It was shown that the resonant coherent tunneling could be more effective between the QW excited state and the QD excited state due to the larger overlap of the electronic wavefunctions. This interaction was observed in the PLE spectra as a peak splitting of the second QW subband resonance, indicating the hybridization and derealization of the electronic states between the QW and the QDs. The possibility of tuning the QW energy

levels by varying the QW composition was demonstrated on a set of samples. When the QW lowest state was tuned in resonance with the WL state, the anti-crossing of the QW and WL resonances was observed in the PLE spectra which indicated a strong coupling between these two states with the tunneling dynamics on a few hundred femtosecond scale.

This dissertation is approved for recommendation to the Graduate Council. Dissertation Director:

Dr. Gregory Salamo

Dissertation Committee:

Dr. Huaxiang Fu

Dr. Min Zou

Dr. Surendra Singh

Prof. Ken Vickers (ex officio)

The following signatories attest that all software used in this dissertation was legally licensed for use by Mr. Vitaliy Dorogan for research purposes and publication.

Mr. Vitaliy Dorogan, Student

Dr. Gregory Salamo, Dissertation Director

This dissertation was submitted to http://www.turnitin.com for plagiarism review by the Turnltln company's software. The signatories have examined the report on this dissertation that was returned by Turnltln and attest that, in their opinion, the items highlighted by the software are incidental to common usage and are not plagiarized material.

Prof. Ken Vickers, Program Director

Dr. Gregory Salamo, Dissertation Director

2011 by Vitaliy Dorogan All Rights Reserved

DISSERTATION DUPLICATION RELEASE I hereby authorize the University of Arkansas Libraries to duplicate this dissertation when needed for research and/or scholarship.

Agreed Vitally Dorogan

Refused Vitaliy Dorogan

ACKNOWLEDGEMENTS I would like to express my sincere gratitude to my academic advisor, Distinguished Professor Gregory J. Salamo, who introduced me to amazing world of nanoscience and inspired me in many ways. Dr. Salamo, who has a great knowledge of both experimental and theoretical physics, set an example of ideal scientist to me. It was a great pleasure for me to work with Dr. Yuriy Mazur, Dr. Vasyl Kunets, Dr. Dorel Guzun, Dr. Georgiy Tarasov, and Dr. Morgan Ware. These people gave me continuous support, encouragement, and help with conducting the experiments and discussing the results. I appreciate a cooperation with Prof. Christoph Lienau who made a big contribution to our understanding of the coherent phenomena. I am grateful to Dr. Euclydes Marega Jr. for long hours spent and great passion shown during the MBE growth of many high quality samples used in this work. I thank Dr. Mourad Benamara for taking numerous TEM images of the samples used in this research. I am also grateful to my dissertation committee members Dr. Huaxiang Fu, Dr. Min Zou, Dr. Surendra Singh, and Professor Ken Vickers for critical reviewing of my dissertation and giving me suggestions that helped me improve it. Finally, I wish to thank all of the postdocs and students in Dr. Salamo's group and MicroEP community for being inspiring and supportive all this time on my long way to achieving this important goal in my life. The research presented in this dissertation was financially supported by the National Science Foundation under Grants # DMR-0520550 and DMR-1008107. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author and do not necessarily reflect the views of the National Science Foundation. vii

DEDICATION I dedicate this dissertation to my parents, Gennadiy and Ida Dorogan. I appreciate that you have never interfered with my decision to become a scientist and always supported me in every possible way.

viii

TABLE OF CONTENTS I. II. INTRODUCTION COUPLING IN DOUBLE LAYER NANOSTRUCTURES 2.1 Double quantum well system 2.2 Double quantum dot system 2.3 Quantum d o t - quantum well system 2.4 Strain-induced QDs in a QW 2.5 Applications of QD-QW system 2.6 Summary III. SAMPLES AND EXPERIMENTAL METHODS 3.1 QD-QW sample design and growth 3.2 Structural characterization of QD-QW samples 3.3 Continuous-wave PL and PL excitation measurements 3.4 Time-resolved PL measurements 3.5 Summary 1 3 4 11 19 26 28 30 31 31 35 39 41 42

IV. EFFECT OF QD STATE FILLING ON TUNNELING PROCESSES IN QDQW NANOSTRUCTURES 4.1 Effect of GaAs spacer thickness 4.2 Effect of AlAs and AlGaAs barrier height 4.3 PL decay measurements in QD-QW samples 4.4 Rate equation model 4.5 Fitting experimental data and determination of tunneling time 4.6 Effect of QW composition variation on tunneling processes 4.7 Strain-induced QDs in dot-in-well structure 4.8 Summary 43 44 47 51 56 61 70 78 86

IX

V.

SIGNATURES OF COHERENT TUNNELING IN QD-QW NANOSTRUCTURES 5.1 Anomalous dependence of QD PLE signal on barrier thickness under resonant QW excitation 5.2 Optical Bloch equation model 5.3 Resonant tunneling time measurement 5.4 Hybridization of the QD and QW excited states variation 5.6 Summary 89 94 100 102 109 114 115 119 126 130 134 88

5.5 Anti-crossing between WL and QW states tuned in resonance by QW composition

VI. CONCLUSIONS AND OUTLOOK References Appendix A: Derivation of Equations (4.13) and (4.14) Appendix B: Description of Research for Popular Publication Appendix C: Executive Summary of Newly Created Intellectual Property Appendix D: Potential Patent and Commercialization Aspects of listed Intellectual Property Items Appendix E: Broader Impact of Research Appendix F: Microsoft Project for PhD MicroEP Degree Plan Appendix G: All Publications Published, Submitted and Planned

135 137 139 143

I.

INTRODUCTION

Nanotechnology, an amazing field of research, technology, and applications, has grown beyond our bravest expectations in the past two decades. Nanostructures, structures made of nanometer-sized pieces of matter with specific shapes, are the elements or building blocks of nanotechnology. By controlling size and shape of nanoscale material systems, it is possible to engineer materials with unique properties which bring innovation in all fields that one can imagine. Products of nanotechnology can be found in wide variety of important applications: electronics, optoelectronics, telecommunications, power generation, military and space applications, all kinds of coatings, paints, lubricants, biomedical applications, etc. Semiconductor nanostructures hold promise to drastically improve the

functionality and increase the variety of devices based on existing semiconductor microelectronic technology. Moreover, semiconductor nanostructures are to bring the quantum phenomena to a practical use, e. g. quantum computation. Among the semiconductor nanostructures are quantum wells (QWs), quantum wires (QWRs), quantum dots (QDs), and all possible combinations of the three, as well as multilayered structures. Unique properties of the QDs allow for realization of single QD devices [1] such as single-electron transistors [2,3], single-photon emitters [4,5,6], devices for storage of a single electron [7], electron spin memory devices [8,9], and a concept of quantum information processing [10]. The next step needed for successful optical quantum computing is a creation of even more complex quantum systems that will contain various nanostructures interacting with each other [11,12]. Such systems that are made of different nanoscale components 1

with quite different properties are called hybrid nanostructures. For example, combining metal nanostructures with semiconductor nanostructures is of great interest because this combination gives an opportunity to utilize a unique ability of metals to confine the light within nanometer-sized structures along with the ability of semiconductors to detect this light and convert it into an electrical signal [13]. These metal-semiconductor hybrid nanostructures are predecessors of future nanophotonic integrated circuits. Of course, hybrid nanostructures could be made of solely semiconductor components that exhibit very different properties. The fundamental questions about the details of mechanisms of interaction, often referred to as coupling, between different nanostructures are still to be answered. One can consider that coupling between two nanostructures exists if exchange of charge carriers and/or energy takes place in such system, but how does one confirm experimentally that two nanostructures are coupled? If one of the two coupled nanostructures is being disturbed externally, for example by light of a certain wavelength, it will change its energy state. The other nanostructure should respond to that change by changing its own state, which can be observed. How does this transfer occur? How many different mechanisms do exist in a particular system? What is the contribution of each mechanism? How does the coupling change if one changes the distance or the surrounding material that separates two nanostructures? An attempt to answer these and many other questions about coupled nanostructures has been made in this dissertation research by exploring optical properties of the quantum dot - quantum well (QD-QW) hybrid nanostructure.

I I . COUPLING IN DOUBLE LAYER NANOSTRUCTURES


To build a quantum well nanostructure one must use at least two kinds of semiconductor materials, one of which has a greater bandgap than the other. The size of a smaller bandgap material in at least one dimension should be only several nanometers thick, close to the de Broglie wavelength of the carrier. The smaller bandgap material should be surrounded by the larger bandgap material. This provides a potential barrier at the interface of the two materials that confines charge carriers within the nanostructure. The schematic way of presenting such nanostructures is the one-dimensional energy diagram along one of the directions of quantum confinement as shown in Figure 2.1.

GaAs

GaAs

r f*I

Conduction Band

CD C

LU

-^-

Valence Band

Direction of confinement Figure 2.1 Schematic representation of the energy band diagram of a semiconductor nanostructure using InGaAs QW as an example.

In the nanostructure, both electrons and holes are confined within a potential well and are allowed to possess only certain discrete energies, quantum states. The values and the number of the discrete energy levels depend on the height of the potential barriers and on the width of the potential well. The most suitable way to study coupling between two semiconductor nanostructures is to make a structure that consists of two layers (each layer contains the same or different type of nanostructures) separated by a barrier material of known thickness, so called double layer nanostructures. Coupling of two nanostructures may involve many different mechanisms that can play more or less significant roles: coherent tunneling [14,15], phonon- and Auger-assisted (incoherent) tunneling [16,17], dipoledipole or Forster interaction [18,19], long-range coupling through the radiation field [20,21], long-range polariton interaction [22]. The main coupling mechanism that takes place at low temperature (~10 K) in the semiconductor nanostructures discussed in this study is quantum (coherent or incoherent) tunneling. The strength of tunnel coupling depends on the distance that separates two adjacent nanostructures, on the barrier material, and on whether the energy levels in the two nanostructures are aligned.

2.1 Double quantum well system Quantum wells, thin layers of semiconductor material sandwiched between a different bulk-sized semiconductor with high quality abrupt interfaces (as in Figure 2.1), were among the first semiconductor nanostructures that crystal growth techniques were able to produce about two decades ago. Hence, a double QW structure was a logical

opportunity

for

exploration

of

coupling

mechanisms

between

semiconductor

nanostructures. There are two types of double QW structures: symmetric double QWs (both wells have the same width and composition) and asymmetric double QWs (the structure consists of a narrow QW and a wide QW, both wells are of the same composition). In the symmetric double QW system, the energy levels of electrons and holes in both wells are perfectly aligned. As one brings two identical QWs together to a distance of several nanometers the wave functions overlap and the energy levels start to interact and push apart from each other. This is a direct analogy to the situation when two separate hydrogen atoms are brought together and form the H2 molecule, their atomic states split into doublets. On the other hand, if the distance between the two QWs is rather large (> 20 nm), each QW can be considered as an identical standalone QW with no any interaction with the other QW. It is shown in Figure 2.2 how the lowest conduction energy states splitting in the coupled QWs depends on the barrier thickness in the systems with (a) 40 A thick QWs and (b) 80 A thick QWs [23]. In superlattices, heterostructures with many identical QWs divided by thin barriers, the splitting of the interacting levels results in a formation of minibands, which is analogous to a formation of bands instead of discrete atomic states when individual atoms are bound together forming a solid matter. An asymmetric double QW system has an advantage over a symmetric double QW system, because the optical signal (for example, PL or absorption) from the two QWs is spectrally separated (see Figure 2.3). Another advantage of an asymmetric double QW system is that by applying an external electric field along the growth direction it is possible to tune the energy levels in and out of the resonance (see Figure 2.4).

160

20

40 60 B(A)

80

100

Figure 2.2 Conduction energy levels splitting as a function of barrier thickness B of (a) 40 A-B-40 A and (b) 80 A-B-80 A symmetric double QWs systems. Inset is a schematic conduction band energy diagram [23].

CB

WQW

NQW

E 2 >E 1

(a)

(b)

VB E2 Energy

Figure 2.3 Asymmetric double QWs system: (a) energy band diagram and (b) corresponding PL spectrum of wide QW (WQW) and narrow QW (NQW). 6

-<

-<

Nonresonant tunneling

Resonant tunneling

(a)

(b)

Figure 2.4 Asymmetric double QWs system: (a) nonresonant tunneling of electrons and holes in an unbiased structure and (b) resonant tunneling of electrons in a biased structure.

Under resonant condition, as shown in Figure 2.4(b), and with sufficiently thin barrier (< 10 nm) between the QWs, coherent tunneling of particles can take place. At a microscopic (quantum) level, particles (electrons, holes, protons etc.) have wave-like properties. If the wave functions of the two states located in two adjacent QWs coherently overlap, then coherent tunneling occurs between these two states. When the two levels are resonantly aligned, it is possible for a particle to tunnel back and forth in an oscillatory manner (coherent oscillations) [24,25]. Coherent tunneling is characterized by a coherent time Tc defined as half of the oscillation period and can be written as
T

h 2AE

(2.1)

where h is a Planck's constant and AE is the energy separation of the two coupled states. Resonant tunneling takes place when nominally |AE|~0 (see Figure 2.5) [25].

Under Bias

Resonance

Over Bias

AE

AE<0

AE-0

AE>0

Figure 2.5 Conduction band profiles of asymmetric double QWs when under biased, in resonance, and over biased by an external electric field. AE is an energy separation between the levels [25]. With thin barriers, AE never reaches zero and is equal to the value of the level splitting that occurs due to the level interaction. This splitting at resonance can be seen in the level anti-crossing phenomenon. Figure 2.6 shows the results of calculations for the lowest energy electron states in both wells as a function of applied electric field [26]. Here, the situation can be described following steps shown in Figure 2.5. Before resonance, the level in the left QW is lower than in the right QW. At resonance, AE reaches its minimum value and the states switch their localization places. After resonance, AE increases again. 80 Anti-crossing ^
>

70

Right

E ^60 E >
c
111

10 20 30 Electric field (kVcm"1)

40

Figure 2.6 Dependence of the lowest conduction band states of 600A600A-500A double QWs on the external bias [26]. 8

Experimentally the anti-crossing can be observed by following the PL or absorption peak positions while changing an external bias [27]. Measuring the anti-crossing allows to distinguish and investigate separately electron and hole tunneling, since the electron states anti-cross at a different bias than the hole states [23]. The process of tunneling is characterized by the tunneling time, an average time that is needed for a particle to tunnel through a potential barrier. Many researchers have tried to compare the resonant tunneling time measured experimentally with the theoretical coherent time Tc. The experimental value always turned out to be larger than Tc [28]. The reason is that in a real, nonideal, system there are many factors and processes, like scattering on defects, interface roughness, alloy fluctuations, and the fact that one experimentally measures the tunneling of the whole ensemble of particles instead of a single particle tunneling, may affect the tunneling time. Under nonresonant condition, as shown in Figure 2.4(a), both electrons and holes can tunnel. However, the question is what exactly tunnels: electrons and holes separately, entire excitons (neutral quasiparticles which are electron and hole bound together via Coulomb interaction), only electrons? The answer to this question strongly depends on a particular system and conditions. For thin barriers (< 40 A) and low temperatures (4.2 K), correlated tunneling of electrons and holes (excitons) takes place [29]. If the electronic levels in the two QWs with thin barrier are tuned in resonance, at high excitation intensities the electrons that have tunneled to a second well can form a crossed exciton with the holes that remained localized in the first well (see Figure 2.7) [30]. The so-called cross excitons exhibit long decay times of PL (~10ns) caused by the reduced wave function overlap. Tunneling of an intact exciton or switching to a cross exciton is even

Figure 2.7 Energy band diagram of biased double QW system with direct (X and X') and cross (indirect in real space) (X" and X'") exciton transitions [30]. more likely to happen in materials that have high exciton binding energy, for example in II-VI semiconductors [31]. Another way to facilitate a fast nonresonant hole transfer is to build a heterostructure that consists of ternary compounds [32]. Hole scattering on the alloy fluctuations turned out to be an effective mechanism for fast hole tunneling. If the barrier becomes thicker (> 40 A), the electrons tunnel faster than the holes (the tunneling rate of the electrons is about one order faster than that of the holes) due to the large difference of the effective masses. This results in a charge buildup at high excitation intensities and, hence, to a slight band bending due to the internal electric field, which in turn shifts the electron and hole energy levels [33,34]. In a nonresonant regime, if the energy level mismatch is larger than the energy of a longitudinal optical (LO) phonon then the electron tunneling transfer via thick barrier turns out to be LO-phononassisted process [34]. If this energy mismatch is smaller than the value of the LO phonon, the nonresonant electron tunneling occurs via scattering on defects and interface imperfections of the double QW heterostructure [35]. The electron tunneling rate decreases exponentially with an increase of the barrier thickness [23]. 10

2.2 Double quantum dot system A quantum dot is often embedded in the matrix made of a material with a larger bandgap which creates a quantum confinement for charge carriers in all three dimensions, and has attracted attention of researchers in the past 15-20 years. The QDs are often called artificial atoms because of their discrete, atomic-like, energy states for the electrons and holes. The properties of the QDs (optical and electrical) depend on the relative energy position of these levels and can be easily engineered by changing the size and chemical composition of the QDs. There are different methods of the QD growth. Spherical QDs with a precise control of the QD size and homogeneity can be obtained by chemical colloidal synthesis. Usually, these nanoparticles are made of II-VI semiconductors and are dispersed in some kind of a chemical solution. Additional procedures are required to attach colloidal QDs to the semiconductor surface if a combination with a planar structure is needed. In epitaxial growth, for example by molecular beam epitaxy (MBE) or metal-organic chemical vapor deposition (MOCVD), the QDs are grown on a semiconductor substrate with an atomiclayer-thickness control. During the growth, a thin layer of semiconductor material with slightly different lattice constant from that of the substrate (for example, between GaAs and InAs lattice constant difference is ~7%) is deposited on the substrate. The QDs form spontaneously due to the strain relaxation in a lattice mismatched structure (StranskiKrastanov growth mode) [36]. Varying the growth conditions and choosing different types of substrates to grow on, one can control QD shape, size, density, and even site location [37,38,39,40]. Using a droplet epitaxy MBE growth, the QD clusters of various configurations can be created [41,42]. All above mentioned techniques are so-called

11

"bottom-up" approaches to the QD formation. In a "top-down" approach, the QDs are fabricated using patterning of the QW (or any other heterostructures that contains a twodimensional electron gas) by means of various techniques such as lithography. Only samples with Stranski-Krastanov based QDs grown by MBE were used in this work. If two QDs are placed close to each other so that the wave functions of the carriers in both dots overlap, coupling may occur between these two QDs. This situation is very similar to a formation of a molecule out of the two atoms. Following this analogy, the system of the two (or more) interacting QDs is often called a "QD molecule". When considering the MBE-grown double layer QDs, one has to deal with two ensembles of QDs. This means that due to the size distribution of the QDs the sum of slightly different discrete energy levels ends up as a wide band in the optical spectrum. Also, it is important to distinguish between lateral and vertical coupling. Figure 2.8 illustrates an ideal case where the QDs in the layer 1 are located directly on top of the identical QDs in the layer 2. In the Figure 2.8, the arrows indicate interdot distances within the layer and between the layers. Because of the flattened hemispherical shape of the QDs, the lateral coupling can be achieved only in dense QD arrays (with QD density

QD Lateral Layer 1 QD Vertical QD Layer 2 Figure 2.8 Schematic cross-section view of a double layer QD system.

12

greater than 1010 cm2) [43,44]. In the vertical QD arrangement, the interdot distance can be significantly reduced compared to the lateral QDs and, therefore, much stronger coupling can be achieved. Moreover, it is more difficult to control the lateral interdot distance as compared to the vertical distance between the QD layers, which is important for the study of the tunneling time as a function of the barrier thickness [45]. By analogy with the double QW system, the double layer QDs can be symmetric and asymmetric, and the carrier tunneling can be resonant and nonresonant. It is reasonable to perform an explicit study of resonant carrier transfer only on a single pair of QDs, where the energy states are sharply and clearly defined in the optical spectra. On the other hand, if the system consists of the two QD ensembles with the same average size, due to the QD size variation, some portion of the QDs from both layers is always inresonance and another portion of the QDs is always off-resonance. Therefore, to study nonresonant carrier transfer between the two QD ensembles it is reasonable to have two layers with different sizes of the QDs with uniform distributions within each layer. In this case, the optical signal from the ground exciton transition of each QD layer has different energy in the spectrum [46,47]. The PL spectra of the asymmetric double QD samples with various spacer thicknesses are shown in Figure 2.9(a) [47]. The QDs in the first (seed) layer formed by 1.8 monolayers (ML) of InAs deposition were smaller in size (small QDs - SQD) and were kept the same for the whole set of samples. For the second layer, QDs were intentionally made larger (large QDs - LQD) by means of 2.4 ML InAs deposition. Most of the QDs in the second layer, correlated by a strain field from the seed layer, grew directly on top of the first layer QDs. Density and size of the QDs in the second layer

13

1.0

1.2 Photon energy (eV)

1.4

Figure 2.9 (a) PL spectra of the asymmetric double QD structures with different spacer thicknesses taken at T= 10 K. (b) Schematic energy band structure along the growth direction of the double QD system with the processes of carrier generation, recombination, and interdot tunneling represented by arrows [47]. depended on the spacer thickness [47,48]. The QD layers were divided by a GaAs spacer with thicknesses of 30, 40, 50, and 60 ML (which corresponded to 85, 113, 142, and 170 A, respectively). As one can see from the Figure 2.9(a), for the smallest spacer (30 ML) only the emission peak from the LQDs in the second layer was present in the PL spectrum. The PL peak from the SQDs in the seed layer started to appear gradually as the spacer thickness increased from 30 to 60 ML. This behavior of the PL spectra indicated that tunneling probability was high with 30 ML spacer: most of the carriers tunneled from the SQD to the LQD and recombined there, giving rise to a single PL peak. With increasing spacer thickness tunnel coupling decreased due to reduced carrier wave function overlap, which eventually resulted in an independent recombination of excitons in both QD layers. The carrier dynamics in such a system (shown by arrows in

14

Figure 2.9(b)) was probed by the time-resolved PL measurements using pulsed laser excitation. The PL transients (curves that show the PL intensity decay after the laser pulse as a function of time) recorded at the emission energies of SQDs and LQDs of the 30 ML sample (Figure 2.10(a)) clearly indicated that the LQDs continued to receive carriers for some time after the laser pulse (rise part of the LQD PL transient) and showed longer PL decay time, while the PL signal from the SQDs started to rapidly decay

10
T3

(b) ^ 6 0 ML W50ML

10"1 "< -

<8 ^ ^ 4 0 ML 1
500

' %

30ML^*j 1 i Zr^iir
1000 Time (ps) 1500

Figure 2.10 (a) Normalized PL intensities detected at the emission peaks of the SQDs and LQDs of the structure with 30 ML spacer thickness as a function of the time delay after the laser pulse, (b) Normalized PL transients detected at the emission peak of the SQDs for the samples with different spacer thicknesses [47]. 15

immediately after the laser pulse. The comparison of PL transients from the SQDs for samples with different spacer thicknesses is depicted in Figure 2.10(b). The PL decay time TPL can be extracted from the slope of the exponential decay of the PL transients. The PL decay time (measured experimentally) is related to the transfer (or tunneling) time rtr from the SQDs to the LQDs by the following equation:

P L

= ^ - , s+Ttr

(2-2)

where rs is the radiative lifetime in the SQDs. The rs can be obtained from the PL transients measured on the reference sample containing solely SQDs grown under the same conditions as the SQDs in the double layer QD samples. The tunneling time between two layers of the QDs depends exponentially on the spacer thickness [45,47]. This dependence is described by the Wentzel-Kramers-Brillouin (WKB) approximation:

rlr cc exp 2* Here, d

W^

(2.3)

is the spacer thickness, rn is the effective mass in the spacer material, Fis the

conduction band discontinuity, and Es is the energy of the lowest quantum state in the SQDs. Several studies [45,47] have shown that the tunneling times in the asymmetric double QD system were longer than those in the asymmetric double QW system. This can be explained by the carrier localization in the QDs and the possibility of tunneling only between the two neighboring dots, whereas in the double QW system, carriers can tunnel at any location in the QW and find available empty states in the second QW. Using near-field optical microscopy [49] or micro-PL techniques together with a shadow mask that has an aperture with diameter of < 1 um [50,51,52], or with processing 16

of the low density QD samples using an electron beam lithography and dry etching [53], mesas are left with supposedly only one QD pair, allowing the study of coupling in a single QD molecule. Exploring single QDs biased by an external electric field gives an opportunity to trace resonant tunneling of single excitons, electrons, and holes. Exciton energy dependences on the electric field applied along the growth direction calculated for the system of asymmetric pair of large InAs QD (4 nm height) and small InAs QD (2.5 nm height) with 4 nm GaAs spacer are shown in Figure 2.11 [51]. The excitons are

100

TopQD siiii

5
E>
a>
c CD

50 -

4 nm

y^WA
0

Bottom QD

c o o a 100 .> JS 50 0

X^s|_^/

(b)

TopQD 4 nm

an
10Y0 10*

/fix"
100

-50 0 50 Electric Field (kV/cm)

B 100

Bottom QD

Figure 2.11 (a) and (b) represent the calculations of exciton energies as a function of electric field in the two types of single asymmetric QD molecule schematically shown on a right-hand side. "B" and "T" denote horizontal lines of a direct exciton transitions in a bottom and top QDs, respectively. Yellow circles indicate the anti-crossings observed in PL experiment [51].

17

labeled asehBlTXQ, where eg and er (hB and hi) indicate the number of electrons (holes) located in the bottom ("B") and top ("T") QDs, Q is the total charge of the exciton. Direct excitons, both electron and hole are in the same dot, ( ^ J 0 and
QIX)

only slightly

depend on the electric bias, whereas indirect excitons, electron and hole are in different dots, (l^X0 and ^X ) exhibit strong linear electric field dependence (Stark shift). When two levels (either electron or hole) in two QDs align the resonance occurs, which appears as anti-crossing in the exciton energy diagrams in Figure 2.11. If the larger QD sits on top of the small QD (Figure 2.11(a)), the electron resonance occurs at positive bias and the hole resonance occurs at negative bias. The situation is reversed for the opposite structure (Figure 2.11(b)). The magnitude of the anti-crossing splitting, a measure of the coupling strength, is related to the resonant tunneling time and depends on the spacer thickness and the effective mass of the particles that tunnel. In case of electron level resonance, the anti-crossing splitting is much larger (one order for the particular case shown in Figure 2.11) than that of the holes [50,51], which reflects the effective mass difference between electrons and holes. Study of nonresonant tunneling times in a single pair of asymmetric double QDs as a function of spacer thickness revealed that at large spacer (-20 nm) there was practically no tunneling, at intermediate spacer (-10 nm) mostly electron tunneling took place, and at small spacer (-2.5 nm) both electrons and holes tunneled from small to large QDs [53]. Transfer times showed an exponential dependence on the spacer thickness [53], which agreed with the WKB approximation. In temperature range between 4 K and 25 K, thermally induced (phonon-assisted) transfer of holes between two neighboring QDs separated by 5 nm spacer has been reported [52]. 18

Thus, double layer QDs have very similar tunneling properties to the double QW structure. However, these two systems possess quite different two-dimensional (2D) and zero-dimensional (OD) densities of states. For the QDs, two scenarios, ensemble and single QDs, should be considered as well as the possibility of lateral and vertical coupling. Another big QD research field, spin interaction and spintronics applications in coupled QDs (see for example [54,55]), was not touched upon in this review since this dissertation research does not include any direct information about the interaction of the electron and hole spins in the QD-QW system.

2.3 Quantum dot - quantum well system Now that the coupling in the double QWs and double QDs has been reviewed, consider a hybrid structure that combines two different nanostructures - QD-QW system, as shown in Figure 2.12. This system offers a unique opportunity to explore and utilize the coupling between OD and 2D quantum states in the QDs and QW, respectively. Double QD system

r*\ rr\ ST\

2LX S

QD-QW system

Double QW system

Figure 2.12 Double layer nanostructures and a hybrid QD-QW nanostructure with arrows defining separation distance between the layers. 19

Resonant and nonresonant tunneling from a QW to a QD have been considered theoretically [16,56]. In the resonant case when the QW ground state is closely aligned with one of the QD excited states, the Bardeen transition-probability approach was used to calculate the tunneling times of electrons and holes depending on the barrier thickness [56]. It is seen from Figure 2.13 that according to this model electrons tunnel from a QW to the QDs much faster than holes (~ 4 orders of magnitude difference at 10 A spacer).
10- 5
^ ^ ^ ^ ^ ^ ^ ^ ^ T ^ m - p T ^ T - p - T

icr
^V

lO"7

^
< D
+->

icr8
10-iu

Hole tunneling

E 10"y
C

"53 10" 11 c c 10" 12 3 hu

Electron tunneling

10"

10" 14 10" 15

10

20 30 Barrier width (A)

40

50

Figure 2.13 Calculated resonant tunneling time for electrons and holes that tunnel from the QW ground state to the QD 1st (long dash line) and 2nd (dotted line) excited states (solid line is a total tunneling time) depending on the separation barrier thickness [56]. In the nonresonant case when energy levels in the QW and QD are not aligned, LOphonon-assisted and Auger-assisted tunneling of electrons and holes can take place [16]. For the former mechanism, that is most effective at low excitation powers, the energy difference between the QW and QD ground states must be equal or larger than the energy of one LO phonon which carrier emits while tunneling from the QW to the QD. The latter mechanism does not have such a limitation on the energy difference between the QW and 20

QD levels and dominates the tunneling at high excitation powers. In Auger-assisted tunneling, two carriers in the QW take part in the process: one carrier transfers part of its energy via Coulomb repulsion to another carrier exciting it to a higher state in the QW and tunnels to the QD state. Tunneling time for both phonon- and Auger-assisted mechanisms can vary between a few and few hundreds of picoseconds depending on a spacer thickness [16]. Several QD-QW structures slightly different by design and material composition have been studied experimentally using steady-state and time-resolved PL. The PL and PL excitation (PLE) spectra of the QD-QW structure are shown in Figure 2.14(b) along QD

QW
i

T = 8 K

lex = 0.1W/cm2

1.2 1.3 1.4 1.5 Photon energy (eV) Figure 2.14 (a) PL spectra of InAs QDs and Ino 3Gao 7As QW reference samples, (b) PL and PLE spectra of the QD-QW sample. PLE signal was detected at the QD and QW exciton ground state energies (shown by vertical down-arrows). Iex = 0.1 W/cm2; T= 8 K [57]. 21

with the PL spectra of InAs QDs and 7.5 nm thick Ino.3Gao.7As QW reference samples (Figure 2.14(a)) grown under the same conditions as the QD-QW sample [57]. The QDs and QW separated by a 4.5 nm GaAs spacer were designed in such a way that the lowest QW electronic state was located above the QD ground state by more than one LO-phonon energy. The PLE spectrum measured at the QD PL maximum clearly showed QW to QD coupling by a spike and abrupt absorption edge at energy position that coincided with the QW PL emission (Figure 2.14(b)). Another evidence to the favor of the carrier transfer from the QW to the QDs was the excitation intensity evolution of the PL spectra depicted in Figure 2.15. At the lowest excitation intensity, only the QD PL peak was present in the

1.15

1.20 1.25 1.30 1.35 Photon energy (eV)

Figure 2.15 PL spectra (normalized with respect to the QD PL maximum) of QD-QW sample at different excitation intensities. The spectra are shifted vertically for clarity. Inset shows the integrated PL intensity ratio IQD/IQW as a function of excitation intensity. T= 10 K [57]. 22

spectrum. As excitation intensity increased the QW PL peak started to gradually grow in intensity and eventually dominated the spectrum at the highest excitation intensities. At low excitation intensity the electrons and holes initially trapped by the QW were effectively transferred to the QDs with emission of LO-phonons in the process. With increased excitation power more QDs became populated with excitons and, due to the Pauli exclusion principle, could not accept additional carriers from the QW which led to the appearance of the QW PL peak (Figure 2.15). Based on the experimental results, it was suggested that the tunneling in this particular system occured in several stages. First, an electron tunneled to the QD while a heavy hole remained for some time in the QW, temporarily forming an indirect exciton with an electron in the QD. In the next stage, a heavy hole tunneled to the QD restoring a direct exciton with an electron in the QD [57]. Time-resolved PL results for the same QD-QW sample fitted by the rate equation model that included QD state filling gave the tunneling time of 1.12 ns [58]. Another timeresolved PL study was done on a QD-QW system that consisted of 7 nm thick In03Gao7As QW and Ino6Ga04As QDs separated by a 2 nm GaAs spacer [59]. This study also indicated the significance of a state filling effect in tunneling process at high excitation powers and the nonresonant tunneling time was determined to be 200 ps. Systematic investigation of the QD-QW tunnel injection structure as a function of a separation spacer thickness was performed by Talalaev et al. [60]. A set of the QD-QW structures with different spacer thicknesses was grown in an opposite order: first, InAs QDs were formed, then the QDs were covered by a GaAs spacer, and 11 nm thick Ino lsGao 85AS QW was grown on top followed by a GaAs cap layer (see Figure 2.16(a)).

23

Figure 2.16 (a) Schematic illustration of the inverse QD-QW structure studied in [60]. (b) Formation of nanobridges at spacers thinner than 5 nm. From high resolution transmission electron microscopy measurements it was determined that in samples with a small separation barrier the QDs were touching the QW forming nanobridges defined as areas that contained more than 15% of Indium (Figure 2.16(b)). The coupling between the QW and QDs was confirmed by the PLE measurements. Tunneling times obtained from the time-resolved PL experiments agreed with the WKB approximation within the barrier thickness range from 10 to 5 nm. The deviation from the WKB approximation was observed below 5 nm which was attributed to the effect of nanobridges that connected QDs to the QW [60]. Nonradiative Forster-type energy transfer, which involves Coulomb interactions as oppose to the wave function overlap, from the 3 nm thick InGaN QW to the CdSe/ZnS core/shell nanocrystals through the 3 nm GaN barrier in a hybrid QD-QW structure was explored experimentally [61] and theoretically [62]. Electron-hole pairs can be created in the QW either optically or electrically (as in Figure 2.17(a)). The PL emission of the QW is in a strong absorption range of the CdSe nanocrystals which favors a dipole-dipole coupling between the QW and QDs. The study showed that the energy transfer was fast enough to compete with recombination processes in the QW resulting in a more than 50%

24

Figure 2.17 (a) Proposed hybrid light emitting QD-QW device that consists of epitaxial InGaN QW sandwiched between GaN barrier layers and a monolayer of densely packed CdSe/ZnS QDs coated with organic molecules placed on top of GaN barrier. Metal contacts on the top and the bottom are used to electrically inject carriers in the QW. ET denotes energy transfer, (b) Schematic illustration of the energy transfer, carrier relaxation and recombination in the hybrid QD-QW structure [62]. of energy transferred to the QDs. Resonantly created electron-hole pairs in the QDs quickly relaxed to the ground state and recombined radiatively (see Figure 2.17(b)). The effectiveness of the energy transfer depended on the temperature and the carrier concentration in the QW. The emission wavelength of the nanocrystals was determined by their geometrical sizes, which can be easily tuned in a wide range. 25

2.4 Strain-induced QDs in a QW Self-assembled epitaxial QDs grown by Stranski-Krastanov mechanism, i. e. formed due to the strain relaxation of the lattice mismatched materials, create a strain field that spreads around a dot in the surrounding matrix. The larger the lattice mismatch and the size of the QD, the greater the strain and the further it extends. Now, if the QW is placed in a close proximity to the QD (Figure 2.18(a)), the strain deformation potential caused by the dot will induce a lateral 2D confinement in the QW (Figure 2.18(b)). The QW itself provides confinement in a third (vertical) dimension, hence, resulting in a 3D carrier confinement which is nominally required for a formation of a QD. Such QD is called strain-induced QD (SIQD) [63]. The strain field creates almost ideally parabolic

(a)

SIQD InAs Island InP barrier InGaAsP QW InP buffer InP subs.

(c)
QD1

0.70 0.75 0.80 PL energy (eV)

-i

0 60 120 Radial distance from island center (nm)

Figure 2.18 (a) Cross-section of the structure with SIQD; (b) lateral confinement potential created by the strain field in the QW plane; (c) typical PL spectrum of the SIQD; (d) potential profile as a function of a stressor dot - QW separation distance d [64]. 26

deformation potential with evenly spaced confined states that give rise to a distinct optical transitions in PL spectra (Figure 2.18(c)). The depth of a deformation potential well in the QW caused by the stressor dot depends on the distance from the stressor island to the QW (Figure 2.18(d)) [64] as well as the size (height) of the stressor island [65]. There are different material combinations suitable for building a system with large stressor island that causes SIQDs in a near QW. Coherent InP QDs with diameter of up to 120 nm can be grown on a GaAs surface; the SIQDs are formed in the InGaAs QW [66]. Both surface [64,65] and buried [67,68] InAs QDs with lateral sizes from 40 to 110 nm have been reported as suitable stressor islands for creating SIQDs. It was shown from the optical and transmission electron microscopy study that when small InAs QDs (40-60 nm in diameter) were used as stressors, well detectable SIQDs were formed only in the QWabove-the-stressor-dot configuration (Figure 2.19(a)), whereas for inverted configuration of the structure, shown in Figure 2.19(b), the formation of SIQDs was possible only for very small distances between the stressor and the QW [68]. Stressor islands made of GaSb were also used to create SIQDs in the GaAs/AlGaAs QW [69].

(a)

(b)

Figure 2.19 (a) Structure with GaAs QW above the stressor InAs QDs and (b) inverted structure with GaAs QW below the stressor InAs QDs [68].

27

Although SIQDs appear to be an ideal model system for the fundamental studies of various physical processes in the QDs and QWs, it may be difficult to find a practical application for these objects. There was one experimental result that may prove a possibility of an exciton storage in an InAs QD coupled to a SIQD [70].

2.5 Applications of QD-QW system Quantum dots with their discrete, delta-function-like transitions and emission wavelength tunable by simply varying the size of the QDs are considered as ideal candidates for building the micro-scale light emitting diodes (LEDs) and semiconductor lasers. Such small lasers could be integrated into the telecommunication schemes and future optical computing circuits. The possibility of the QD laser operation has been successfully demonstrated using InGaAs/GaAs QDs emitting at 1.3 urn at room temperature [71]. However, there were limitations to the efficiency of the QD lasers caused by a random QD size and shape distribution, low density, inefficient carrier collection and redistribution among the dots, room temperature thermal carrier escape and hot carrier relaxation. These problematic issues can be partially resolved if a QW is fabricated in a close proximity to the QD layer. In this case, the QW acts as an efficient carrier collector (due to its larger area and higher density of states) and reservoir that supplies carriers uniformly to all QDs via tunneling. The carriers that tunnel through the barrier to the dots are already "cold" and, hence, do not cause the problems related with the hot carrier relaxation. Also, the strain from the QW affects the growth of the QDs improving their uniformity and increasing density [72,73]. At high temperature, when thermally activated
28

carriers escape from the QDs, the QW recaptures some of those carriers and again injects them into the QDs increasing the probability of their radiative recombination inside the QDs [74]. It has been reported that QD-QW diode lasers based on single layer InP QDs (Figure 2.20(a)) resonantly coupled to the InGaP QW through a 2 nm tunnel barrier were capable of room temperature continuous-wave laser operation in visible range (630680 nm) and had a steeper current-voltage characteristic as compared to the same InP QD diode laser without a QW, as illustrated in Figure 2.20(b) [74,75].

(a)

(b)

Figure 2.20 (a) Schematic structure of a QD-QW diode laser, where CL is a cladding layer, WG is a waveguide, B is a barrier, LB = 20A and LQW = 20A. (b) Comparison of I-V characteristics of the QD and QD+QW diode lasers [75]. The efficiency of the carrier tunneling strongly depends on the relative position of the energy levels in the QW and QDs. Therefore, the QD-QW structure should be carefully engineered so that the tunnel coupled states are in resonance [74,75] or the energy difference between these levels is equal to the LO-phonon energy [76,77]. The coupled InAs QD - InGaAs QW lasers target the near-infrared operation region (at 1.3 and 1.55 urn) which is solicited by the fiber-optic telecommunication industry [76,77]. 29

The possibility of a white light generation based on the Forster type energy transfer from the epitaxially grown InGaN QWs to the core/shell CdSe/ZnS colloidal nanocrystals has been proposed and experimentally demonstrated [78]. The color of light generation in this type of LEDs can be tuned by changing the size of the colloidal QDs.

2.6 Summary From the above discussion, a QD-QW hybrid nanostructure makes a convenient system to explore coupling between different types of semiconductor nanostructures. The main coupling mechanism observed in this system is quantum tunneling of electrons, holes, and excitons. Effectiveness of tunneling strongly depends on the relative positions of energy levels in the QW and QDs as well as the tunneling barrier thickness. The most efficient tunneling occurs when the energy levels in the QW and QDs are resonantly aligned or the energy difference between the levels equals to LO-phonon energy. Another coupling mechanism that may play a dominant role in systems that consist of colloidal QDs coupled to an epitaxial QW is nonradiative Forster energy transfer that is governed by the Coulomb interaction of electron-hole pairs. The QD-QW hybrid structure can be successfully used in the LED and laser diode fabrication for emission in visible and nearinfrared (telecommunication range) parts of spectrum. Systematic study of the QD-QW system with different spacer and energy level configurations at various excitation conditions is still needed in order to clarify the contribution of different coupling mechanisms and carrier dynamics in the coupled semiconductor nanostructures.

30

III. SAMPLES AND EXPERIMENTAL METHODS

3.1 QD-QW sample design and growth For comprehensive study of the QD-QW system, a variety of high quality samples were to be grown. The structure was designed in such a way that the PL peaks from the ground states of the QW and QDs were spectrally separated from each other allowing several QD excited states to find room between the QW and QDs ground states. Such a design may prove useful for clarifying the role of QD excited states in tunneling processes of carriers and their further relaxation within the QDs. Furthermore, if the energy of emission from one of the QD excited states overlaps with the QW excitonic emission, the phenomenon of coherent tunneling can be thoroughly explored in such system. The strength of the tunnel coupling can be effectively varied by changing the thickness and/or height of the potential barrier between the QW and QDs. A series of QD-QW samples with one layer of InAs QDs and an InxGai_xAs QW separated by a barrier with variable thickness and height were grown using molecular beam epitaxy (MBE) Riber 32 system. In order to smooth the surface of semi-insulating GaAs (001) substrate, a 0.3 um thick GaAs buffer layer was first deposited at temperature of 580C. After that the temperature decreased to 530C and a 14 run thick InxGai_xAs QW with x = 0.15 was grown followed by a 30 s growth interruption. The thickness of the next layer, GaAs spacer, was varied between 1 and 20 nm in the main set of samples. In a number of samples an additional Al(Ga)As barrier with thickness of 1 to 4 nm was inserted in the middle of the GaAs spacer (see Figure 3.1(a)). Then, 2 monolayers of InAs were deposited on top of the GaAs spacer followed by a self-assembled QD formation.
31

(a)
GaAs cap layer InAs QDs AlAs or AIGaAs barrier s (optional) WL GaAs spacer

(b)
Surface InAs QDs

GaAs buffer layer

GaAs buffer layer

GaAs (001) substrate

GaAs (001) substrate

Figure 3.1 Schematic structure of QD-QW samples used for (a) PL measurements and (b) AFM studies.

The entire structure was capped with 50 nm GaAs layer. Samples containing only QW and only QDs were grown under the same conditions to be used as reference structures. An additional set of QD-QW samples was grown with variation of In composition (x = 0.07, 0.10, 0.15, and 0.18) in the QW, whereas the thickness of GaAs spacer was kept the same, 4 nm. Also, several QD-QW samples with different GaAs spacers were left uncapped for atomic force microscopy (AFM) studies as shown in Figure 3.1(b). A reference QD sample with an additional layer of uncapped QDs on top of the structure was used in AFM studies as well. The quality of each sample was tested after every growth by means of PL measurements and, if necessary, corresponding corrections were made in the parameters of subsequent growths. The summary of all samples used in this dissertation is given in the Table 3.1.

32

Table 3.1 Classification of the QD-QW samples by the barrier type and thickness as well as by In content in the QW. Each shaded cell represents one sample. Reference Reference lno.i5Gao85As In As QDs QW

Dot-Well separation
d
s P

1.5 nm

2nm

3 nm

5 nm

8 nm

11 nm

15 nm

20 nm

GaAs spacer AlAs 1 nm AlAs 2 nm AlAs 3 nm AlAs 4 nm AIGaAs 1 nm AIGaAs 2 nm Uncapped QDs Indium content in QW Reference QWs QD-QW dSD = 4 nm Figure 3.2(a) illustrates the comparison of PL spectra from the reference InAs/GaAs QD sample (upper part), reference Ino.15Gao.85As/GaAs QW sample (middle part), and representative QD-QW sample with 11 nm thick GaAs spacer (bottom part) taken at low excitation intensity and T= 10 K. The PL peaks originated from the ground state exciton radiative recombination in the QW at 1.345 eV and in QDs at 1.145 eV. 33 7% 10% 15% 18%

AlAs Barrier. r V. 1 -4 nm | | QDs In As

1.1 1.2 1.3 1.4 Photon Energy (eV)

i i L J

Figure 3.2 (a) PL spectra of reference QD sample, reference QW sample, and QD-QW sample with dsp = 11 nm measured at low excitation intensity 7exc = 2xlO"57o (/o = 1000 W/cm2, T= 10 K). High excitation (7eXc = 10/0) PL spectrum of reference QD sample shown by dashed line, (b) Electron energy band diagram of the QD-QW system based on the growth parameters, TEM and PL results. Recombination and relaxation processes are shown by straight and wavy down-arrows, carrier transfer process is shown by a horizontal wavy arrow [79]. It is easy to see from this comparison that the PL peak positions from the QD-QW structure coincided within 5% with the PL peaks from both reference samples. The PL spectrum of the reference QDs taken at high excitation power shown by dashed line revealed additional emission peaks that identified the transition energies of the QD excited states. It is shown by vertical dashed lines that the QW emission coincided with the third excited state of the QDs. Thus, in this structure the resonant tunnel coupling should be observed. According to the growth parameters, TEM, and PL data the energy scheme of the conduction band (CB) and valence band (VB) for such hybrid QD-QW

34

structure can be suggested as depicted in Figure 3.2(b). The QDs have at least 4 bound states (E, E1, E2, and E3), whereas the QW has two states. One state is also present in the wetting layer (WL). The QW ground state is closely aligned with the third excited QD state allowing the resonant tunneling of carriers through the GaAs (AlAs) barrier.

3.2 Structural characterization of QD-QW samples For successful study of tunneling processes, knowing the exact parameters such as distance between the nanostructures is crucial. At a nanometer scale, even a small deviation from the nominal dimensions can cause a great difference in physical properties. It is also very important for comparison of many samples to be valid that all structural features of each sample be identical except for the one varied intentionally. The size of the QDs (it is height in case of MBE-grown self-assembled QDs) determines the energy position of the QD PL peak, given the chemical composition of the QDs remains the same. The size distribution of the QDs is reflected by the width of the QD PL peak. The AFM 1 um x 1 um images and corresponding QD size distributions of (a) uncapped reference QD sample and uncapped QD-QW samples with (b) 3 nm, (c) 5 nm, and (d) 8 nm spacers are shown in Figure 3.3. QD height and density extracted from the AFM images are collected in Table 3.2. Table 3.2 Structural data obtained from AFM analysis. QD Height (nm) Sample QD Density (cm-2) Reference QDs QD-QW, d sp = 3 nm QD-QW, d sp = 5 nm QD-QW, d sp = 8 nm 4.5 0.5 5.0 0.5 4.5 0.5 5.0 0.5 (1.220.05)x10 1u (1.150.05)x10 1u (1.11 0.05)x10 lu (0.860.05)x10 1u

35

15 00 nm

0.00 nm

QD height (nm)
15.00 nm

(b)
w

o
O 4

0 00 nm

1 2

Jl3

QD height (nm)
15 00 nm

(C)

o O

0.00 nm

IIIIII
1 2

llllll 7

QD height (nm)
15.00 nm

0.00 nm

1 2

QD height (nm)

Figure 3.3 1 um x 1 um AFM scans and QD size (height) distributions of (a) reference QDs and QD-QW samples with (b) 3 nm, (c) 5 nm, and (d) 8 nm GaAs spacer between QW and QDs.

36

Structural analysis to examine the quality of interfaces and verify the thicknesses of the QW and the spacers between the QW and QDs was done by the cross-sectional transmission electron microscopy (TEM) measurements using a FEI Titan 80-300 TEM equipped with an image Cs-corrector (CEOS) and an image filter (Gatan Tridiem). All samples were prepared using standard procedure which included mechanical polishing, dimpling (using Fischione Dimpling Grinder 200), and low-angle Ar+ ion-milling (using Fischione Ion Mill 1010). Two-beam or multi-beam conditions at one of the (002) reflections were used to carry out diffraction contrast imaging. For high resolution TEM (HRTEM) measurements (magnification up to 380000x), the sample orientation along the [110] axis was used. Bright-field cross-sectional TEM images of QD-QW samples with 15nm and 8 nm thick GaAs spacers are depicted in Figure 3.4 (a) and (b), respectively. One can clearly distinguish regions with QDs, WL, QW, and GaAs spacer. The change in contrast seen in the two images taken under the same conditions is related to the thickness differences in the different areas of the sample. QD-QW samples with smaller features were examined by the HRTEM in order to check the quality of interfaces and precisely determine the QW and GaAs spacer (or AlAs barrier) thicknesses. Figures 3.4 (c) and (d) illustrate HRTEM images taken from the QD-QW samples with 2 nm AlAs barrier embedded in 8 nm GaAs spacer and with just 2 nm GaAs spacer, respectively. The QDs have the shape of truncated cones with average height of 5 nm and diameter of -20 nm. The thickness of the QW was found to be 14 0.5 nm. All spacer and barrier thicknesses were within 0.5 nm error of the nominal dimensions set up during the MBE growth. The TEM image along with the schematic structure of the QD-QW sample with different indium content (x = 0.18) and ^sp = 4nmis shown in Figure 3.5.

37

QD

Figure 3.4 Cross-sectional TEM images of QD-QW samples: (a) and (b) were taken in bright-field mode from samples with 15 nm and 8 nm GaAs spacers, respectively; (c) and (d) are HRTEM images from sample with 2 nm AlAs layer symmetrically embedded in 8 nm GaAs spacer and from sample with 2 nm GaAs spacer, respectively. The QW composition was Ino.15Gao.85As. White dashed line in (d) defines QD contour [79].

In As QDs

Figure 3.5 (a) Multi-beam bright-field TEM image of the QD-QW structure with the QW indium content of x = 0.18 and GaAs spacer thickness of dsp = 4 nm. (b) Schematic structure that mirrors the TEM image.

38

Based on all TEM images taken from many QD-QW samples with different spacer thicknesses, an important fact was confirmed. The strain field caused by the InAs QDs did not penetrate more than 2 nm down into the GaAs spacer or further into the QW. Thus, the possibility of formation of strain-induced QDs inside the QW was excluded in this case. Overall, from the AFM and TEM analysis, the following conclusion was made: the MBE-grown QD-QW samples were of high quality and well fitted for the study of coupling effects in such systems.

3.3 Continuous-wave PL and PL excitation measurements Most of the experimental studies in this dissertation research were done using various PL methods. Figure 3.6 depicts the experimental setup used for continuous wave (cw) PL and photoluminescence excitation (PLE) measurements. The sample was mounted inside the closed-cycle helium optical cryostat (Janis) which allowed varying the temperature within the range of 10 to 300 K. The cryostat was designed in such a way that the sample was placed in a close proximity (< 1 cm) to the window, so that the short focus objective could be used for illumination and collection of the signal. For the excitation above the GaAs bandgap, a 532 nm line of the frequency-doubled neodymium doped yttrium aluminum garnet (Nd:YAG) laser (Coherent Verdi-VIO) was used. For the excitation below the GaAs bandgap for example resonant excitation in the WL or QW states, a tunable Tksapphire laser (Coherent MIRA 900) was used in the cw mode. The excitation power was varied by a set of neutral density filters over the range of (10~7-102) mW. The laser beam was focused on a sample surface using a near-infrared objective 39

Figure 3.6 Schematic diagram of the experimental setup for PL and PLE measurements. In the figure: M is mirror, FM is flip mirror, BS is beam splitter, LP Filter is long pass filter, ND Filter is neutral density filter, CCD is InGaAs photodiode detector array.

(Mitutoyo NIR HR) with 50x magnification to a spot with diameter of ~20 urn. The PL signal was collected with the same objective, dispersed by a 0.5 meter imaging triple grating monochromator (Acton SpectraPro 2500i), and detected by a liquid nitrogen cooled InGaAs linear photodiode array (Princeton Instruments OMA V). In case of PLE measurements, the excitation wavelength from the Ti: sapphire laser was tuned from 750 nm to 970 nm with increments of 1 nm while recording the intensity of the PL signal at a fixed detection wavelength of either QD or QW exciton ground state emission. The output power of the laser was carefully monitored and kept at the same level of 500 mW throughout the tuning range.

40

3.4 Time-resolved PL measurements The dynamics of the carrier excitation, distribution, tunneling, relaxation, and recombination in semiconductor nanostructures can be traced by time-resolved PL (TRPL) measurements where the time evolution of the PL intensity is being recorded. The experimental setup for TRPL is shown in Figure 3.7. The scheme was very similar to the cw PL measurements with the difference in excitation and detection. The Ti:sapphire laser was tuned to a wavelength X = 750 nm and switched to a mode-locked regime. The laser produced 2 ps pulses at a repetition rate of 76 MHz. The excitation density on a sample surface was varied between 5><109 and 9xl0 13 photons/(pulse x cm2) which corresponded to peak excitation intensities in a range from 7><102 to 1.3><107 W/cm2.

Figure 3.7 Schematic diagram of the experimental setup for TRPL measurements. In the figure: M is mirror, FM is flip mirror, BS is beam splitter, SP is semitransparent plate, T is trigger, LP Filter is long pass filter, ND Filter is neutral density filter. 41

The PL emission collected by the objective was focused to an entering slit of the spectrometer (Bruker) with a synchroscan streak camera (Hamamatsu C5680) attached to it. The streak camera was equipped with an infrared enhanced SI cathode, which allowed detecting of a reliable signal up to the wavelength of -1150 nm (QD ground state emission from the QD-QW samples was around 1100 nm). The overall time resolution of the system was 15 ps. The result of the measurements by this technique was either the PL transient, a curve that shows the PL intensity decay as a function of time at a given wavelength, or the PL spectrum at a given time delay. By fitting the PL transients with exponential function it was possible to extract the exciton recombination time in the semiconductor nanostructure under investigation.

3.5 Summary Several sets of the QD-QW samples with various spacer thicknesses and barrier heights as well as different Indium concentrations in the QW were grown by MBE. Reference samples that contained only QW or QDs were grown as well. The crystal quality of the samples and all geometrical parameters such as QD height, diameter and density, QW thicknesses, and spacer thicknesses were determined from the optical, AFM, and TEM characterizations. Experimental setups for the cw PL, PLE, and TRPL measurements were described in details.

42

IV. EFFECT OF QD STATE FILLING ON TUNNELING PROCESSES IN QD-QW NANOSTRUCTURES

The dynamics of carrier transfer from the QW to the QDs, and then of carrier relaxation within the QDs, define an efficiency of lasing from the QD ground state. It is important that the carrier relaxation from the higher QD excited states occurs at a faster rate than the tunneling from the QW to the QDs [80]. This fast relaxation helps to avoid the state filling in the QDs that slows down the tunneling process [81]. In this chapter, an attempt to answer the questions of "What coupling mechanism does play the major role in the InAs QD - InGaAs QW hybrid nanostructure?" and "How does the QD state filling influence the carrier transfer time?" was made by means of continuous-wave and time-resolved PL experiments. To determine the mechanism of carrier transfer, a variation of the separation barrier thickness and height as well as tuning the energy of the coupled QW and QD states were realized. Two sets of the QD-QW samples were grown with changing profile configurations (includes height and thickness) of the Ga(Al)As barrier and changing Indium content in the InGaAs QW. To explore the effect of the QD state filling on the carrier tunneling, continuous-wave and time-resolved PL spectra of the QD-QW samples were measured as a function of laser excitation power. Experimental results were analyzed using the rate equation model, and parameters of the carrier dynamics between the QW and QD coupled states and between the excited and ground states inside the QD were determined.

43

4.1 Effect of GaAs spacer thickness In this study, the QD-QW structure was designed in such a way that the QD 3 rd excited state emission energy (EgD ~ 1.350 eV) overlapped with the QW ground state exciton emission (EQW ~ 1.345 eV), as shown in Figure 3.2(a). For this arrangement of energy levels in the QW adjacent to the QDs, fast carrier tunneling of resonant type was expected from the QW ground state to the QD 3 rd excited state (Figure 3.2(b)). The intensity of the QW PL peak in the spectra of the QD-QW structures was anticipated to be very sensitive to the separation distance between the QW and QDs, thus indicating the coupling strength in the hybrid QD-QW system. Indeed, a clear demonstration of the QW PL peak sensitivity to a variation of the GaAs spacer thickness can be seen in Figure 4.1 that shows the PL spectra of six QD-QW samples with GaAs spacer thicknesses, dsp, ranging from 2 to 20 nm taken under different excitation intensities 7exc- All spectra were normalized with respect to the maximum of the QD ground state PL peak and vertically shifted for convenience. When the GaAs spacer was 20 nm thick, the evolution of the PL spectra of both QW and QDs with increasing excitation intensity were essentially the same as those of the reference QW and QD samples. The PL peaks from the lowest states of both QW and QDs were detected at the lowest excitation intensity (Figure 4.1(a)), which indicated that the QW and QDs were practically decoupled at 20 nm GaAs separation and acted as independent nanostructures. The PL peaks from the QD excited states gradually appeared as the excitation intensity increased. When GaAs spacer was reduced to 15 nm, the QW peak was barely detectable under the low excitation conditions and became stronger only with increase of excitation power (Figure 4.1(b)). Further reduction of the spacer thickness resulted in an increase of 44

4xnrv
2 2x105/

1.1 1.2 1.3 1.4 Photon Energy (eV) cfsp = 20 nm

1.1 1.2 1.3 1.4 Photon Energy (eV) dsp = 15 nm

1.1 1.2 1.3 1.4 Photon Energy (eV) dsp = 11 nm

~LT QW

QDU

~LT QW

QD

~LT QW

QDU

1.1 1.2 1.3 1.4 Photon Energy (eV)

1.1 1.2 1.3 1.4 Photon Energy (eV)

1.1 1.2 1.3 1.4 Photon Energy (eV)

unr
QW UQD

cfsp = 8 nm

i_nr
QW

d SD = 5 nm SP
J

uir
QW UQD

dsp = 2 nm

UQD

Figure 4.1 Normalized PL spectra of the QD-QW samples with (a) 20 nm, (b) 15 nm, (c) 11 nm, (d) 8 nm, (e) 5 nm, and (f) 2 nm thicknesses of GaAs spacer measured at different excitation intensities denoted on each curve. The spectra are vertically shifted for clarity. Diagrams of the conduction band profile illustrate the GaAs spacer thicknesses of corresponding samples. 7o= 1000 W/cm2, T= 10 K [79].

45

the QD-QW coupling strength, which was reflected in the behavior of the QW PL peak. As shown in Figure 4.1(c)-4.1(f), the QW PL signal emerged at higher excitation intensities Iexc for smaller spacers dsp. This behavior correlated with the carrier tunneling that occurs faster for smaller spacers, whereas the recombination rate in the QW remained the same for all spacer thicknesses. The increasing supply of carriers that tunneled through smaller GaAs spacers made the PL peaks from the QD excited states appear earlier (at lower excitation intensities) as compared to the reference QD sample and the QD-QW samples with thick GaAs spacers (15 and 20 nm), as can be seen from Figure 4.1(f). When the excitation power became high enough to populate the 3rd excited state of the QD and block the tunneling channel, the QW PL peak started to show up in the spectrum. Thus, intensity dependent state filling of the QDs controls the tunneling channel and, hence, the appearance of the QW PL signal. The PL results in Figure 4.1 qualitatively show the increasing coupling strength tendency with decrease of the GaAs spacer thickness (or separation distance between the QW and the QDs) in the QD-QW hybrid nanostructures.

46

4.2 Effect of AIAs and AlGaAs barrier height The coupling between the nanostructures can be also affected by the height of the separation barrier. To determine the influence of the barrier height on the tunneling efficiency in the QD-QW system, the structures were designed with a complex barrier ds that consisted of the GaAs spacer with fixed total thickness dsp which included a symmetrically incorporated barrier (AIAs or Alo.3Gao.7As layer) of variable thicknesses, as schematically shown in the right part of Figure 4.2. The effective barrier d-& was

4 nm

d = cL
sp GaAs

+ dA
AIAs 0

= 8 nm

/
exc

= 0.004 /

"E
3

n
>_ "re

AI* <nm)

c
_i Q.

1.1

1.2

1.3

Photon Energy (eV)

"U^

8 nm-

Figure 4.2 PL spectra of the QD-QW samples with constant 8 nm thick spacer dsp that contained AIAs effective barrier JAIAS of 1 -4 nm thickness inserted in the GaAs layer as schematically shown on the right. All spectra were taken at excitation intensity Iexc = 0.004x/0, I0 = 1000 W/cm2, r=10K[79].

47

defined by the following expression:

dB = [spJV(z)-Edz,

(4.1)

where E is the starting level energy and V(z) is the profile of the potential barrier. The PL spectra of the QD-QW samples with dsp = 8 nm and embedded AlAs layers of variable thicknesses from 1 to 4 nm measured at the same excitation intensity, along with the schematic representation of their conduction band barrier profiles, are depicted in Figure 4.2. There was significant carrier tunneling that suppressed the QW PL peak in the QD-QW sample with pure 8 nm GaAs spacer. Inclusion of just a 1 nm thick AlAs barrier immediately "turned on" the QW PL signal, while increasing of the AlAs barrier thickness up to 4 nm completely cut off the tunneling channel between the QW and QDs practically making them independent, non-interacting nanostructures. It is clearly seen from Figure 4.3 that the integrated PL signal of the QDs dropped exponentially (shown

<t
c 250
3 .Q
>-

I
\ \ ( sp \\ \
X X X

I
= 8 nm
AlAs

GaAs

+d

200 150 100 1

c
-J Q. Q

X X

'

- ,>

I
.

PL Signal of Ref QD

.
1
r

) C

I ^5

cL. AlAs (nm)


Figure 4.3 Dependence of the integrated QD PL intensity on the AlAs barrier thickness. The size of the circles reflects an experimental error, red line gives the value of the integrated PL intensity of the reference QD sample [79].

48

with dashed line) as the AlAs layer thickness increased and approached the reference QD PL signal at C/AIAS = 4 nm. This meant that with 4 nm of AlAs the QDs did not receive any additional carriers from the QW. Another example of how the incorporated barriers affect the tunneling processes are shown in Figure 4.4. Here, QD-QW samples with a fixed 3 nm thick spacer were used. Figure 4.4(a) demonstrates that the carrier transfer was reduced by increasing the barrier height. When the QW and QDs were separated just by 3 nm thick GaAs spacer, the QW signal was not observed in the PL spectrum (curve 1). Embedding of 1 nm thick Alo.3Gao.7As barrier in the middle of the spacer resulted in the increase of the QW PL intensity by one order of magnitude (curve 2). When the height of the effective barrier was increased even more by incorporating a 1 nm thick AlAs layer, it gave rise to two more orders of the QW PL intensity increase (curve 3). Almost the same result can be achieved by making the Alo.3Gao.7As effective barrier thicker. As shown in Figure 4.4(b) (curve 3), adding 1 nm of thickness to the effective barrier results in about 2.5 orders increase of the QW PL peak. Thus, by varying the aluminum content in the embedded effective barrier one can effectively control the coupling strength between adjacent nanostructures while keeping the separation distance constant. The results of these PL experiments with various barrier profiles indicated that in the QD-QW system the quantum tunneling is the main coupling mechanism and that the electromagnetic and polariton coupling could be neglected in further consideration [79].

49

QW

QD

"LT*
~LT
I
1.1 1.2 1.3 1.4

r3nm

Photon Energy (eV)

QD

"Li

1.1

1.2

1.3

1/ 1

3 nm

Photon Energy (eV) Figure 4.4 PL spectra of the QD-QW samples with constant 3 nm thick spacer dsv that contained AlAs or Alo.3Gao.7As effective barrier t/B- (a) and (b) demonstrate influence of the effective barrier height and width on tunneling, respectively. Corresponding conduction band energy profiles are shown on the right side. The spectra are normalized, vertically shifted and plotted on a logarithmic scale. I0 = 1000 W/cm2, T= 10 K.

50

4.3 PL decay measurements in QD-QW samples The time-resolved PL measurements allowed one to study the dynamics of the carrier tunneling and relaxation processes in the QD-QW nanostructure. Figure 4.5 shows the time evolution of the reference QD PL spectrum after the excitation laser pulse. At 0 ps delay, the TRPL signal reflected all the features present in a cw PL spectrum at high excitation intensity (shown by the red solid line in Figure 4.5) with a strong WL peak, which quickly decayed in first 100 ps. The QD ground state exhibited the slowest

800

900

1000

1100

Wavelength (nm)
Figure 4.5 Time-resolved PL spectra of the reference QD sample taken at various delay times (from 0 to 1000 ps) after the laser pulse. A cw PL spectrum (red solid line) is shown for comparison. TRPL spectra are vertically shifted for clarity. X,exc= 750 nm, T= 10 K.

51

recombination rate, whereas every higher QD state decayed faster than the previous (lower) state. The time-resolved PL spectra recorded at a number of delay times of the QD-QW sample with 3 nm GaAs spacer, along with the cw PL spectrum recorded from the same sample at high excitation intensity (red solid line), are shown in Figure 4.6. The strong emission from the QW ground state and the WL dominated the spectrum at 0 ps delay. A part of the spectrum that consisted of PL peaks from three QD states remained practically unchanged (peak intensity decayed less than 10%) up to the delay time (around 1000 ps)

Figure 4.5 Time-resolved PL spectra of the QD-QW sample with 3 nm GaAs spacer taken at various delay times (from 0 to 1000 ps) after the laser pulse. A cw PL spectrum (red solid line) is shown for comparison. TRPL spectra are vertically shifted for clarity. A.exc= 750 nm, T= 10 K. 52

when the QW PL peak vanishes. After that, the QD peaks decayed in a cascade fashion described above for the reference QDs. This was a real-time observation of the QW feeding the QDs via tunneling until it ran out of the carriers excited by the laser pulse. The PL intensities as a function of time, PL transients, were measured at energies of a QD ground state and QD excited states (denoted as E, E1, E2, and E3 in Figures 4.5 and 4.6) of the reference QD sample and the QD-QW sample with dsp = 3 nm (plotted in Figure 4.7). Nonresonant excitation above the GaAs bandgap (kexc = 750 nm) was used for the PL transient measurements with an excitation power of Ip = 1.3 x 10 5 W/cm2. A convex shape of the reference QD ground state (E) and 1st excited state (E1) PL transients (Figure 4.7(a)) indicates the presence of the state filling effects, which was well known for the QDs optically excited with high intensity [82,83]. In the presence of the proximal QW the PL transients exhibited two stages of decay (see Figure 4.7(b)). If in the case of the reference QDs optically excited carriers relaxed to the dot states only from the WL and GaAs barrier, then in the QD-QW system an additional portion of carriers tunneled to the QDs from the QW affecting the dynamics of relaxation processes in the QDs. In the first stage the QD states received extra carriers from the QW, which resulted in slower decay rates. When the QW was completely exhausted, the PL decay from the QD states became faster without additional supply (second stage). For each lower energy QD state the second stage started later because it still had additional carrier supply from the QD states with higher energies. The PL transient from the 3rd QD excited state (E3) was not possible to measure since its peak position overlapped with the QW ground state exciton emission energy and played a role of an acceptor state for the carriers that transferred from the QW.

53

1000

1500

500

1000

1500

Time (ps)
Figure 4.7 PL decay transients recorded at the energy positions of each QD state (E, E1, E2, and E3) for (a) reference QD sample and (b) QD-QW sample with 3 nm GaAs spacer. The pulse laser excitation at 750 nm had a peak intensity of Iv = 1.3x105 W/cm2 [79]. Similarly, at high excitation powers the QW PL transients became affected by the state filling and exhibited clear convex shapes (Figure 4.8(a)). The QW PL decay transients of the QD-QW sample with dsp = 5 nm shown in Figure 4.8(a) were fitted by a monoexponential decay (in order to determine the QW excitonic emission time) only at low excitation powers (Ip = 1.3 xlO4 W/cm2). Figure 4.8(b) demonstrates the dependence of the QW PL decay transients on the GaAs spacer thickness taken at low power. It was observed that for thin spacers the QW PL decay occurred at a faster rate than for thick spacers. As GaAs spacer thickness increased in the QD-QW samples, the QW PL decay time approached the reference QW PL decay time (see the slopes in Figure 4.8(b)). 54

100

200

300

Time (ps)
Figure 4.7 (a) QW PL transients of the QD-QW sample with 5 nm spacer taken at different intensities, (b) QW PL transients taken from the QD-QW samples with different GaAs spacer thicknesses and the reference QW sample at low excitation intensity (7P = 1.3><104 W/cm2) [79].

55

4.4 Rate equation model The experimental results presented in the previous sections of this chapter described mostly qualitative information about the state filling and tunneling processes in the QD-QW system. To obtain quantitative information from the experimental data, a valid comprehensive model that describes carrier generation, relaxation, recombination, and tunneling processes in the QD-QW system was developed and applied to fit the experimental data. A dynamic rate equation model based on the QD-QW energy level scheme shown in Figure 3.2(b) was build with the following assumptions: (1) under nonresonant optical excitation, where carriers are generated in the GaAs matrix with following relaxation to the QW states, WL states, and QDs states (directly or through the WL), the coherent effects quickly disappear during the relaxation processes and, therefore, play a minor role in the dynamics of tunneling and intersublevel relaxation within the QW and QDs; (2) excitonic tunneling is considered in the QD-QW system (which is a significant simplification as compared to the separate electron and hole tunneling and relaxation). The dynamic picture in the QD-QW system can be described by introducing the number of carriers n, that occupy each fth state of the QD (i = 0 for the QD ground state and / = 1,2,3 for the QD 1st, 2nd, and 3 rd excited states, respectively) and the number of carriers nw on the QW ground state. These numbers change in time due to the generation, transfer, and relaxation processes schematically illustrated in Figure 4.9. The rates of the QW and QD state population change can be written in the following equation set [79]:
dnw{t) dt
=rw(J

x nw(i)
Vexc) _R

nw{t)
_NR NR

^nw
Z_i ^
T

tr

N,

(4.2)

56

dnl it) dt

N,-n,(t) N,

+
j

W-nXOW)
N

(4.3)

,[tf,-n,(0],(0
J
N

n,(0 n,(f)
_NR

J*,J

where Gw (Iexc) and Gf (Iexc) are the carrier generation rates in the QW and f QD state, respectively; rw, xw and rt , r ; are radiative and nonradiative recombination times in

the QW state and the ith QD state, respectively; x\r is a carrier transfer time from the QW state to the z'th QD state; T is a time of intersublevel relaxation between the z'th and7th QD states; N, is the fh QD state degeneracy (A^; = 2/ for QDs of semispherical shape); [Nt n,(t)]/N, is the probability that the z'th QD state is empty which takes into account the state filling. Thus, if the carrier occupies one of the QD excited states, it has three choices of relaxation: radiative recombination, nonradiative recombination, nonradiative relaxation to the lower state.

QW

zw
5 o
X

o
CO

**<*
Transfer

re

Figure 4.9 Schematic illustration of the optical carrier generation, relaxation, and transfer processes in the QD-QW system between the QW th lowest state and the i QD state. 57

Equations (4.2) and (4.3) are general and include all possible processes and can be used to derive expressions that describe both time-dependent and steady-state cases of excitation. There are two excitation limits: high intensity and low intensity. In first case, almost all the states are filled, N, ~ n,(t), and the carrier decay is described by the effective recombination time fR defined as
1 -R 1
+ TR TNR

1 =*
T

TRTNR ' T?+Tm ^


}

and
1
-R
T

1
R ^
T

TRTNR

= + -
NR

=> f
~^
l

=
T

ww

(4 51
NR
y+-J)

tV

R ,

~w

w+ Tw

for the rth QD state and the QW state, respectively. In case of low excitation intensity, most of the states are empty, N, n,(t), and equations (4.2) and (4.3) can be reduced to

= ^(/)-=#-X^(/).

(4-6)

at where

ir(U=i^=^
is a carrier transfer rate term (coupling term) and

(4.8)

^=fe^]
I > Ttr)

(4-9)

is an effective transfer time. Assuming that the QDs are free of defects, the radiative decay time from each of the QD states is about the same [84], and, therefore, the effective 58

recombination time of z'th state rR is replaced with the general relaxation time xR. Equations (4.6) and (4.7) can be used to describe time-dependent behavior of the optical signal from the QD-QW system under the low excitation intensity condition. In case of continuous excitation (steady-state regime) - = 0 and dt

' dt

= 0, and equations (4.6) and (4.7) can be rewritten in the following form:

Gw(Iexc)-i-ZTr(IeJ

= 0,

(4.10)

In these equations, n% and ns' are the quasi-stationary carrier densities. Equations (4.10) and (4.11) are directly related to the cw PL measurements and allow the derivation of an expression that can be used to fit the experimental data. The integrated PL intensities for the QW and QDs can be respectively defined as

QW=a-fw
?w

aild

QD=a^lT>

(4-12)

where a is a factor essentially the same for the QW and QDs which depends on the particular experimental setup. Finally, using equations (4.10) and (4.11) together with the expressions for PL intensities (4.12), an expression for the ratio of the integrated PL intensities of the QW to QDs for the hybrid QD-QW system was derived as following (for derivation details see Appendix A):
hybr _

QW

1
+ R

1 1
r +T

=r*V'
T,R

Af

(4.13)

59

jref

where A

ref

=-^r

is the integrated PL intensity ratio for the reference QW and QDs, and

=. Expression (4.13) was used to fit experimentally measured PL intensities of

TR

the QD-QW samples at low excitation power as a function of the GaAs spacer thickness. If it is assumed that the carrier relaxation between the neighboring states within the dot is much faster than between non-neighboring states, one can denote the relaxation times between the neighboring states as r = rint and neglect the relaxation between nonneighboring states. In the low excitation intensity limit under continuous optical pumping, a solution of the equation (4.3) for the z'th QD state occupation numbers that define the PL signal is [79]

*> =miL '

ifo+ori, l
L

(r

k=m+1 D w

Nk
1-/?

r mt r
11

=R

mt

J '

s-T

( 4 14^) '

R r i n t

Here, L denotes the highest QD excited state, T^ =n^/r'tr.

Derivation details for

expression (4.14) can be found in Appendix A. In case of rapid relaxation between the neighboring states, rint fR, the population of the /' QD state under continuous

excitation is determined only by the optical generation, Gf, transfer from the QW, T^, and relaxation to the lower lying state, r mt :

<=r m t 2X D +7f).
k=i

(4.15)

60

4.5 Fitting experimental data and determination of tunneling time Now, having the model for carrier dynamics in the QD-QW system developed, it was be applied to fit the experimental data and deduce the carrier relaxation parameters. In order to do that, the PL results obtained from the reference QD sample were analyzed first, so that the relaxation parameters in the QDs not affected by the QW could be determined. The cw PL spectra of the reference QD sample taken at a series of excitation intensities were analyzed. The integrated PL intensities (TpLihxc), i = 0, 1,2, 3) of all QD states, E\ were obtained from the Gaussian fit of each PL peak (symbols in Figure 4.10(a)). Using the solution (4.14) and setting T^(Iexc) = 0 (no tunneling) and Gf (Iexc) = crj^ (where <r; is a scaling factor), the low power part of experimental data

was modeled, resulting in parameters r int , rf, and a. With these parameters, the whole power dependence of the PL from all QD states was fitted using equation (4.3) in steadystate regime. Assuming that all the carriers were captured in the QD via the highest (z-3) excited state and could relax only between the closest states, equation (4.3) was rewritten as a following set of equations [79]:
aI eXC

n3(N2-n2) N2rmt

n3 _Q f3R ' n2
T2
=Q

n3{N2-n2)
N2Tmt

n2(Nx-nx)
NrTmt

t \ n2{Nx-nx)
NSM

! \ n\N0-nQ)
N0Tmt

' nx ^ Q
T*

(4-16)

^V,t

r0R

0.

61

25 In 20 E = 15

. RefQD

5-

i 5
(A

(a)
1 . 1 . 1

c fl) c
_J

00 :

Excitation Intensity ///Q


- - RefQD - 0 - RefQW
JO'

"5 io

g107
c ~106 ! * * * * "
10"

J-' ^V"-*-*
O-"*
(b)
10 u

10"'

10"n

Excitation Intensity ///Q


Figure 4.10 (a) Integrated PL intensities of emission from all QD states of the reference QD sample (symbols) and their fit (solid lines) using equations (4.16) as a function of excitation intensity, (b) Integrated PL intensities of the reference QW (open circles) and QD (filled circles) samples versus excitation intensity. The data points are connected with lines for better eye guidance [79]. From the fit of the experimental data shown by solid lines in Figure 4.10(a) using equations (4.16) the following parameters were determined: the inter sublevel relaxation time rmt = 25 ps; recombination times from each of the QD states r^ = 790 ps, ff =911 ps, r2R = 863 ps, and * = 980 ps. The small intersublevel relaxation time was in a good agreement with the previously reported relaxation time of <40 ps determined by the same methods [83,85,86]. To confirm that the parameters rmt and T* were consistent, the modeling of the time dependent PL of the reference QD sample was performed using the rate equations (4.3) as in [84] and some parameters obtained from 62

the cw measurements by means of equations (4.16). The time-dependent PL measurements on another reference sample, single QW, gave an important parameter, the recombination time from the ground state of the QW, that was used in a process of deducing the time of a carrier transfer from the QW to QDs. The QW recombination time was determined to be F^ = 550 ps at low excitation power [79]. The next step was to use the expression (4.13) on the cw PL data of the QD-QW samples with different barriers to acquire quantitative information about the carrier tunneling processes. Figure 4.10(b) presents the integrated PL intensities of both QW and QD reference samples as a function of the excitation intensity. These intensities were required for the analysis since the expression (4.13) contained a term Aref = Ir^w j'TQD . Integrated PL intensity ratio values IQW jIQD of QD-QW samples measured at different excitation intensities are plotted as a function of the GaAs spacer thickness in Figure 4.11(a). Figure 4.11(b) shows the intensity ratios of QW to QD versus the AlAs effective barrier thickness measured in the QD-QW samples with constant total dsp = 5 nm. It is seen from Figure 4.11(a) that for a 20 nm thick GaAs spacer the intensity ratios became close to those of the reference QW and QD samples obtained from the data shown in Figure 4.10(b). When the spacer was thin, the Ahyhr sharply decreased for low excitation intensity indicating a strong tunnel coupling. For higher excitation intensities, this Ahybr drop was less pronounced due to the state filling that blocked the tunneling channel. At low excitation intensity (/eXc = 4.3xlO~4/o), when the condition N, n, was satisfied, the experimental intensity ratio was modeled using the equation (4.13).

63

2
V)

8 d
sp

10 12 14 16 18 20 (nm)
v

'

+-

10 r

f 10 iii.10 I L 10^
3

0A

V--"
cf = cL A + dA1A = 5 n m
sp GaAs AlAs

io-2c007V
'0.03 /

(b)
1
AlAs

Figure 4.11 (a) Integrated PL intensity ratios AhybT of QW to QDs recorded at various excitation intensities versus the GaAs spacer thickness. Inset: theoretical fit (solid line) of the lowest excitation data (symbols) using equation (4.13). (b) Integrated PL intensity ratios Ahybr of QW to QDs recorded at various excitation intensities versus the AlAs barrier thickness. AlAs layers were embedded in the GaAs spacer with total constant thickness of 5 nm. All dashed lines connecting data points were plotted for eye guidance [79].

The expression for the transfer time derived from (4.13) is


Ahybr(T,RT*+ArefTR)
T' R

(Aref

-Ahybr)

(4.17)

Then, using the Ahybr data form Figure 4.12(a) for every spacer thickness and excitation intensity, the QW to QD intensity ratio Aref of the reference samples, the reference QW

64

10"7
(0

10"

10"

10"

Excitation Intensity l/lQ

no"1

10"4

10"3

10"2

10"1

Excitation Intensity l/lQ


Figure 4.12 (a) Integrated PL intensity ratios AhybT of QW to QDs versus excitation intensity for QD-QW samples with various spacer thicknesses, (b) Experimental data (open circles) and theoretical fit using equation (4.13) (solid line) of the intensity ratio A y r versus excitation intensity for QD-QW sample with 5 nm thick GaAs spacer [79].

radiative recombination time rw = 550 ps, and assuming that nonradiative recombination is small and can be neglected (fR = TR), the transfer times were calculated from the cw PL data by formula: AhybrT*(l + Aref) ftr(CW) = - jref _ J hybr (4.18)

The transfer times calculated by equation (4.18) for the GaAs spacer thicknesses of 5, 8, 11, 15, and 20 nm are depicted in Figure 4.13. The carrier transfer time in the QD-QW 65

system exponentially decreased by 5 orders of magnitude from 20 to 2 nm spacer suggesting that this dependence can be described by the WKB approximation: Ttr=T0xexp(a-d ). (4.19)

This dependence allowed one to make a conclusion that the quantum tunneling of carriers was the dominant coupling mechanism between the QW and QDs in this system at low temperature. Moreover, subpicosecond tunneling times for spacers less than 2 nm

10" rQD- QW j
r

(0

E Mo 1

10 r vf c)

c 5

12
nm

16

20

dp (

Figure 4.13 The transfer (tunneling) times (filled circles) obtained from the cw PL measurements and fit (solid line) using WKB approximation (4.19). The symbol size reflects the experimental error.

suggested that it was primarily electrons that took part in tunneling process, since the holes were expected to tunnel with much longer times [56,87]. From the fit shown in Figure 4.13 by the solid line parameters of equation (4.19), r 0 = 0.3 ps and a = 0.5 nm-1, were determined. Since the PL signal resulted from the recombination of electron-hole pairs, it was suggested that the holes were captured into the QD directly from the WL or GaAs barrier and rapidly relaxed to the ground state [88] to compliment the electrons that tunneled from the QW. 66

Thus, the rate equation model satisfactorily described the tunneling dynamics of the carriers in the QD-QW system at low excitation regime which was seen from the fitting curves in the inset of Figure 4.11(a) and in Figure 4.12(b). However, some deviations of the model from the experimental data were observed for the high excitation powers. These deviations were attributed to the electron-electron scattering and Augerassisted transfer processes [16]. The values of the tunneling times discussed so far have been deduced from the cw PL experiments. Time-resolved PL data presented another, additional way to extract the tunneling times from the direct measurements of the PL decay dynamics. As was already mentioned above, the PL transients were very sensitive to the state filling effects that influence the early stages of the carrier dynamics after the laser pulse at high excitation power (see Figure 4.8(a)). Therefore, the low excitation intensity (Ip= 1.3 xlO4 W/cm2) was chosen for the measurements of the QW PL transients of the QD-QW samples with various GaAs spacer thicknesses as shown in Figure 4.8(b). The PL decay transients were fitted by the monoexponential function that allowed extraction of the QW PL decay times rD (dsp) for different spacer thicknesses:
f

WW(0)exp
V

t^
T

(4.20)

Dj

The values of fD (open circles in Figure 4.14) approach the reference QW radiative recombination time f^= 550 ps (shown as a dashed line in Figure 4.14) when the GaAs spacer thickness approached 20 nm. The effective carrier transfer (tunneling) time can be determined as

67

104

r\r\

n>\i\r.

m3
w
Q. 102 O

^x
" j~i"*T

)~

E 101

U^

r-;

o - rD ;
A -rlr(TR) -Tlr(0W)

j y
Vy

c)

12

; I !
16

20

Figure 4.14 Carrier transfer (tunneling) times versus the GaAs spacer thickness obtained from cw (filled circles), ftr (CW), and time-resolved (filled triangles), flr (TR), PL measurements. fD (open circles) is the PL decay time measured in the QW which for 20 nm thick spacer approaches the value T (dashed line) measured in the reference QW [79].

1
*Z>(0
T

1
W

+-

Ttr( sP)

T-TD(d

(4.21)

The carrier transfer times obtained from the TRPL measurements are displayed in Figure 4.14 (filled triangles) together with the data obtained from the cw experiments (filled circles). Although the carrier transfer times from the QW to QDs measured by the two methods were expected to be the same, the difference between the data sets was obvious. The TRPL data also followed the exponential dependence on the spacer thickness, confirming the tunneling mechanism of coupling. However, the slope of the fit deviated from that of the cw data, especially in the thin spacer region. Interestingly, the tunneling times deduced from the TRPL measurements coincided well with the results obtained before for the QD-QD and QD-QW coupled systems using similar methods [47,60]. 68

From deeper analysis of the experimental details of the two methods, one can find logical reasons for the discrepancy between the spacer thickness dependences of the tunneling time values extracted from the cw and time-resolved PL measurements. The PL decay transients reflected all stages of carrier relaxation down to the QW ground state, which was tunnel coupled to the QDs, starting from the moment of the laser excitation of the carriers deep into the conduction band. This relaxation was usually very fast for the QW (< 1 ps), but it still added to the QW PL decay time fD. On the other hand, the cw PL intensity included recombination only from the lowest QW state. Also, the PL transients were very sensitive to the state filling. Although the TRPL measurements were done at the lowest possible excitation intensity (Ip= 1.3 x 10 4 W/cm2), it was still much higher than the excitation intensity used in cw PL measurements (Iexc = 4.3x10_1 W/cm2). Therefore, the state filling effects in the QW could significantly influence and increase the PL decay times obtained from the TRPL experiments. On the contrary, the rate equation model that was used to describe the cw PL results included the state filling effects in both QW and QDs. Therefore, the effective transfer times extracted from the cw PL data were believed to be more accurate and reliable as compared to those deduced from the TRPL data. In addition, for thick GaAs spacers, the difference (T -fD) that

was in the denominator of the expression (4.21) for ftr(TR) became vanishingly small, hence, increasing the error of the transfer time determination (shown as error bars in Figure 4.14). For thin GaAs spacers, where the QW PL decay time was approximately equal to the transfer (tunneling) time and was expected to be ~1 ps, the lowest PL decay time measurement was limited by 15 ps, the time resolution of the streak camera.

69

To conclude this part, the tunneling times obtained from the cw PL measurements turned out to be more accurate than the values obtained from the time-resolved PL experiments at nonresonant excitation due to the limitations of the TRPL experimental arrangement. To extract the correct tunneling times directly from the time-dependent measurements for this QD-QW system, the optical pump-probe technique should be used at low intensity and resonant excitation of the QW ground state.

4.6 Effect of QW composition variation on tunneling processes In the previous sections, the strength of coupling between the QW and QDs was varied by the composition and geometrical dimensions (thickness and height) of the separation barrier. Another way to probe the effectiveness of tunnel coupling is to tune the energy band alignment of the QD-QW structure. The energy levels can be shifted by an applied electric field along the growth direction of the structure. However, in this study, variation of the QW layer composition was used to tune the QW energy between the lowest QD state and the WL energies in the set of QD-QW samples. Figure 4.15 shows low temperature (Y=10K) PL spectra measured under increasing excitation intensity, 7exc, for the QD-QW sample with In content* = 0.15 in the QW and GaAs spacer thickness of dsp = 4 nm. At the lowest excitation intensity ^exc= 1-5x10"6io (7o = 1000 W/cm2), the QD exciton ground state was found at EQD ~ 1.10 eV (AQD = 1127 nm) and is fitted well by a single Gaussian with a full width at half-maximum (FWHM) of r o = 32 meV, which reflected the highly homogeneous size distribution of the QD ensemble. With increasing 7exc, additional PL bands emerged

70

900

950

1000 1050 1100 1150 1200

Wavelength (nm)
Figure 4.15 PL spectra normalized to the QD ground state emission of the QD-QW structure with 4 nm GaAs spacer and x = 0.15 in the InxGai.xAs QW measured at various excitation intensities. The spectra are vertically shifted for clarity. I0 = 1000 W/cm2, T= 10 K. at A'QD (i= 1, 2, 3...) that corresponded to the QD excited state emission resulting from the state filling at higher excitation intensities [83,84]. The maxima of these additional bands were separated by an energy of -66 meV and the corresponding PL bands were also narrow, r, = 34 3 meV. In addition to the observed QD excited states, a strong PL band reflecting the QW exciton emission developed at Exw = 1.347 eV (/lxw = 920 nm). An important correlation between the QW and QD emissions was seen in Figure 4.15. The QW excitonic emission appeared only if all QD states at energies below EQW were filled at high 7exc. At low excitation power the QW PL signal was detected. This implied that the tunneling induced dot-well coupling in the samples with thin spacer (JjP = 4nm) was sufficiently strong to fully suppress the QW PL [79]. Due to this 71

coupling the carriers effectively tunneled from the QW states into the energetically lower-lying QD states quenching the QW emission at low excitation power. Raising the excitation intensity increasingly populated the lower-lying QD states with carriers. This state filling reduced the carrier transfer probability thus populating the QW long enough to observe its exciton state emission in PL spectrum. This created a picture of the QDQW hybrid system where the total carrier (exciton) capacity of the QDs was determined by the total number of the QD states located below the QW ground state, but the carrier flow rate from the QW to the QDs was determined by the QD-QW coupling strength. In order to test this simple picture of the QD-QW system, the structure was examined as the energy of the QW ground state increased approaching the energy of the InAs WL. For this purpose, QD-QW samples with QW composition of x = 0.07, 0.10, 0.15, and 0.18 and the spacer thickness of dsp = 4 nm were chosen. Using the PL spectra shown in Figure 4.15 and previous results [79], the energy band structure and carrier dynamics of this set of QD-QW samples was presented as depicted in Figure 4.16. The excitonic transitions of the QW lowest state are shown for the four QD-QW samples with different values of x relative to the energies of the QD excited states. Carrier transfer from the QW states to the QD excited states is shown with thick horizontal arrows whereas the radiative transitions are identified by wavy downward arrows. It was seen from Figure 4.16 that a decrease in x increased the QW exciton energy EQW . In this case the number of QD states with energy less than EQW increased. Thus, the excitation intensity needed to saturate these QD states and observe a noticeable QW PL signal also had to be increased. Figure 4.17 illustrates the PL spectra of all four QD-QW samples along with the reference QD sample measured at the same excitation intensity Iexc = 10/o72

QW
lnxGai.xAs
Sample Sample Sample Sample 4 3 2 1

QDs
d, spacer In As

Figure 4.16 Energy band alignment of confined states in hybrid QD-QW structures with various QW compositions: x = 0.18 (sample 1), x = 0.15 (sample 2), x = 0.10 (sample 3), and x = 0.07 (sample 4). Carrier dynamics is shown by different types of arrows. The normalized PL spectra of corresponding reference QWs measured at low excitation intensity are shown in the inset of Figure 4.17. For x = 0.07 and 0.10 the emission energy EQW of the QW was higher than the emission energy EA0D of the QD 4th excited state, while for x = 0.15 and 0.18 the QW emission was at energy below that of the QD 4th excited state. At the same time, the intensity of the QW emission forx = 0.7 and 0.10 was more than one order of magnitude lower compared with the intensity for x = 0.15 and 0.18 measured at the same excitation conditions. The PL spectra of all hybrid QD-QW samples at energies below EQW were very similar to that of the reference QD sample in the respective spectral range. It was also found that the amplitude of the GaAs free exciton PL band ( XQOAS = 824 nm) remained

73

106

850

900

950 1000 1050 1100

Wavelength (nm)
Figure 4.17 PL spectra of the QD-QW samples with the same spacer thickness (dsp = 4 nm) and various In content in the InxGai_xAs QW and of the QD reference sample measured at the same high excitation intensity (IeXc= 10io). Inset depicts normalized PL spectra of the reference QW samples with corresponding In compositions detected at low excitation intensity (Iexc = 1.5xl0"6/0). I0 = 1000 W/cm2, T= 10 K. constant for all samples under study. This observation indicated that the incorporation of the QW into the QD samples did not introduce new trapping centers in the hybrid QDQW structures in comparison with the reference QD sample. Also, the QWs did not introduce extra losses for the carrier transfer from the GaAs barrier to the QD states under nonresonant excitation (Aexc = 532 nm). The QW essentially acted to collect the carriers that relaxed from the GaAs barrier and redistributed them uniformly among the QDs. This redistribution was clearly seen from Figure 4.17, which demonstrated the cutoff of the QD excited state emission at high excitation power with subsequent switching

74

to the QW emission. This switching had a threshold that depended on the relative energy position of the QW exciton within the energy range between the WL and the QD ground state and provided a mechanism to control the QW emission intensity by tuning the excitation intensity in hybrid QD-QW nanostructure. The QW emission in a hybrid QD-QW structure can be described by the following simple model. It is assumed that under nonresonant cw excitation in the GaAs barrier, the hot carriers relax into the QW states and, after rapid thermalization, generate NQW(Iexc) carriers per unit time in the QW ground state. The further relaxation of these carriers should fulfill the relation NQW(LJ = NRQW+NNQRW+NTQW, (4.22)

where NQW is the number of carriers that recombine radiatively per unit time in the QW and contribute to the QW PL signal IPQW ; NQ% is the number of carriers that recombine non-radiatively per unit time in the QW; NTQW is the number of carriers that tunnel from the QW states to the lower-lying QD states per unit time. Then the QW steady-state population can be expressed as nSQW(Iexc) = NQW{Iexc){rR + YNR + YT)~1 , (4-23)

where yR , yNR, and yT are the radiative recombination, non-radiative recombination, and carrier transfer rates, respectively. In this model, the back-transfer of carriers from the QD states to the QW state is neglected, since the relaxation to the lower QD states is much faster process than any of the rates defined in the equation (4.23). The QW PL intensity can be defined as
IQW=P-KW,

(4-24)
75

where /? is a factor that depends on the details of an experimental setup. Now, using equation (4.22), the QW PL expression can be rewritten as

The NQW term in equation (4.25), which describes the carrier flux from the QW to the QD states, can be represented in a form that takes into account the QD state filling:

KwV-c) = 4w(iejt{N,~n,(Iexc))>
1=0
T

( 4 - 26 >

T,i

where nXIexc), Nt, and TTI are the z'th QD state population, the z'th QD state degeneracy, and the tunneling time from the QW to the r'th QD state; Q denotes the highest index i for which the QD excited state energy E'QD lies below the QW exciton energy EQW . If the tunneling time rT, weakly depends on the index /, and if it is assumed that degeneracy is Nl - 2(i +1), then the number of the QD states (capacity of the QD system) that are filled by carrier tunneling from the QW can be given by n iV, = ( Q + l)(Q + 2).
(=0

(4.27)

Using equations (4.25) and (4.27) one can derive an expression that can be used to determine the threshold excitation intensities over which the QW PL signal for various positions of EQW with respect to the QD excited states is non-zero: n N, NQAhJ-KwileJll-^-Kw
i=0 TT
j

*QW\*exc)

= a ( / ) k / c ) - ( Q + l)(n + 2)], (4.28)

where

76

* ( / ) = /?
and

OW\*exc)

Mimy

rT(NQAIeJ-N)
n

QW\*exc)

In this model, a finite QW emission occurs if the term [A(Iexc)-(Q. + l)(Q + 2)] in equation (4.28) is positive. This introduces a threshold intensity for the QW PL which is defined largely by Q , i.e. the total number of QD states that are below the QW energy. Equation (4.28) contains two parameters at a fixed Iexc and can be applied to fit the amplitude of the QW PL intensity. Figure 4.18 depicts the dependence of
IPQW

on Q for

the case of 4 nm spacer between the QW and QDs and an excitation intensity Iexc = 107o, when the QD excited states were substantially filled. The good agreement of this fit with the experimental data points supports the reliability of the model that describes the carrier tunneling from the QW to the QDs including the influence of the QD state filling.

10

.a

a > 1

_
I

calculation - 1 : x = 0.18 - 2 : x = 0.15 * - 3 : x = 0.10 A - 4: x = 0.07


I . I .

a.

"l

, .4
3 versus the Q index calculated

QD excited state index


IPQW

Figure 4.18 The QW PL intensity

using equation (4.28) (shown by solid line). Experimental values for each sample are shown by the symbols. 77

4.7 Strain-induced QDs in dot-in-well structure Buried QDs create nonuniform elastic strain which affects the properties of the surrounding matrix [89,90]. In Section 2.4, it was discussed that QDs can create potential profile in a proximal QW. The carrier confinement arises from the modulation of a deformation potential created above and/or below the QDs in the QW. As a result, selfassembled QDs can act as stressors and facilitate the formation of the strain-induced QDs (SIQDs) in the QW [67]. Such systems have been studied in the structures with buried InAs QDs near the GaAs/AlxGai_xAs QW [68]. On the contrary, no evidence of the SIQDs formation in the QW was found in the QD-QW samples (even with very thin -1.5-2 nm GaAs spacer) studied in this dissertation work (see for example Figure 3.4). However, the situation became different if the InAs QDs were directly overgrown with the InGaAs QW as shown in Figure 4.19. The TEM image was recorded off the [110] zone axis under 2-beam conditions with (002) reflecting planes. The InAs QD was clearly visible as a dome or platelet and clearly lay on top of a wetting layer. The dark contrast underneath and above the QD was the result of the tensile strain produced by the lattice mismatch between the dot and its surrounding materials. It was noteworthy that the resulting strain field in the well region extended along the [001] growth direction

Figure 4.17 (a) Cross-sectional bright-field TEM image of a QD buried in the QW. (b) Schematic structure that mirrors the TEM image with denoted regions. 78

into the GaAs cap layer. This observation confirmed that the QD formation was directly responsible for strain effects that go beyond the QW. Figure 4.20 illustrates the low temperature PL spectra measured for a range of excitation intensities from 4x 10"7/o to 4/0 (/o = 1000 W/cm2) from the sample shown in Figure 4.19. At low excitation intensity a single PL band was observed with a maximum at EQD = 1.090 eV and full width at half-maximum of r~30meV. This band corresponds to the exciton ground state transition in the QDs. Comparing the EQD value for the InAs QDs capped with a strain-reducing Ino.15Gao.85As layer with that of the reference InAs QD sample, the PL peak of reference QDs was found to be higher in

InAs QD

S\QD\ InGaAs QW

1.1

1.2

1.3

Photon Energy (eV)


Figure 4.18 PL spectra of InAs QDs capped by 14 run thick Ino.15Gao.85As QW measured at various excitation intensities. The spectra are normalized with respect to the QD ground state peak and shifted vertically. Each PL peak is identified and denoted by vertical dashed (QD), dash-dotted (SIQD), and short-dashed (QW) line. I0 = 1000 W/cm2, T= 10 K. 79

energy by -50 meV. This was due to the lowering of the potential within the QDs from the strain redistribution and lower potential barrier of the InGaAs layer. An increase in Texc resulted in additional PL bands that subsequently emerged at higher energies E'QD (i = 1,2) and were related to the dipole-allowed interband transitions between QD excited states. The PL bands at E'0D were separated from each other by ~66 meV and had linewidths of 30 to 40 meV. Gradual appearance of these PL peaks with increasing excitation intensity was attributed to filling of the lower energy states in the QDs with photogenerated carriers. At high excitation power, emission from a ground state and up to two excited states of InAs QDs overgrown with the InGaAs QW were observed in the PL spectrum (see Figure 4.20). Similar structure of the QD excited states emission was observed in the PL spectrum of the reference QD sample, thus facilitating the QD state identification in the strain-reduced samples. In addition to the PL bands related to the QD states, the emission from the Ino.15Gao.85As QW developed at high excitation intensities. The absence of the QW band in the PL spectrum (Figure 4.20) under low excitation power was an indication of coupling between the QW and QD subsystems. Due to this coupling photo-excited carriers relax from the GaAs barrier to the lowest energy states available in the system. Those were the states of the buried InAs QDs. After filling the lowest (QD) states the, higher lying (QW) states gained a finite population at high excitation conditions. As a result the QW emission became dominant in the PL spectra of strain-reduced samples due to its large density of states. In addition to the identified QD and QW states, PL bands were found at 1.247 eV and 1.297 eV, which did not follow the excited state pattern of the QDs and were located at energies too low to be from the QW or the InAs WL. It was 80

believed that these bands resulted from the recombination of carriers confined within the region defined by the strain created by the QD inside the QW, so-called strain-induced QD. For better understanding of the SIQD origins in this system, the effects of strain resulting from the InAs QDs were qualitatively described. According to Grundmann et al. [91], there existed a strong hydrostatic compressive strain which extended a very short distance into the QW as a results of the relaxed InAs QD pushing against the QW material. This compressive strain resulted in an increase in the total bandgap energy therefore creating a narrow barrier between the SIQD and the InAs QD [91]. At the same time, there existed a biaxial tensile strain which extended and decayed more slowly through most of the thickness of the QW. This tensile strain contributed to a slight decrease of the bandgap energy in that region, primarily due to a shift in the valence bandedge [92]. This redshift created a weak but non-zero confinement in the lateral direction of the QW confining carriers in a QW region just above the InAs QDs and forming the SIQDs. The resulting conduction band diagram is drawn schematically in Figure 4.21. The SIQDs energy levels were located below the lowest QW state and above the highest InAs QD state. In order to better understand the coupling in this system, the PLE spectra of the reference and strain-reduced QD samples were measured at low temperature. Figure 4.22 shows these PLE spectra detected at l det = JiQD which corresponded to the PL maximum of the InAs QDs at EQD in both samples. The QW PL spectra measured at two high excitation intensities (io and 47o) in the strain-reduced sample are shown too. Both PLE spectra revealed strong absorption resonances at XxGaAs =818 nm corresponding to the 81

QW n =1 n =0 n = 1 ^SIQDs n =0

n =2 n =1 n =0 InAs QDs

Figure 4.21 Conduction band energy diagram of the structure with the InAs QDs capped by the Ino.15Gao.g5As QW.

w 105
.Q
i_

> ^
^

0 QDs capped with QW -D-RefQDs |nGaAsQW n=1n=0


1 1

_
(A
*>

.5. "<5

^
"*"~
i

' fry*\
' / ft

"E

#10*

^ ^ "1 i^ L J?I J^tesss> 0


if
j

W .a _
\4/

o
(0
* J

iV

ft

%SS& )

If ]

i,

75 c o
</> _i Q_

3*

0 X

*io 3 :

_ ;

_ ;0

\,

/
T&, /

/' ' /
/
/

\ A
if0

^det ~ ^ Q D

800

850

900

950

Wavelength (nm)
Figure 4.22 PLE spectra of InAs QDs capped by the Ino.15Gao.g5As QW (open circles) and reference InAs QDs (open squares, left ordinate) both detected at a wavelength of the QD ground state emission. Q W PL spectra from the sample with QDs capped with QW measured at high excitation intensities (To and 4Io). The ground state and first excited state of the QW are denoted by the vertical dashed lines. IQ = 1000 W/cm2, T= 10 K.

82

GaAs exciton. The peaks at 830 and 854 nm in the reference QD spectrum were identified as light and heavy hole excitonic states of the WL, respectively. These states most likely were unbound in the WL of the QW capped QD sample due to the lower band offset between InAs and InGaAs as compared with GaAs. As a result, all of the remaining features in the PLE spectrum of the QD-QW system were confined states of the QW. The strong absorption resonances at
XXQW =

928 nm and

XXQW =

898 nm were

attributed to the exciton ground state (n = 0) and 1st excited state (n = 1) in the QW, respectively. This was confirmed by direct comparison with the QW PL peaks also shown in Figure 4.22 at AQW = 934 nm and /lg^=913nm, which exhibited strong Stokes shifts of ~7 meV and ~18 meV, respectively. Thus, a spectroscopic comparison of the PL and PLE spectra from InAs QDs capped with Ino.15Gao.85As or capped with GaAs revealed all expected changes caused by capping in both the InAs QDs and InxGai.xAs QW through the related changes of the characteristic spectral features of the QDs and QW. On the contrary, the two PL peaks that developed at energies ESIQD = 1.247 eV and E\IQD = 1.297 eV under elevated excitation intensities could not be attributed to the QD or QW emission. It could not be related to transitions between the WL states either, because the WL states, if they were bound, should be observed energetically above the QW states (below in wavelength) as can be seen from the PLE spectrum (Figure 4.22). These PL bands lay below the region of the QW transitions by -80 meV and -30 meV, respectively. Increasing the excitation intensity strengthened the PL bands and shifted their maxima towards lower energies, as can be seen from Figure 4.20. Similar behavior was observed in the PL spectra of SIQDs created in a nearby GaAs/AlxGai.^As QW by the strain field coming from the buried self83

assembled InAs stressor QDs [68]. Due to the lateral confinement of excitons in the QW caused by the strain, the formation of bound, zero-dimensional, states was possible at energies lower than the QW ground state energy. It has been shown that the PL spectrum of excitons laterally confined in the QW by strain is red-shifted as compared to the PL of unstrained QW [93,94]. Thus, based on the above discussion and evidence of strained region in the QW on top of the QD (Figure 4.19(a)) the additional features in the PL spectrum shown in Figure 4.20 was attributed to the SIQDs emission. The PL spectra of the sample with InAs QDs capped with the InGaAs Q W versus temperature measured at high excitation intensity (7exc= 1000 W/cm2) are depicted in Figure 4.23(a). The horizontal axis is the photon energy normalized to the QD ground state energy which allowed tracing the temperature shifts of all PL peaks with respect to the QD ground state PL peak. Figure 4.23(b) shows the positions of the QD 1st excited state, QW ground state, and SIQD state with respect to the QD ground state energy as a function of temperature. It was seen that with increasing temperature relative position of the QD 1st excited states did not shift more than 1 meV, whereas the position of both the QW state and the SIQD state exhibited a red-shift by 7 meV over 200 K. The temperature shift of the SIQD PL peak followed the behavior of the QW PL peak (with 80 meV offset). The SIQD PL signal quenched at about 250 K. If the QW was depleted of carriers at high temperature due to the thermal escape, the SIQDs, being a part of the QW, also lacked carriers at high temperatures. In addition, the SIQD PL band was found to broaden significantly at higher temperatures. This broadening was assigned to the thermal populations of the SIQD excited states. From analysis of the SIQD energy spectrum performed for the GaAs/AlxGai.xAs QWs [68] it was found that the excited states in the

84

SIQDs were separated from each other by -1-9 meV. Such a small separation can explained the gradual filling with carriers of the higher energy states as temperature increased, thus resulting in the SIQD PL band broadening.

InAs QD

InGaAs QW

Figure 4.23 (a) PL spectra of InAs QDs capped by the Ino.15Gao.85As QW measured at various temperatures under high excitation intensity (/exc = 1000 W/cm2). The horizontal axis is normalized to the QD ground state emission energy. All spectra are normalized with respect to the intensity of the QD ground state PL peak, (b) Relative positions of the QD 1st excited state (red triangles, left ordinate) and QW ground state (open green circles, right ordinate) with respect to the QD ground state energy as a function of temperature. Additionally, the difference EI0D - EQD (filled blue circles, right ordinate) shifted up by 80 meV in comparison with the corresponding QW difference is shown.

85

4.8 Summary The coupling in the QD-QW structure was studied using cw and time-resolved PL methods under various excitation conditions. The strength of coupling and the excitation transfer between the QW and QDs was controlled by the barrier width (from 2 to 20 nm) and height through the material composition (GaAs, AlGaAs or AlAs) of the effective barrier. Using the rate equation model that took into account state filling effects, a systematic analysis of the experimental QW and QD PL intensities in the QD-QW samples as a function of barrier thickness was performed. The results of this analysis suggested that the main transfer mechanism in the studied QD-QW system was a quantum-mechanical tunneling of carriers from the QW ground state to the QD 3rd excited state. It was observed that at low excitation intensities in samples with a thin barrier (less than 8 nm) only a QD emission peak was present in the PL spectrum, which suggested that most of the carriers effectively tunneled from the QW to QDs. At high excitation intensities, when the QD states were completely populated, the tunneling from the QW to QDs was effectively blocked due to the Pauli exclusion principle, which allowed the QW PL signal to be detected. The tunneling times extracted from the cw PL data exhibited exponential dependence on the barrier thickness. For a 2 nm barrier the carrier tunneling times were in the order of 1 ps, whereas for 20 nm barrier it was more than 10 ns. On the contrary, the tunneling times extracted from the time-resolved PL measurements under non-resonant excitation turned out to be longer than those obtained from cw data. This was explained by the relaxation processes of the non-resonantly

86

photogenerated carriers and the state filling effects at relatively high excitation conditions. Another way to control the coupling regime between the QW and QDs - tuning the alignment of the energy levels - was explored in a set of the QD-QW samples with different QW compositions. The number of the QD states that lay below the energy of the QW exciton determined the QD exciton capacity and, hence, the excitation intensity needed to fill all these states in order to block the tunneling channel and detect the QW PL signal. When the InAs QDs were directly overgrown with the InGaAs QW, the strain field surrounding the QDs caused lateral confinement in the QW in the vicinity of the QDs, which created so-called strain-induced QDs. The PL emission at energy lower than the QW ground state from the excitons localized in the SIQDs was observed at high excitation intensities when all the lower states of the InAs QDs were filled and the QW was able to supply carriers to the SIQDs. No evidence of such SIQDs were found in any of the QD-QW samples where the InAs QDs were grown on top of the InGaAs QW with a GaAs spacer between them.

87

V.

SIGNATURES OF COHERENT TUNNELING IN QD-QW NANOSTRUCTURES

The functionality of many electronic and optoelectronic devices is based on the transfer (or injection) of charged carriers (electrons and/or holes) from one part of the device into the other. In most cases this carrier injection is believed to be an incoherent process. On the other hand, it has been demonstrated that with semiconductor nanostructures it is possible to utilize controlled coherent interactions which may be of critical importance for further development of the quantum computing [95]. Strong coherent coupling with hybridization of the confined states was achieved by electrically tuning the energy levels of the two individual QDs into the resonance [14,15]. In this work, the coherent phenomena in the hybrid system consisting of nanostructures with different dimensionalities, OD (QD) and 2D (QW), were investigated. The carriers were resonantly injected from the QW ground state through the tunnel barrier into one of the QD excited states, followed by relaxation to the QD ground state and recombination. The strength of the coupling was varied by changing the thickness and/or height of the separation barrier between the QW and QDs. The main tool used to probe the tunnel coupling in this hybrid system was photoluminescence excitation spectroscopy which allowed resonant excitation of the Q W exciton state and detection of the PL response from the QD ground state. This PLE signal as a function of the barrier thickness revealed information about a kind of the coherent tunneling regime in the hybrid QD-QW nanostructure [96].

88

5.1 Anomalous dependence of QD PLE signal on barrier thickness under resonant QW excitation The evolution of the PL spectra with increasing excitation intensity shown in Figure 5.1 for (a) reference InAs QD sample and for QD-QW samples with (b) dsp = 2 nm and (c) dsp = 11 nm were discussed in Section 4.1. Here, the PL spectra are shown to make a comparison with the PLE spectra (shown by blue open circles in Figure 5.1) measured at the energy of the QD ground state PL emission peak for each sample. Again, it is important to emphasize that the QW PL emission (at -1.35 eV) of the QD-QW samples was observed in the same energy region as the 3 rd QD excited state (at -1.34 eV) in the spectrum of the reference QD sample (Figure 5.1(a)). In the PLE spectrum of the reference QD sample, a clear absorption band that corresponded to the 3rd QD excited state was observed in the spectral range between 1.3 eV and 1.4 eV (Figure 5.1(a), open circles). This band, with maximum at -1.35 eV and FWHM of 35 meV, appeared to be Stokes shifted as compared to the PL peak of the 3 rd QD excited state [96]. The PLE spectra of the QD-QW samples with 2 nm (Figure 5.1(b), open circles) and 11 nm (Figure 5.1(c), open circles) GaAs spacers were quite different. For the sample with an 11 nm spacer, two main resonances appeared in the spectrum (at 1.3 eV and 1.42 eV) which corresponded to the absorption of the two lowest states in the QW. For the sample with a 2 nm spacer, the lowest QW state absorption peak became more asymmetric with pronounced tails on both sides. This asymmetry was believed to be a manifestation of the Fano resonance [97] due to the interference of a narrow QW peak with a broad peak from the 3 rd QD excited state [98,99]. The absorption peak of the QW excited state became broad and split into several

89

Figure 5.1 Normalized PL spectra (solid lines) measured at various excitation intensities and PLE spectra (open circles) detected at energy of the QD ground state emission of (a) reference InAs QDs and QD-QW samples with (b) 2 nm and (c) 11 nm GaAs spacer. The PL spectra are shifted vertically for clarity. I0 = 1000 W/cm2, T= 10 K. [96]

90

resonances which was believed to be an indication of the QD and QW state mixing and hybridization due to coupling [96]. The shape analysis of the PLE spectra from all QDQW samples as a function of the GaAs spacer thickness is described in detail in Section 5.4. The proposed exciton energy structure of the QD-QW hybrid system is shown in Figure 5.2 [96]. This scheme is in a good agreement with the experimental PL and PLE results and is also supported by the calculations that take into account geometrical size and shape of the QD and QW obtained from the TEM images. It shown that the QW exciton ground state |l) happens to be in a close resonance with the 3 rd QD excited state 12). When a carrier arrives to the 12) state, it rapidly relaxes to the QD ground state 13),

QW
InGaAs WLl

QDs
In As

ID

n
<T23

|2>

'Tio

|3>
^30

|0>
Figure 5.2 Simplified exciton energy structure of the QD-QW system with arrows showing the carrier dynamics under resonant excitation conditions [96].

91

where a radiative recombination takes place. Such alignment of the energy levels makes this system particularly interesting and suitable to study coherent tunneling processes in a hybrid 0D-2D nanostructures. First, the ratio of the integrated PL intensities of the QW to the QD measured under nonresonant excitation conditions was studied as a functionof the GaAs spacer thickness (or transmission through the barrier) as shown in Figure 5.3(a) for two excitation intensities 50 W/cm2 (open circles) and 0.15 W/cm2 (filled circles). The carrier transmission T(E) through the potential barrier was calculated taking into account the effective barrier profile of each QD-QW sample T(E) oc expj - yV(x) - Edx (5.1)

where V(x) is the barrier height and E is the initial energy of a carrier (QW lowest state) before tunneling. The intensity ratio droped for thinner spacers (higher transmission) which meant that the relative QD PL intensity increased (more effective transfer of carriers from the QW to QDs). At higher excitation intensities, the tunneling became less effective due to the QD state filling and the intensity ratio became higher for the same spacer thicknesses (open circles compared to filled circles in Figure 5.3(a)) [96]. The resonant excitation conditions gave more direct insight into the carrier tunneling processes between the QW and QDs. In the PL excitation experiment, the laser was tuned in resonance with the energy of the QW exciton absorption, while the PL signal was detected at the QD ground state emission energy (around 1.145 eV). The excitation intensity was chosen to be low (7eXc= 1.5><10~67o), so that the QD state filling did not affect the carrier tunneling. The QD PLE intensity recorded for each QD-QW

92

dp<nm)

11

Transmission
Figure 5.3 (a) PLE intensity (open triangles) detected at the QD ground state emission wavelength of the QD-QW samples under resonant QW excitation as a function of barrier transmission calculated for various GaAs spacer thicknesses (shown on top horizontal axis). Ratios of the QW to QD integrated PL intensities measured for two excitation intensities under nonresonant excitation at Xexc = 532 nm. (b) The same dependencies as in (a) measured in the QD-QW samples with various thicknesses of AlAs effective barriers that were embedded in GaAs spacer with total constant thickness of 8 nm. [96] sample is plotted versus the GaAs spacer thickness (or transmission through the barrier) in Figure 5.3(a) with open triangles. With the largest (20 nm) GaAs spacer the coupling between the QW and QDs was practically absent and the observed QD PLE signal was mainly due to the recombination of carriers directly excited in the QDs. As the spacer thickness decreased, the QD PLE signal rapidly increased because of stronger coupling 93

(more carriers tunneled to the QDs and contributed to the emission). The QD PLE intensity reached saturation at the 8 nm spacer thickness and even starts to drop for spacers thinner than 3 nm (see Figure 5.3(a)). This is a surprising result that cannot be explained based on incoherent carrier tunneling. Figure 5.3(b) demonstrates similar data on an integrated PL ratio (filled circles) and PLE signal (open triangles) for the set of QD-QW samples with a fixed 8 nm total thickness spacer made of AlAs layers of various thicknesses (1-4 nm) symmetrically embedded in the GaAs material. It was observed that the QD PLE intensity decreased by two orders of magnitude (suppressed tunneling) as the AlAs barrier became thicker [96].

5.2 Optical Bloch equation model In order to better understand the ambiguous PLE results, an attempt was made to model the QD-QW system using the optical Bloch equations. The simplified four-level energy scheme shown in Figure 5.2 was considered in the modeling. This is the exciton energy structure with the crystal ground state 10), the lowest QW state 11), the QD excited state 12) which is in resonance with the QW state 11), and the QD exciton ground state 13). The energies of these states are Pica,, where i = 0, 1, 2, 3 corresponding to the state numbers. It was assumed that the exciton in the QW was localized at the shortest distance from the closest adjacent QD [96]. This localization could be facilitated by the strain field from the QDs in the samples with very thin (< 3 nm) GaAs spacers. Optical polarizations p0l with i = 1, 2 ( p are the elements of the density matrix)

were created in the QW (state |l)) and QD (state |2)), respectively, by the light 94

excitation with the amplitude of an electrical component E0(a>L) and frequency coL. The Bloch equations of motion that describe polarizations of states 11) and 12) are [96]

-^21 = -iSxpm -iClx(pu at dp,02 dt

-p00)+ UpQ2 -rlPoi,

(5-2)

= -idiPoi ~ i&2 (Pn - Poo) + iJPoi - /2P02>

(5.3)

where St = coL a>l + & > 0 give the detuning from the resonance frequency, Q ; are the Rabi frequencies given by Q, = d,E0 (coL) (here dt are the transition dipole moments), the dephasing rates are denoted as y[, and J is the matrix element that describes the tunnel coupling between the states 11) and 12). The resonant tunnel coupling of the two states may result in Rabi oscillations. This coherent polarization between the states |l) and |2) is described by the following equation of motion [96]: dPn = ~iS3Pn - iJ(p22 ~ A1) - KPn > dt (5-4)

where S3 = cox - co2. The populations of the states 11), 12), and 13) are given by the following Bloch equations, respectively [96]: - ^ = -2Qi I m ( p 0 1 ) - 2 J I m ( p 1 2 ) - y w p n , dt dp22 dt dp3i dt
YnP-n-YwPyi-

(5.5)

= -2Q 2 Im(p 02 ) + 2Jlm(pn)

-(y23 + y20)p22,

(5.6)

(5-7)

The states 11) and 12) are excited resonantly, as oppose to the state 13). The processes of radiative recombination (relaxation to the system ground state 10)) are described by the 95

relaxation rates yl0. The relaxation from the QD excited state 12) to the QD ground state 13) accompanied by the phonon emission is described by the rate y2i. Comparison of the experimental PLE results with the numerically calculated data from the optical Bloch equations in the steady-state regime (dp^/dt^O) is shown in Figure 5.4(a). In these

calculations, the relaxation rate y2i was considered as the fastest rate among the other

cfsp * - in(|J|2) + const.


0 5 S" 10 15 20

-100-50

50 100

-100-50

50 100

Detuning from QW resonance (meV)


Figure 5.4 (a) Normalized PLE intensity (open circles) detected at the QD ground state emission wavelength of the QD-QW samples under resonant QW excitation as a function of GaAs spacer thickness dsp (bottom horizontal axis). PLE intensity calculated from the Bloch equation model (solid line) versus -ln(|J|2) (where J is a coupling constant), (b) QD PLE intensity map as a function of spectral detuning from the QW resonance (horizontal axis) and term -ln(|J|2) (vertical axis) calculated using Bloch equation model, (c) Calculated QD PLE spectra (horizontal cross-sections of the map in Fig. 5.4(b)) in the QW resonance region taken at different values of -ln(|jf). [96] 96

relaxation rates in the system. Indeed, it was reported that the time of carrier relaxation from the highest excited state to the ground state within the InAs QD was as short as several picoseconds [100,101,102]. The PL decay times measured in the QW ground state and in the QD ground state were -550 ps and ~1 ns, respectively [79]. The coupling strength J used in the calculations was correlated with the spacer thickness dsp by taking

in the Fermi's golden rule, from where dsp = -2 ln(|J|) + constant (top horizontal

axis in Figure 5.4(a)). In the model, it was assumed that the states |l) and |2) were exactly in resonance; however, in the real set of samples used in this study the strain and slight variations in growth conditions summed up to an ensemble-averaged mismatch between the levels no more than 15 meV. When this value of level misalignment was introduced into the calculations, the variation was within 5% and it did not make a difference in the qualitative result shown in Figure 5.4(a). It is seen from Figure 5.4(a) that the model reproduces the experimental QD PLE data points as a function of spacer thickness showing saturation and decrease of the calculated QD PLE signal for thin spacer thicknesses. Better understanding of the QD PLE behavior can be achieved from the analysis of the two-dimensional map shown in Figure 5.4(b) which demonstrates how the calculated QD PLE intensity changes depending on the laser detuning from the resonance with the QW exciton state, as well as on the tunnel coupling term -ln(|J| 2 ). The coupling term is directly related to the barrier thickness between the QW and QDs, as has been discussed above. The calculated PLE spectrum profiles around the QW absorption resonance taken at different values of coupling strength -ln(|./|2) are shown in Figure 5.4(c). For 20 run spacer thickness, the coupling between the QW and QD was 97

practically absent and the tunnel coupling strength J was small (-ln(|J|2) = 37.1). The QD PLE spectrum showed only an absorption peak of the QD excited state. At this distance between QW and QDs (20 nm), the radiative recombination time in the QW Tio was shorter than the tunneling time IT and therefore all the carriers recombined within the well. With thinner spacers, the tunnel coupling strength J increased (-ln(|J|2) decreased) and the tunneling time IT became comparable with Tio, which made it possible for part of the carriers to tunnel to the QDs. Thus, the PLE spectrum was a superposition of both broad QD excited state absorption and sharp QW ground state absorption peaks (see curve denoted as 21.4 in Figure 5.4(c)). As the spacer thickness decreased, the number of carriers that tunneled through the barrier increased exponentially, as did the QD PLE intensity. When IT became shorter than Tio, all carriers optically excited in the QW tunneled to the QDs and the PLE signal reached saturation. The width of this plateau depended on the values of Tio and T23. In this weak coupling regime, the carriers tunneled to the QD excited state and quickly relaxed down to the QD ground state, since this relaxation time T23 was shorter than the tunneling time TT and recombination time Ti0 in the QW. For the spacers thinner than 5 nm, the tunneling time TT approached the relaxation time T23 between the QD levels 12) and 13), and below 3 nm these quantities became comparable. This meant that there was a finite probability for carriers to tunnel back and forth between the lowest QW state |l) and the QD excited state |2^, so-called Rabi oscillations. This was a direct result of the QW and QD state hybridization. In this intermediate coupling regime (J ~ hj23) the Rabi oscillations were strongly damped and the PLE signal decreased due to broadening of the resonance (curve with -ln(|J| ) = 0.7 in 98

Figure 5.4(c)). In the strong coupling regime, when -ln(|J| ) < 1, the hybridization of the states |l) and |2) was so strong that it caused a splitting of the PLE resonance as shown in Figure 5.4(c) for -ln(|J|2) = -2.5 [96]. Apparently, this strong tunnel coupling limit was not reached and distinct PLE splittings were not observed for these QD-QW samples. This could be related to the fact that one was not dealing here with a discrete state of a single QD but rather with an ensemble of the QDs varying in sizes, which resulted in broadening of the QD states that coherently coupled with the QW state. A vivid picture that summarizes different coupling regimes experimentally observed in the spacer thickness dependence of the QD PLE signal at resonant QW excitation of the QD-QW system can be presented schematically as shown in Figure 5.5. TwOTgtingh time T T ctecregses
T23~tT ^23<'CT<X10 TT ~ T10 XT > Tio

5 10 15 20 GaAs spacer thickness dsp (nm)


Strong ; coupling No _j coupling V Weak coupling ;

V.

Intermediate coupling

Figure 5.5 Schematic analysis of the QD PLE dependence on the spacer thickness. Different regions of the curve describe various coupling regimes that depend on correlation between the tunneling time and other relaxation times in the QD-QW system. 99

5.3 Resonant tunneling time measurement The QD PLE dependence on spacer thickness provided qualitative information about the carrier dynamics in different coupling regimes and suggested that tunneling times as short as a few picoseconds should be observed for spacers thinner than 3 nm. The time-resolved PL measurements performed under nonresonant excitation, which were discussed in previous chapter, did not provide the time resolution needed to extract such short times. Another technique, ultrafast pump-probe reflection/absorption

spectroscopy, can give a resolution of -200 femtoseconds and allows measurements under resonant excitation conditions. Pump and probe laser pulses perpendicularly polarized to each other were spectrally tuned to a wavelength (-920 nm) which corresponded to the QW exciton energy Ex. The time duration of each pulse was 200 fs with a 10 nm spectral width. The results of the pump-probe measurements on a QD-QW sample with 5 nm spacer are presented in Figure 5.6. The differential reflectivity spectra

AR/R (xio4) 7.0 _ ;

-2.0 ^ ^ ^ ^ ^ ^ ^ ^ ^ ~1 912 914 916 918 920 Probe wavelength (nm)

50 100 150 Time delay (ps)

Figure 5.6 (a) Differential reflectivity spectra of the QD-QW sample with 5 nm GaAs spacer thickness as a function of time delay between pump and probe pulses, (b) Time dependence of the differential reflectivity probed at the QW exciton absorption wavelength (open circles) and a fit of experimental data using a single exponential decay (solid line). [96] 100

defined as AR(A,At) R0(A)


=

R(A,At)-R0(A) R0(A)

are plotted as a function of time delay with the color being an amplitude of ARIR0 signal (Figure 5.6(a)). Spectral features observed at negative time delays (At<0) were

attributed to the perturbed free-induction decay of the QW exciton polarization [103,104,105]. At positive time delays (At>0), the bleaching of the QW exciton

differential reflectivity signal was the dominant feature on the AR IRQ intensity map. The shape of a spectral line taken at a particular positive delay (for example 1 ps) was essentially the same as the shape of the QW resonance in the QD PLE spectrum. The time evolution of the AR IR0 signal recorded at the QW exciton resonance (A. = hclEx), depicted in Figure 5.6(b) with open circles, was fitted with the single-exponential decay function (solid line in Figure 5.6(b)) which allowed extracting of the QW exciton lifetime. Additional small peaks seen at delay times of 15 ps were caused by the reflection of the pump and probe pulses from the back side of the sample. The QW exciton lifetime of 25 ps was determined from the fit of the time dependence of AR IR0 for the QD-QW sample with 5 nm spacer. The lifetime measured on a reference QW sample in this work (-500 ps) and the QW lifetime known from literature [106] were much longer than this lifetime obtained from the resonant excitation measurement. Therefore, the conclusion was made that the carrier dynamics in the QW was strongly affected by the fast tunneling processes that suggested the tunneling time of -26 ps. Such a short tunneling time was in agreement with the PLE results of this work. According to

101

this result, the tunneling time for the QD-QW samples with the thinnest spacers could reach 1 ps as predicted by the optical Bloch equation model. Another important conclusion was drawn from the pump-probe results. At the beginning, an excitonic tunneling was assumed for the simplicity of modeling. The cw PL and PLE experiments did not provide such information as what kind of particles take part in the tunneling process. However, the tunneling time of 26 ps obtained from the ultrafast technique strongly suggested that it was electrons that tunneled in this system, not holes, as the tunneling times calculated for holes were several orders of magnitude longer due to the heavy effective mass [16,56].

5.4 Hybridization of the QD and QW excited states The QD-QW system studied in this work was designed in such a way that the QW exciton ground state and the 3 rd QD excited state appeared to be in resonance. If two nanostructures with energy levels being in resonance are only several nanometers apart the wavefunctions of resonant states will mix and hybridize, resulting in splitting of energy states. The magnitude of this splitting depends on a thickness of a separation barrier [23]. This resonant alignment of energy levels should lead to enhanced tunnel coupling between the QW and QDs which strongly depends on the separation distance between the nanostructures. The PL spectra of one of the QD-QW samples (c/sp =11 nm) measured at various excitation intensities (Figure 5.7(a)) showed that the QW PL emission wavelength /lQW was spectrally located where the emission of the 3 rd QD excited state AQD should reside.

102

900

950
5

1000 1050 Wavelength (nm)


1

1100

s
104 Signal (arb.
i

A'QW

H
i . ... 1 .
JL

\ (b)

(0

c
V) a

*
det QW

LU -J

7 IV
ri .T 1

o
N

820

840

860 880 900 920 Wavelength (nm)

- ^ 1 '

- Wafj
940

Figure 5.7 (a) PL spectra of the QD-QW sample with dsp= \\ nm measured at various excitation intensities (Xexc = 532 nm) with identified PL peaks from QW and QD states, (b) PLE spectra of the QD-QW sample with dsp = 11 nm detected at the PL maximum of the QD ground state (open circles) and of the QW ground state (filled circles). The QW PL spectra taken at low (7 ex c = 10 -4 ^) and high (7exc = 2.5/o) excitation intensities of the same QD-QW sample are shown for comparison, /o = 1000 W/cm2, T= 10 K. [107] The PLE spectra gave a direct insight into the energy/carrier transfer in the QD-QW coupled system. Figure 5.7(b) illustrates the PLE spectra measured on the same QD-QW sample as shown in Figure 5.7(a). The detection of the PLE signal was taken at the QD ground state emission wavelength (shown with open circles in Figure 5.7(b)) and at the 103

QW ground state emission wavelength (shown with filled circles in Figure 5.7(b)). The PL spectra recorded at two different excitation intensities are also shown in Figure 5.7(b) to help identify the QW absorption resonances AQW and fiQW that were seen on the PLE spectra at 923 nm and 874 nm, respectively [108]. The rest of the PLE resonances present in the spectrum were identified as a GaAs free exciton absorption (at 818 nm) and the WL light hole (at 832 nm) and heavy hole (at 852 nm) exciton absorptions [109]. Presence of these resonances in the PLE signal detected at the QD ground state energy meant that optically created excitations in the GaAs matrix were being effectively transferred through the WL into the QW and QD states. The QW resonances seen in the QD PLE signal gave a direct evidence of coupling and carrier transfer from the QW to the QDs, and their intensity was related to the strength of coupling showing strong dependence on the GaAs spacer thickness dsp. The spacer thickness dependences of the PLE spectra detected at AQD and XQW of the QD-QW structures are shown in Figure 5.8(a) and (b), respectively. The PLE intensity of the QW ground state exciton versus dsp was already discussed in details in Sections 5.1 and 5.2. Therefore, here more attention was paid to the behavior of the PLE resonance of the second QW subband observed at JilQW. The probability of carrier tunneling was vanishingly low in the sample with a 20 nm thick spacer, which was reflected in weak intensities of both QW PLE resonances detected in the QD emission (Figure 5.8(a)). As the spacer thickness decreased, the QW PLE signal for both QW subbands increased at first, then reached a saturation, and finally started to decrease for GaAs spacers thinner than 5 nm. The behavior of the QW excited state resonance at A\w (Figure 5.8(c)) was essentially the

104

Figure 5.8 (a) PLE spectra detected at the QD ground state emission wavelength of QD-QW samples with various GaAs spacer thicknesses. PLE spectrum of the reference QD sample is shown for comparison. The spectra are vertically shifted for clarity, (b) PLE spectra detected at the QW ground state emission wavelength of QD-QW samples with various GaAs spacer thicknesses. PLE spectrum of the reference QW sample is shown for comparison. All spectra were measured under the same excitation conditions at temperature T = 10 K. (c) PLE peak intensities and (d) peak splitting values measured at /LlQW as a function of GaAs spacer thickness. [107]

105

same as that of the QW ground state exciton at AQW (Figure 5.4(a)) [96]. The PLE resonance of the QW ground state maintained its spectral shape and only slightly broadened and redshifted for spacers less than 3 nm thick. This broadening was a sign of the stronger coupling which meant that a wavefunction mixing and a state hybridization started to take place for thin spacers. The red shift could be caused by the strain due to the close proximity of the QDs. The spectral shape of the QW excited state observed at fiQW showed doublet splitting for spacers thinner than 3 nm. The magnitude of the peak splitting is plotted as a function of GaAs spacer thickness in Figure 5.8(d). In case of resonantly coupled electronic states separated by a thin potential barrier, a coherent tunneling of carriers occurs between these two states. If the tunneling rate is higher than the dephasing rates of the excitations on each of the two states (strong coupling regime), then the carriers will oscillate between these two coupled states with a period TR -rcl' COR, where a>R is the Rabi frequency. This strong coupling results in derealization of the carrier wavefunctions and the state splitting. The energy of the peak splitting in the PLE spectra was directly related to the Rabi period giving the coherent transfer times of 230, 115, and 160 fs for spacer thicknesses dsp of 3, 2, and 1.5 nm, respectively [107]. Subpicosecond tunneling times clearly indicated that this process had to do with the electronic tunneling only, whereas the holes could be excluded as their tunneling times were expected to be several orders of magnitude longer [56]. This difference in the coupling strength observed in the QW ground state and first excited state can be explained using the following fact: the wavefunction of the QW excited state penetrated farther into the

106

barrier than that of the QW ground state and, hence, overlaped more with the QD wavefunction [107]. An interesting observation came from the PLE spectra detected at the QW ground state emission wavelength tfQW shown in Figure 5.8(b). Appearance of the WL light hole and heavy hole resonances in the spectra indicated the existence of coupling between the WL and QW (as in coupled asymmetric double QW system) and showed the possibility of nonresonant carrier transfer from the WL states to the QW ground state. Another way to demonstrate the change of tunnel coupling strength was to insert an effective barrier of different thicknesses into the spacer which thickness was kept constant. The PLE spectra of such a system with AlAs effective barrier embedded in 8 nm thick GaAs spacer are shown in Figure 5.9. For the QD-QW sample with no AlAs layer
G?AIAS
=

0, the intense QW absorption resonances reflected the high efficiency of

tunnel coupling between the QW and QDs. As the AlAs layers with increasing thickness were inserted, the QW peaks were gradually quenched in the QD PLE spectra showing the tunneling became effectively suppressed. On the other hand, an increased QW PLE signal along with the disappearance of the WL resonances observed for the QD-QW sample with 4 nm thick AlAs barrier (Figure 5.9(b)) reflected an effective QW ground state emission indicating practically no coupling to the QD layer [107].

107

QD-QW (cf = c/naAe + cAlAs L c = 8 nm)

840

880

920 840 Wavelength (nm)

880

Figure 5.9 (a) PLE spectra detected at the QD ground state emission wavelength of QD-QW samples with constant 8 nm thick GaAs spacer dsp that contained AlAs effective barrier ^AIAS of 1 -4 nm thickness symmetrically inserted in the GaAs layer, (b) PLE spectra of the same set of samples as in panel (a) detected at the QW ground state emission wavelength. PLE spectra of the reference QD and reference QW samples are shown for comparison. All spectra were measured under the same excitation conditions at temperature T= 10 K. The spectra are vertically shifted for clarity. [107]

108

5.5 Anti-crossing between WL and QW states tuned in resonance by QW composition variation In Section 4.6, it was shown that the energy position of the QW states could be tuned by composition (indium content) variation in the InxGai_xAs QW. In such a way one can choose which state in the QD layer should be aligned in resonance with the QW ground state. The QD-QW sample with indium content of x = 0.7 in the QW (sample 4 in Figure 4.16) was of particular interest because its QW exciton energy EQW = 1.422 eV was very close to the energy of the heavy-hole {hh) exciton transition in the WL of the reference QD sample E^hhl = 1.427 eV. The light-hole (Ih) exciton transition in the WL of the reference QD sample was found at E^'hl = 1.485 eV. In this case, the QW and

WL formed an asymmetric double QW system with efficient near resonant tunnel coupling through a 4 nm thick GaAs spacer. Distinct evidence of this direct tunnel coupling between the WL and QW was obtained from the PLE spectra of the QD-QW samples shown in Figure 5.10. All spectra were measured at the wavelength of the QD ground state emission AQD for each sample under the same excitation conditions at temperature T= 10 K. Each spectrum showed intense exciton absorption resonance of the GaAs matrix at 818 nm. The same intensity of these peaks indicated that the carrier trapping from the GaAs into the QDs was not affected by the embedding of the QW layers with different compositions. The PLE spectrum measured on the reference QD sample showed clear resonances from the Ih and hh exciton absorptions at 833 and 861 nm, respectively. The WL hh exciton peak was observed in the PLE spectra of each QD-QW sample. Its resonance exhibited broadening

109

820

840

860

880

900

920 940

Wavelength (nm)
Figure 5.10 PLE spectra detected at the QD ground state emission wavelength of the QD-QW samples with various InxGai_xAs QW composition and 4 nm thick GaAs spacer. PLE spectrum of the reference QD sample is shown for comparison. The spectra were measured at the same excitation intensity Iexc = 1.5><10~67o and vertically shifted for clarity. 70 = 1000W/cm2, r = 1 0 K . and slight red shift (shown with a broken solid line in Figure 5.10) from 1.44 eV in the reference QD sample to 1.41 eV in the sample 1 with x = 0.18. In addition, the PLE
e spectra of the hybrid QD-QW samples displayed clear peaks at the ground Eow and QW e

first excited EQW

exciton resonances of the InxGai_xAs QW. As expected, these

resonance energies redshifted when increasing x, as indicated by the solid lines in Figure 5.10. The shapes and intensities of the EQW e resonances were essentially the same (within 5% error) for samples 1, 2, and 3. However, for sample 4 the PLE spectrum 110

appeared to be more complicated. Here, the energies of the WL hh transition and the QW ground state exciton resonance were almost coincident. The multiple peaks found within this energy range, indicated by the dashed lines in Figure 5.10, can be interpreted as a quantum mechanical anti-crossing between these two states. The energy splitting of this anti-crossing was ~7 + 1 meV, which corresponded to electron tunneling times of a few hundred femtoseconds between the WL and QW states in this sample. The resonant QW and WL electronic states became strongly mixed, hybridized, suggesting that this QWWL double layer essentially acted as a "super-wetting layer" which efficiently traped carriers in either the GaAs barrier, the QW or the WL. Subsequently, the carriers were injected into the QD states allowing for recombination. This apparent, pronounced hybridization made this sample interesting for time-resolved experiments aimed at tracking the microscopic carrier relaxation dynamics in the tunnel-coupled QD-QW hybrid nanostructures in real time. The results of a preliminary time-resolved PL measurement carried out on these QD-QW samples with various QW compositions are shown in Figure 5.11(a). The samples were excited with laser pulses at a nonresonant wavelength of 750 nm, i.e. above the energy of the GaAs bandgap. The PL transients were detected at the maximum of the QW emission band. The detection wavelength was varied between 930 nm and 863 nm in samples 1 to 4 according to the composition induced shift. The excitation intensity was chosen sufficiently low in order to insure that the state-filling effects did not affect the relaxation dynamics. The transient curves were fitted by a simple monoexponential decay model, convoluted with a 15 ps instrument response function. Figure 5.11(b) shows the QW PL decay times extracted from the fit versus the QW spectral position. The value of

111

100

200

300

400

500

Time (ps) w 70 L
Q
X

QW WLrefQD

a.

>

o 60
s u 50 40 30

/ ^

\ r

4 y
880

1: 2: 3: 4:

x=0.18 x=0.15 x=0.10 x=0.07

(b)
i . i

860

900

920

940

Wavelength (nm)
Figure 5.11 (a) PL transients recorded at the wavelength of the QW exciton emission in the QD-QW samples with various QW composition. The laser pulse used for excitation of time-resolved PL (Xexc = 750 nm, peak intensity Ip = 1.3x104 W/cm2) is shown for comparison. The experimental data were fitted using single exponential decay function (solid lines). The curves were normalized and vertically shifted for clarity, (b) PL decay times (filled circles) as a function of the QW exciton emission wavelength along with the PL decay time measured in the WL of the reference QD sample (open circle). T= 10 K.

112

the WL PL decay time measured on the reference QD sample is also shown on the plot. It was seen that the values xQW decreased from 70 ps (for sample 1) to 28 ps (for sample 4) as the QW exciton energy EQW tended towards the WL energy E^L, ultimately approaching the same PL decay time in the WL (T^ L =30ps). This was another indication of the state hybridization in sample 4, where tunnel-coupled states acted as one super-WL system. Evidently, the time resolution of such measurements using the streak camera and nonresonant excitation was limited by the finite instrument response time (15 ps here). Also, it was limited by rather slow carrier relaxation dynamics from the GaAs barrier down to the WL and QD states. This clearly calls for experiments with selective resonant excitation of the WL, QW and QD states to further elucidate the microscopic carrier tunneling dynamics in such QD-QW hybrid structures. The PLE results, specifically the observed QW-WL hybridization in Figure 5.10, suggest that such experiments under resonant excitation should be performed with a time resolution of at least 100 fs and may reveal evidence for coherent tunneling dynamics even in QD ensemble measurements.

113

5.6 Summary In this Chapter, the coherent aspects of tunnel coupling in the QD-QW structure were explored. The PLE experiments were used to gain direct insight into coupling between the QW and QD states. Nonmonotonic behavior of the spacer-thickness dependence of the QD PLE signal under resonant QW excitation was observed and explained in terms of resonant carrier tunneling between the QW exciton ground state and the 3 rd QD excited state. In QD-QW samples with thin GaAs spacers (<isp < 5 nm) resonant tunneling induced hybridization of states which led to broadening and intensity decrease of the PLE resonance of the QW exciton transition. The experimental data were successfully described by the optical Bloch equation model and the PLE intensity drop for thin spacers was ascribed to the intermediate coherent coupling regime between the QW lowest state and the 3rd QD excited state. Fast tunneling time of -26 ps deduced from the pump-probe experiment for QD-QW sample with 5 nm spacer indicated that the electron resonant tunneling was primarily observed in this system, whereas the hole tunneling was negligible. It was also shown that for the QW excited state with its broader wavefunction a strong coherent tunneling regime was reached in the samples with dsp < 3 nm, resulting in state splitting and derealization of the electron wavefunctions between the QW and QD. Strong hybridization and mixing of states was also observed when the QW ground state was tuned into resonance with the WL state by varying indium concentration in the InxGai_xAs QW. Ultrafast coherent tunneling with -100 fs time resulted from this strong coupling which transformed the QW-WL system into a super-WL. This super-WL effectively collected the carriers and transferred them to the QDs, considerably enhancing their exciton emission.

114

VI. CONCLUSIONS AND OUTLOOK

The QD-QW hybrid structure, which consists of OD nanostructure (QD) and 2D nanostructure (QW) separated by a potential barrier (spacer), is a very convenient system to study mechanisms of coupling between nanostructures with different dimensionalities. The efficiency of coupling strongly depends on the barrier thickness and/or height and the energy level alignment between the two nanostructures. For a systematic study of the QD-QW structure, a number of InAs/GaAs QDs InxGai_xAs/GaAs Q W samples with barriers of various thicknesses and heights as well as with different indium concentration in the QW layer were grown by the MBE technique. The QD-QW samples were investigated theoretically and experimentally using various methods: continuous-wave PL, time-resolved PL, and PL excitation spectroscopy; TEM and AFM microscopy. It was shown that for thick separation spacers the coupling was weak and the exciton emission peaks from both QW and QDs were observed in the PL spectrum even at low excitation intensity. For thin separation spacers, on the other hand, the coupling was strong, the carriers were effectively transferred from the QW to the QDs and only QD ground state emission was detected in the PL spectrum under low excitation conditions. The results of the PL measurements at various excitation intensities on the QD-QW samples with different barrier thickness and height were analyzed using the rate equation model. The main coupling mechanism in these QD-QW samples deduced from the analysis of PL results was the resonant tunneling of electrons from the QW ground state to the third QD excited state. The tunneling was strongly affected by the QD state

115

filling at high excitation intensities and was completely blocked if all QD states were populated so that the additional carriers from the QW could not be accepted due to the Pauli exclusion principle. The tunneling times extracted from the ratio of the QW and QD integrated PL intensities showed an exponential dependence on the spacer thickness which was in agreement with the Wentzel-Kramers-Brillouin approximation. For spacers thinner than 2 nm, the tunneling times were predicted to be ~1 ps or even shorter. In case of 20 nm thick spacer, obtained tunneling times were longer than 10 ns. The tunneling times obtained from the time-resolved PL measurements under nonresonant excitation also exhibited an exponential dependence on the spacer thickness, however they deviated from the cw PL data. Further analysis suggested that the measurements by this method did not reflect correct values of tunneling times for thin spacers because of several reasons: (i) these measurements included the initial carrier relaxation dynamics from the nonresonant optical excitation into the GaAs barrier down to the coupled QW states; (ii) the lowest possible excitation intensity used for the measurement was still high enough to introduce the effects of the state filling; (iii) the time resolution of the streak camera was limited to 15 ps. For direct measurement of the resonant tunneling times an ultrafast pump-probe technique should be used at resonant excitation of the QW state. Indeed, from the preliminary pump-probe experiment at resonant QW excitation on a QD-QW sample with 5 nm thick spacer the tunneling time of ~26 ps was obtained, which was closer to the value extracted from the cw PL, thus confirming the validity of cw PL results and the rate equation model. It was shown that using different indium concentrations in the InxGai_xAs QW layer the energy position of the QW ground state was tuned across the range of several

116

QD excited states up to the WL exciton state. The excitation intensity required to fill up all the QD states that have lower energy than the QW ground state in order to block the tunneling and promote the appearance of the QW PL peak strongly depended on the energy position of the lowest QW state. The more QD states were located below the energy of the QW state, the higher excitation intensity was needed to "turn on" the QW PL signal. In the QD-QW sample where InAs QDs were directly overgrown by the InGaAs QW layer, the strain-induced QDs formed in the QW on top of the InAs QDs. The lateral confinement was created by the strain field from lattice-mismatched InAs QDs, whereas the vertical confinement was provided by the QW. The exciton emission peaks from the SIQDs appeared below the QW emission energy and above the second QD excited state, and were detected only at the high excitation intensity when all the InAs QD states were populated with carriers. The nonmonotonic dependence of the QD PLE intensity at resonant QW excitation on the spacer thickness was observed in the QD-QW system. The experimental data were analyzed and modeled using optical Bloch equations and simplified four-level energy scheme. The PLE intensity decreased for spacers < 5 nm was ascribed to the hybridization of resonant QW and QD states due to the intermediate coherent tunnel coupling. It was found from the QD PLE spectra that a strong tunnel coupling regime was reached between the QW first excited state and resonant to it QD excited state in the samples with spacers thinner than 3 nm. The broader wavefunction of the QW excited state penetrated further into the barrier allowing for a greater overlap with the

117

wavefunction of the QD excited state, which resulted in derealization and the state splitting observed experimentally in the PLE spectra. In case of tuning the QW ground state in resonance with the WL state by the change of the QW composition the PLE peak splitting caused by the hybridization of the resonant WL and QW states was also experimentally observed. The magnitude of splitting (~7 meV) of the PLE peaks predicted resonant tunneling times on the order of a few hundred femtoseconds. The combination of the resonantly coupled QW and WL formed a super-WL that effectively collected the carriers and supplied them to the QDs. The future prospective work may include the ultrafast pump-probe experiments devoted to resonant tunneling time measurements as well as detection of the Rabi oscillations in the QD-QW samples with spacers thinner than 2 nm. Additional work can be done using the single QD spectroscopy along with the external electric field application on the low density QD-QW samples. Another interesting research field, coherent spin-polarized electron tunneling, could be also explored in this system by means of the polarized optical spectroscopy. Using QD-QW samples with type II nanostructures in the spectroscopic experiments could help separate electrons and holes in the tunneling process so that one can study the details of tunneling of only one type of carriers.

118

References 1. P. Michler (ed.), Single Quantum Dots: Fundamentals, Applications and New Concepts, Topics in Applied Physics Vol.90 (Springer-Verlag, Berlin, 2003). H. Ishikuro, T. Hiramoto, Appl. Phys. Lett. 71, 3691 (1997). L. Zhuang, L. Guo, S. Y. Chou, Appl. Phys. Lett. 72, 1205 (1998). V. Zwiller, H. Blom, P. Jonsson, N. Panev, S. Jappesen, T. Tsegaye, E. Goobar, M.-E. Pistol, L. Samuelson, G. Bjork, Appl. Phys. Lett. 78, 2476 (2001). Z. Yuan, B. E. Kardynal, R. M. Stevenson, A. J. Shields, C. J. Lobo, K. Cooper, N. S. Beattie, D. A. Ritchie, M. Pepper, Science 295, 102 (2002). M. H. Baier, E. Pelucchi, E. Kapon, S. Varoutsis, M. Gallart, I. Robert-Philip, I. Abram, Appl. Phys. Lett. 84, 648 (2004). H. Drexler, D. Leonard, W. Honsen, H. P. Kotthaus, P. M. Petroff, Phys. Rev. Lett. 73, 2252 (1994). S. Cortez, O. Krebs, S. Laurent, M. Senes, X. Marie, P. Voisin, R. Ferreira, G. Bastard, J.-M. Gerard, T. Amand, Phys. Rev. Lett. 89, 207401 (2002). M. Kroutvar, Y. Ducommun, D. Heiss, M. Bichler, D. Schuh, G. Abstreiter, J. J. Finley, Nature 432, 81 (2004).

2. 3. 4.

5.

6. 7.

8.

9.

10. A. Imamoglu, D. D. Awschalom, G. Burkard, D. P. Di Vincenzo, D. Loss, M. Sherwin, A. Small, Phys. Rev. Lett. 83, 4204 (1999). 11. I. Fushman, D. Englund, A. Faraon, N. Stoltz, P. Petroff, J. Vuckovic, Science 320, 769 (2008). 12. J.-J. Li, K.-D. Zhu, Nanotechnology 22, 055202 (2011). 13. E. Ozbay, Science 311, 189 (2006). 14. H. J. Krenner, M. Sabathil, E. C. Clark, A. Kress, D. Schuh, M. Bichler, G. Abstreiter, J. J. Finley, Phys. Rev. Lett. 94, 057402 (2005). 15. E. A. Stinaff, M. Scheibner, A. S. Bracker, I. V. Ponomarev, V. L. Korenev, M. E. Ware, M. F. Doty, T. L. Reinecke, D. Gammon, Science 311, 636 (2006). 16. S. W. Chang, S. L. Chuang, N. Holonyak Jr., Phys. Rev. B 70, 125312 (2004). 119

17. M. Reischle, G. J. Beirne, R. RoBbach, M. Jetter, H. Schweizer, P. Michler, Phys. Rev. B 76, 085338 (2007). 18. A. O. Govorov, Phys. Rev. B 68, 075315 (2003). 19. T. Unold, K. Mueller, C. Lienau, T. Elsaesser, A. D. Wieck, Phys. Rev. Lett. 94, 137404 (2005). 20. M. Scheibner, T. Schmidt, L. Worschech, A. Forchel, G. Bacher, T. Passow, D. Hommel, Nat. Phys. 3, 106 (2007). 21. Yu. I. Mazur, V. G. Dorogan, E. Marega Jr., G. G. Tarasov, D. F. Cesar, V. LopezRichard, G. E. Marques, G. J. Salamo, Appl. Phys. Lett. 94, 123112 (2009). 22. G. Parascandolo, V. Savona, Phys. Rev. B 71, 045335 (2005). 23. R. Ferreira, G. Bastard, Rep. Prog. Phys. 60, 345 (1997). 24. K. Leo, J. Shah, E. O. Gobel, T. C. Damen, S. Schmitt-Rink, W. Schafer, K. Kohler, Phys. Rev. Lett. 66, 201 (1991). 25. C. Juang, Phys. Rev. B 44, 10706 (1991). 26. P. Harrison, Quantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor Nanostructures, 2n edition (John Wiley & Sons, Chichester, UK, 2005) p.99. 27. K. Leo, J. Shah, E. O. Gobel, J. P. Gordon, S. Schmitt-Rink, Semicond. Sci. Technol. 7, B394 (1992). 28. D. Y. Oberli, J. Shah, T. C. Damen, C. W. Tu, T. Y. Chang, D. A. B. Miller, J. E. Henry, R. F. Kopf, N. Sauer, A. E. DiGiovanni, Phys Rev. B 40, 3028 (1989). 29. R. Sauer, T. D. Harris, W. T. Tsang, Phys. Rev. B 39, 12929 (1989). 30. R. Ferreira, P. Rolland, Ph. Roussingol, C. Delalande, A. Vinattieri, L. Carraresi, M. Colocci, N. Roy, B. Sermage, J. F. Palmier, B. Etienne, Phys. Rev. B 45, 11782 (1992). 31. S. Haacke, N. T. Pelekanos, H. Mariette, M. Zigone, A. P. Heberle, W. W. Rtihle, Phys. Rev. B 47, 16643 (1993). 32. M. F. Krol, S. Ten, B. P. McGinnis, M. J. Hayduk, G. Khitrova, N. Peyghambarian, Phys. Rev. B 52, R14344 (1995). 33. R. Sauer, K. Thonke, W. T. Tsang, Phys. Rev. Lett. 61, 609 (1988). 120

34. T. H. Wang, X. B. Mei, C. Jiang, Y. Huang, J. M. Zhou, X. G. Huang, C. G. Cai, Z. X. Yu, C. P. Luo, J. Y. Xu, Z. Y. Xu, Phys. Rev. B 46, 16160 (1992). 35. D. H. Levi, D. R. Wake, M. V. Klein, S. Kumar, H. Morkoc, Phys. Rev. B 45, 4274 (1992). 36. I. N. Stranski, L. Krastanov, Sitzungsber. Akad. Wiss. Wien, Math.-Naturwiss. Kl. lib 146, 797 (1938). 37. B. L. Liang, Zh. M. Wang, J. H. Lee, K. A. Sablon, Yu. I. Mazur, G. J. Salamo, Appl. Phys. Lett. 89, 043113 (2006). 38. P. M. Lytvyn, Yu. I. Mazur, E. Marega Jr., V. G. Dorogan, V. P. Kladko, M. V. Slobodyan, V. V. Strelchuk, M. L. Hussein, M. E. Ware, G. J. Salamo, Nanotechnology 19, 505605 (2008). 39. J. H. Lee, Zh. M. Wang, V. G. Dorogan, Yu. I. Mazur, M. E. Ware, G. J. Salamo, J. Appl. Phys. 106, 073106 (2009). 40. Yu. I. Mazur, W. Q. Ma, X. Wang, Z. M. Wang, G. J. Salamo, M. Xiao, T. D. Mishima, M. B. Johnson, Appl. Phys. Lett. 83, 987 (2003). 41. Zh. M. Wang, B. L. Liang, K. A. Sablon, J. H. Lee, Yu. I. Mazur, N. W. Strom, G. J. Salamo, Small 3, 235 (2007). 42. B. L. Liang, V. G. Dorogan, Yu. I. Mazur, N. W. Strom, J. H. Lee, K. A. Sablon, Zh. M. Wang, G. J. Salamo, J. Nanosci. Nanotechnol. 9, 3320 (2009). 43. D. L. Huffaker, D. G. Deppe, Appl. Phys. Lett. 73, 366 (1998). 44. J. W. Tomm, T. Elsaesser, Yu. I. Mazur, H. Kissel, G. G. Tarasov, Z. Ya. Zhuchenko, W. T. Masselink, Phis. Rev. B 67, 045326 (2003). 45. A. Tackeuchi, T. Kuroda, K. Mase, Y. Nakata, N. Yokoyama, Phys. Rev. B 62, 1568 (2000). 46. Yu. I. Mazur, Zh. M. Wang, G. G. Tarasov, Vas. P. Kunets, G. J. Salamo, Z. Ya. Zhuchenko, H. Kissel, J. Appl. Phys. 98, 053515 (2005). 47. Yu. I. Mazur, Zh. M. Wang, G. G. Tarasov, M. Xiao, G. J. Salamo, J. W. Tomm, V. Talalaev, H. K i s s e l , ^ / . Phys. Lett. 86, 063102 (2005). 48. B. L. Liang, Zh. M. Wang, Yu. I. Mazur, G. J. Salamo, E. A. DeCuir Jr., M. O. Manasreh, Appl. Phys. Lett. 89, 043125 (2006).

121

49. T. Unold, K. Meuller, C. Lienau, T. Elsaesser, A. D. Wieck, Phys. Rev. Lett. 94, 137404 (2005). 50. E. A. Stinaff, M. Scheibner, A. S. Bracker, I. V. Ponomarev, V. L. Korenev, M. E. Ware, M. F. Doty, T. L. Reinecke, D. Gammon, Science 311, 636 (2006). 51. A. S. Bracker, M. Scheibner, M. F. Doty, E. A. Stinaff, I. V. Ponomarev, J. C. Kim, L. J. Whitman, T. L. Reinecke, D. Gammon, Appl. Phys. Lett. 89, 233110 (2006). 52. W.-H. Chang, H. Lin, S.-Y. Wang, C.-H. Lin, S.-J. Cheng, M.-C. Lee, W.-Y. Chen, T.-M. Hsu, T.-P. Hsieh, J.-I. Chyi, Phys. Rev. B11, 245314 (2008). 53. M. Reichle, G. J. Beirne, R. RoBbach, M. Jetter, H. Shcweizer, P. Michler, Phys. Rev. B 76, 085338 (2007). 54. M. Bayer, P. Hawrylak, K. Hinzer, S. Fafard, M. Korkusinski, Z. R. Wasilewski, O. Stern, A. Forchel, Science 291, 451 (2001). 55. J. - P. Leburton, S. Nagaraja, P. Matagne, R. M. Martin, Microelectron. J. 34, 485 (2003). 56. S. L. Chuang, N. Holonyak Jr., Appl Phys. Lett. 80, 1270 (2002). 57. Yu. I. Mazur, B. L. Liang, Zh. M. Wang, D. Guzun, G. J. Salamo, Z. Ya. Zhuchenko, G. G. Tarasov, Appl. Phys. Lett. 89, 151914 (2006). 58. Yu. I. Mazur, B. L. Liang, Zh. M. Wang, G. G. Tarasov, D. Guzun, G. J. Salamo, T. D. Mishima, M. B. Johnson, J. Appl. Phys. 100, 054313 (2006). 59. M. Syperek, P. Leszczyhski, J. Misiewicz, E. M. Pavelescu, C. Gilfert, J. P. Reithmaier, Appl. Phys. Lett. 96, 011901 (2010). 60. V. G. Talalaev, J. W. Tomm, N. D. Zakharov, P. Werner, U. Gosele, B. V. Novikov, A. S. Sokolov, Y. B. Samsonenko, V. A. Egorov, G. E. Cirlin, Appl. Phys. Lett. 93, 031105(2008). 61. M. Achermann, M. A. Petruska, S. Kos, D. L. Smith, D. D. Koleske, V. I. Klimov, Nature 429, 642 (2004). 62. S. Kos, M. Achermann, V. I. Klimov, D. L. Smith, Phys. Rev. B 71, 205309 (2005). 63. F. Boxberg, J. Tulkki, Rep. Prog. Phys. 70, 1425 (2007). 64. J. Riikonen, J. Sormunen, H. Koskenvaara, M. Mattila, A. Aierken, T. Hakkarainen, M. Sopanen, H. Lipsanen, Nanotechnology 17, 2181 (2006).

122

65. J. Sormunen, J. Riikonen, M. Mattila, M. Sopanen, H. Lipsanen, Nanotechnology 16, 1630 (2005). 66. C. Lingk, W. Heifer, G. von Plessen, J. Feldmann, K. Stock, M. W. Feise D. S. Citrin, H. Lipsanen, M. Sopanen, R. Virkkala, J. Tulkki, J. Ahopelto, Phys. ev. 5 62, 13588(2000). 67. J. H. Davies, Appl. Phys. Lett. 75, 4142 (1999). 68. W. V. Schoenfeld, C. Metzner, E. Letts, P. M. Petroff, Phys. Rev. B 63, 205319 (2001). 69. T. Wang, A. Forchel, J. Appl. Phys. 86, 2001 (1999). 70. T. Lundstrom, W. Schoenfeld, H. Lee, P. M. Petroff, Science 286, 2312 (1999). 71. D. L. Huffaker, G. Park, Z. Zou, O. B. Shchekin, D. G. Deppe, Appl. Phys. Lett. 73, 2564 (1998). 72. T. Chung, G. Walter, N. Holonyak Jr., Appl. Phys. Lett. 79, 4500 (2001). 73. G. Walter, T. Chung, N. Holonyak Jr., Appl. Phys. Lett. 80, 1126 (2002). 74. G. Walter, N. Holonyak Jr., J. H. Ryou, R. D. Dupius, Appl. Phys. Lett. 79, 1956 (2001). 75. G. Walter, N. Holonyak Jr., J. H. Ryou, R. D. Dupius, Appl. Phys. Lett. 79, 3215 (2001). 76. Z. Mi, P. Bhattacharya, S. Fathpour, Appl. Phys. Lett. 86, 153109 (2005). 77. W. Rudno-Rudzinski, G. Sek, K. Ryczko, M. Syperek, J. Misiewicz, E. S. Semenova, A. Lemaitre, A. Ramdane, Appl. Phys. Lett. 94, 171906 (2009). 78. S. Nizamoglu, E. Sari, J.-H. Baek, I.-H. Lee, H. V. Demir, New J. Phys. 10, 123001 (2008). 79. Yu. I. Mazur, V. G. Dorogan, E. Marega Jr., Z. Ya. Zhuchenko, M. E. Ware, M. Benamara, G. G. Tarasov, P. Vasa, C. Lienau, G. J. Salamo, J. Appl. Phys. 108, 074316(2010). 80. J. Urayama, T. B. Norris, J. Singh, P. Bhattacharya, Phys. Rev. Lett. 86, 4930 (2001). 81. L. Zhang, T. Boggess, K. Gundogdu, M. E. Flatte, D. G. Deppe, C. Cao, O. B. Shchekin, Appl. Phys. Lett. 79, 3320 (2001). 123

82. M. Grundmann, C. Bimberg, Phys. Rev. B 55, 9740 (1997). 83. S. Malik, E. C. Le Ru, D. Childs, R. Murray, Phys. Rev. B 63, 155313 (2001). 84. S. Raymond, S. Fafard, P. J. Poole, A. Wojs, P. Hawrylak, S. Charbonneau, D. Leonard, R. Leon, P. M. Petroff, J. L. Merz, Phys. Rev. B 54, 11548 (1996). 85. R. Heitz, A. Kalburge, Q. Xie, M. Grundmann, P. Chen, A. Hoffmann, A. Madhukar, D. Bimberg, Phys. Rev. B 57, 9050 (1998). 86. D. Morris, N. Perret, S. Fafard, Appl. Phys. Lett. 75, 3593 (1999). 87. P. Bhattacharya, S. Ghosh, S. Pradhan, J. Singh, Z. K. Wu, J. Urayama, K. Kim, T. B. Norris, IEEE J. Quantum Electron. 39, 952 (2003). 88. S. Trumm, M. Wesseli, H. J. Krenner, D. Schuh, M. Bichler, J. J. Finley, M. Betz, Appl. Phys. Lett. 87, 153113 (2005). 89. J. H. Davies, J. Appl. Phys. 84, 1358 (1998). 90. A. E. Romanov, P. Waltereit, J. S. Speck, J. Appl. Phys. 97, 043708 (2005). 91. M. Grundmann, O. Stier, D. Bimberg, Phys. Rev. B 52, 11969 (1995). 92. J. Tulkki, A. Heinamaki, Phys. Rev. B 52, 8239 (1995). 93. K. Kash, J. M. Worlock, M. D. Sturge, P. Grabbe, J. P. Harbison, A. Scherer, P. S. D. Lin, Appl. Phys. Lett. 53, 782 (1988). 94. K. Kash, R. Bhat, D. D. Mahoney, P. S. D. Lin, A. Scherer, J. M. Worlock, B. P. Van der Gaag, M. Koza, P. Grabbe, Appl. Phys. Lett. 55, 681 (1989). 95. D. Loss, D. P. DiVincenzo, Phys. Rev. A 57, 120 (1998). 96. Yu. I. Mazur, V. G. Dorogan, D. Guzun, E. Marega Jr., G. J. Salamo, G. G. Tarasov, A. O. Govorov, P. Vasa, C. Lienau, Phys. Rev. B 82, 155413 (2010). 97. U. Fano, Phys. Rev. 124, 1866, (1961). 98. M. Kroner, A. O. Govorov, S. Remi, B. Biedermann, S. Seidl, A. Badolato, P. M. Petroff, W. Zhang, R. Barbour, B. D. Gerardot, R. J. Warburton, K. Karrai, Nature (London) 451, 311 (2008).

124

99. P. A. Dalgarno, M. Ediger, B. D. Gerardot, J. M. Smith, S. Seidl, M. Kroner, K. Karrai, P. M. Petroff, A. O. Govorov, R. J. Warburton, Phys. Rev. Lett. 100, 176801 (2008). 100. T. F. Boggess, L. Zhang, D. G. Deppe, D. L. Huffaker, C. Cao, Appl. Phys. Lett. 78, 276 (2001). 101. E. W. Bogaart, R. Notzel, Q. Gong, J. E. M. Havekort, J. H. Wolter, Appl. Phys. Lett. 86, 173109(2005). 102. K. W. Sun, A. Kechiantz, B. C. Lee, C. P. Lee, Appl. Phys. Lett. 88, 163117 (2006). 103. H. Wang, K. Ferrio, D. G. Steel, Y. Z. Hu, R. Binder, S. W. Koch, Phys. Rev. Lett. 71, 1261 (1993). 104. T. Guenther, C. Lienau, T. Elsaesser, M. Glanemann, V. M. Axt, T. Kuhn, S. Eshlaghi, A. D. Wieck, Phys. Rev. Lett. 89, 057401 (2002). 105. J. M. Shacklette, S. T. Cundiff, Phys. Rev. B 66, 045309 (2002). 106. J. Feldmann, G. Peter, E. O. Gobel, P. Dawson, K. Moore, C. Foxon, R. J. Elliott, Phys. Rev. Lett. 59, 2337 (1987). 107. Yu. I. Mazur, V. G. Dorogan, E. Marega Jr., M. Benamara, Z. Ya. Zhuchenko, G. G. Tarasov, C. Lienau, G. J. Salamo, Appl. Phys. Lett. 98, 083118 (2011). 108. R. Heitz, M. Veit, N. N. Ledentsov, A. Hoffmann, D. Bimberg, V. M. Ustinov, P. S. Kop'ev, Zh. I. Alferov, Phys. Rev. B 56, 10435 (1997). 109. M. Hugues, M. Teisseire, J.-M. Chauveau, B. Vinter, B. Damilano, J.-Y. Duboz, J. Massies, Phys. Rev. B 76, 075335 (2007).

125

Appendix A: Derivation of Equations (4.13) and (4.14)

In order to derive Equation (4.13) let us simplify the look of equations and denote

Z^(/)=^.
IJG?(Iexc) = GD(Iexc) = G
Gw{Iexc) = Gw,

I X =nD
in Equations (4.10) and (4.11). Then the steady-state rate equations will look as follows:
=0
V W ' tr J

Cr

Next, nw and nD can be expressed from these equations as


G TwTtr -R . -

w ~

*V+7 r ,

=R

GD +

=T

GD

r^w-R"
T

^
T

W tr
+f

= TR

QD

, ^

W
'tr J

{fw

tr)*tr)

' W ^

For the reference samples rtr = < = o , therefore |rcre/ = G^r f l

K=GDr*
PL intensity can be written as / = , where rR is the time of radiative recombination (without a bar). 126

The PL signals from the QW and QDs of hybrid QD-QW samples are nw rR
L

CLW-R-

QW

(T

+T )TR '

nD
QD

fR

GD+Tw

Then the intensity ratio can be written as


iybr _
1

J QW

fW-RT T W tr

QD +

^J ~w
T

(-iWR

IlQD

(fR +r

)rR

tr .

For the QD and QW reference samples PL intensities are given as:


jref _ L QW
n

y,ref w 1

WR Gw vR '

jref 1 QD

_ ~

"p R

Guf
-R

DR '

and the ratio


jref _
1

jref QW jref I QD

<J
w

^WRR TWT QDRfR 'W'-

Now, inserting Aref in Ahybr one gets


jhybr _ W
T

WTtr

l* f~iWR
R R

WTtr

(f

+f

)TR

QD

*-r ?w *w+*tr
R_R

(f +ftr)r GD

1+
+

r^wR

GD{TW'+Ttr)_
-1

*"

fR

1
(Jw+hr)

A*

+-

T,

(T+rlry

(*W+*,r)

A"1

-R

(TZ+fjT

, =-

->R

which is exactly Equation (4.13) with r' = ^-.

127

Equation (4.14) was derived as follows. In case of Nl nt only the first order terms by nt are preserved in Equation (4.3) and

only intersublevel relaxation between the neighboring QD levels is taken into account. Then Equation (4.3) takes the following form
iN n (G?+T" '- > i i
,
J
w

,+i

= 0

N,(

T mt

Tint

TR

or

fe +r)-|

GD+TW N,

1 1 n+!hL = 0. + +7 s r
int y

If one denotes the highest QD level to which the carriers are transferred from the QW as i = L, then previous equation can be transformed into the set of equations for each QD level

(G+7f)-| G" +?7 J _ J_

nL=0
n

L-l+-

l L - i

(G.2+T?_2)-

( /~<D 'L-2

, rpW ' xL-2


^ 1 -

1 " N
l

+ nL_2 +

L-\

=0

up to i = 1

Solutions of this equation system for each QD state are given by the following set

nT

GD
N,

iD

, rpW rpW ,

D , rpW 1 G+T; - +
'int

+
*

128

G?_,+T" +( ^-1 i" 1

GL +1L

Gr+T" 1 1 " + + ^r
N T T

CD

-l-TW

1 G L_x+Tl, , 1 , P Af r r^
D

(<,+# Gf+2f
r

1 1

(Gf+rf)

G?+T7
#,

1 \ r^o <,+*+J_
#i-i
7

_ +J
T

^m

*" y

mt

and so on until i = 1.

From this set of solutions a general formula (4.14) was deduced.

129

Appendix B: Description of Research for Popular Publication Coupling between semiconductor nanostructures: How nanostructures talk to each other Great progress in creating new electronic and optoelectronic devices has been recently made and expected to evolve even more in a near future owing to discoveries of nanoscience and developments of nanotechnology. At the University of Arkansas in Fayetteville, research group led by Dr. Gregory Salamo, Distinguished Professor in Physics, has been working on growing semiconductor nanostructures and studying their optical and electrical properties for more than a decade. "State-of-the-art molecular beam epitaxy (MBE) is a powerful growth tool that allows to grow semiconductor materials atom by atom", as Dr. Salamo used to say to his students and visitors. "We would like to learn more about how different nanostructures interact with each other, how they talk", Dr. Salamo ones said at the group meeting encouraging students to explore this issue. Semiconductor nanotechnology obtained a head start due to well developed microelectronic industry. The properties of semiconductor nanostructures - quantum well (QW), quantum wire (QWr), and quantum dot (QD) - have been extensively studied and put into use. For example, QDs with their unique atomic-like electron states found a great application potential in light emitting devices. Another huge advantage of nanostructures is that they exploit the laws of quantum mechanics and, therefore, can make possible the creation of a quantum computer, a dream of many scientists. Mr. Vitaliy Dorogan, a PhD student directed by Dr. Salamo, says, "To make this dream come true, researchers must learn how to combine nanostructures with various properties in a controllable fashion and predict behavior of such complex and extremely small systems." First attempts to study interaction (or, in other words, coupling) were made on double layer nanostructures that consisted of either two QWs, two layers of QWrs, or QDs, as shown in Figure B-l. Those studies have shown that coupling occurs when the charge carriers and/or energy are being transferred from one nano structure to another and the strength of this coupling strongly depends on a distance between nanostructures and the mutual alignment of their energy levels. 130

QW

QWr

Figure B-l Double-layer QW, QWr, and QD nanostructures To progress further, the coupling between nanostructures of different

dimensionalities should be studied Dr Yuriy Mazur, a research professor in Dr Salamo's group, proposed to study a hybrid system that combined a single QW with a layer of QDs separated by a thin barrier. Interest to this hybrid nanostructure is driven by its possible application in optoelectronic devices such as lasers, LEDs, and electronic displays etc The QW here serves as additional reservoir and distributor of carriers (Figure B-2) that can be transferred to the QDs improving their light emission characteristics.

Figure B-2 QD-QW system for light emission applications 131

Mr. Vitaliy Dorogan conducted systematic experiments on QD-QW structures using photoluminescence (PL) methods. By resonantly exciting the QW with a tunable laser and watching the PL signal from the QDs the coupling was studied as a function of separation between the nanostructures. As the spacer thickness decreased, the QD PL signal rapidly increased and saturated at about 10 nm separation. But what happened at smaller distances confused both Mr. Dorogan and Dr. Salamo: signal started to drop as the spacer thickness decreased below 5 nm. Increase and saturation of the QD PL signal can be logically explained by the increased tunneling of carriers from the QW to the QDs up to the point when all the carriers have tunneled (saturation). However, decrease of PL signal at smaller distances remained a mystery to the researches for quite some time. Only after careful study of the system's energy level scheme and considering the coherent effects the mystery was solved. Figure B-3 helps to clarify all the processes occurring in a coupled QD-QW hybrid system. Because energy levels |1> and |2> are in resonance, at small separation

QW
InGaAs

QDs
In As

wu ID nnelii
^23

|2>

Laser excitation

T10

|3>

^30

PL signal
|o>

Figure B-3 Energy level scheme of the QD-QW system studied here. 132

distances the coupling becomes so strong that electrons can tunnel back and forth in an oscillatory fashion. Such oscillations are called Rabi oscillations and are only possible in coherent quantum systems that are resonantly excited. Apparently, in this non-ideal system an average electron performs only one oscillation coming back to the QW and thus reducing the PL emission in the QDs. "I was amazed that I could observe the quantum mechanics work in a not so complicated experiment", Mr. Dorogan expressed his excitement about his work, "And it is such a great feeling to know that you can understand and explain the results you see."

Thanks to Dr. Salamo's efforts, nanoscience program at the University of Arkansas was raised on a new level when a cutting-edge Nanoscale Materials Science and Engineering Building opened its doors on campus in Fayetteville. This building was specially designed to accommodate advanced facilities and people interested in research at a nanoscale and ready to share and mix their ideas for innovations in the field of nanotechnology for the future advances of a global society.

133

Appendix C: Executive Summary of Newly Created Intellectual Property Understanding the details of interaction between different nanostructures is crucial for further research and development of new technologies based on

semiconductor nanostructures. The findings of this dissertation research that are listed below may be considered and used in design of quantum dot based light emitting devices or any other type of devices that require coupling between nanostructures. 1. Unique energy level system of coupled QDs and QW with resonantly aligned QW exciton state and third QD excited state was used to study quantum tunneling for the first time. 2. For the first time non-trivial behavior of the QD PLE signal as a function of barrier thickness at resonant QW excitation was observed and explained based on coherent effects. This result shows that coherence should be considered in tunnelcoupled nanostructures with barrier thicknesses less than 5 nm. 3. The method of tunneling time determination based on the ratio of cw-PL intensities from coupled QW and QDs and rate equation model, which includes state filling effects, was proposed and proven to be very accurate compared to the time-resolved PL measurements. 4. The strain field surrounding self-assembled QDs plays important role in localization of carriers which may improve the tunneling probability. Also, this strain can create additional confinement and even facilitate strain-induced QDs in a QW layer near the stressor QDs.

134

Appendix D: Potential Patent and Commercialization Aspects of listed Intellectual Property Items D.l Patentability of Intellectual Property Some of the four listed items could be potentially patented with the following reasoning: 1. The energy scheme itself used in this research cannot be patented. However, the principle of resonant coupling to a QD (or any nanostructure) excited state if used in any device design could be patented. 2. The QD PLE signal as a function of barrier thickness is a fundamental scientific result which so far does not have a direct device application and, therefore, could not be a subject of patent. 3. The method of extracting the tunneling times from the cw PL measurements of the QD-QW hybrid structure could be patented. 4. The strain-induced QDs are already known among scientific community and cannot be patented alone. However, if the strain-induced QDs are used as an active part of some device, then it could be included in a patent as a part of device design.

D.2 Commercialization Prospects The items listed as the newly created intellectual property of this dissertation research could be used in a process of development of new technologies for light emitting diodes, lasers, displays, photodetectors, and other future devices that may explore quantum properties (coherent tunneling) of hybrid nanostructures. Thus, the findings of

135

this work can potentially be commercialized in a future after additional research and device development is done which should include patenting of each particular device design. 1. The energy scheme of QD-QW structure should not be patented. 2. The QD PLE signal of QD-QW structure as a function of barrier thickness should not be patented. 3. The method of extracting the tunneling times from cw PL data should not be patented. 4. The strain-induced QDs in QD-QW structure should not be patented in current form.

D.3 Possible Prior Disclosure of IP All of the items listed were discussed in public forums (conferences and meetings) or have been published in peer-reviewed journals which places under question the patentability of the results obtained in this dissertation research in its current form. 1. The energy scheme of QD-QW structure was published in Journal of Applied Physics 108, 074305 (2010). 2. The QD PLE signal of QD-QW structure as a function of barrier thickness was published in Physical Review B 82, 155413 (2010). 3. The method of extracting the tunneling times from cw PL data was published in Journal of Applied Physics 108, 074305 (2010). 4. The strain-induced QDs in QD-QW structure was submitted for publication in Journal of Applied Physics (2011).

136

Appendix E: Broader Impact of Research

E.l Applicability of Research Methods to Other Problems The optical methods used to study coupling between semiconductor

nanostructures in this research can be applied to other materials systems such as colloidal QDs, metal nanoparticles, biological objects, hybrid systems that contain various materials interacting with each other. The problem of ultrafast optically induced charge carrier injection needs to be overcome for successful creation and development of faster optoelectronic devices, for example high-speed photodiodes, quantum cascade lasers, tunnel-injection QD lasers, optical switches, photodetectors, and photovoltaic devices. The method for tunneling time determination can be applied in any kind of experiments that involve carrier tunneling and light emission including biological objects and their combination with the semiconductor nanoparticles.

E.2 Impact of Research Results on U.S. and Global Society The research results of this dissertation directly deal with the problems of nanoscience and nanotechnology which have become one of the major players in hightech development trends of the world leading countries. Billion dollar investments are being made in research and development of new artificial materials and structures that can solve the energy problems of the global society and support a booming market of high-tech gismos like cell-phones, laptop computers, and tablet computers along with all possible electronic applications in every aspect of peoples lives. Understanding coherent phenomena will advance the progress towards optical quantum computation which

137

should open a completely new era allowing for fast processing of massive amounts of information and maybe even artificial intelligence.

E.3 Impact of Research Results on the Environment All the samples used in this research were grown in MBE chamber under ultrahigh vacuum conditions with no chemical pre- or after-treatment and no any toxic waste. The equipment and materials used in all optical experiments were also environmentfriendly.

138

>
Task Name START Graduate School Events Doctoral Dissertation Committee Doctoral Dissertation Title Research Proposal writing (Part 1) Research Proposal presentation Candidacy examination (Part 2) Defining Research Topic Discussion with Dr Yu Mazur Meet with Dr Salamo Independent Research to Undestend Existing Body of Knowledge Sorting available papers Search for new papers Near-field spectroscopy of single quantum dots at room temperature A close look on single quantum dots Low temperature near-field photoluminescence spectroscopy of InGaAs single q Scanning cathodoluminescence microscopy A unique approach to atomic scale < Single Quantum Dots Fundamentals, Applications and New Concepts Principles of Nano-Optics Photoluminescence study of lnSb/Alxlnx-1 Sb quantum well Excrton determination of strain parameters in InSb/AllnSb quantum wells Semiconductor optics Interdot carrier transfer in asymetric bilayer InAs/GaAs quantum dot structure Excitonic transfer in coupled InGaAsJGaAs quantum well to InAs quantum dot Materials and Supplies Samples QD+QW QD + QWB0,10,20A " Surface QD + QW QD-QW samples wrth different composition in QW QD + QW for electric field Data Acquisition PL General and lntensrty_Depenclence QD-QW sample testing QD*QWFA320,FA321 QD+QWFA326.FA327 QD+QWFA326.FA327 QDQWFA330,FA333 QD+OWFA334.FA335 QDtQW FA320, FA326, FA327 QDfQWFA23?, FA336, FA337_, FA333 QD*QWFA337,FA341 QD+QWFA342.FA343 QD+QW FA343 surface QD+QWFA344 QD*QW FA347, 348,349, 350,351 Ref QDs FA353 ^-Dr. Eucrydes Marega | Dr. Eucrydes Mnrega Ml Q3 * I Q4 09:01 2007 I Q1 I Q2 I Q3 I Q4 2008 I Q1 I Q2 I Q3 I Q4 2009 I Q1 I Q2 I Q3 I Q4 2010 I Q1 I Q2 I Q3 I Q4 2011 I 01 I 0 2 I Q3 I Q4 2012 I Q1

n a a

n
O
+ 0M8

as O

o
>i

*d

a
o M crei
| Dr. Vasyl Kimets.Dr. Euclydes Marega

ID 47 48 49 50 51 52 53 54 56 57 58 59 60 61 62 63 65 66

Task Name Q3 QD+QWFA355FA356 PL Excitation (PLE) Sample SC596 Sample SC596 second time PLE of all QD+QW samples PLE intensrty and temperature dependences PLE QD O r t samples with shifted QW PLE strain induced QDs Temperature Dependence CWPL QD*QWFA353 QD+QW FA327 QD+QW FA334 Temperature Dependence of all samples Temperature Dependence of srain induced QDs Time Resolved PL Temperature dep of TR PL PL + Electric Field TEM Sample preparation for TEM TEM measurements Data Analysis I Q4

2007 01

1 Q2

1 Q3

1 Q4

2008 Q1 I Q2

I Q3

I Q4

2009 Q1 I Q2

2010 I

Q3 I Q4

Q1 I Q2 I Q3 I Q4

2011 Q1 I Q2

I Q3

I Q4

2012 Q1

1 h

rflH

""
1 1 1

-i i
- M B

68 69 70 71 72 73 74 75 77 78 79 80 81 82

Processing Data Broadband s d 10(100)2nm s d 13(100)4nm Small Peak s d 10(100)2nm, s d 13(100)4nm Reference Sample for s d 10(100)2nm and s d 13(100)4nm Capped and uncapped QDs Samples 296 297 Temperature dependences Temperature dependence of strain induced QDs analysis Intensity dependences Reference Sample for s d 10(100)2nm and s d 13(100)4nm Broadband s d 10(100)2nm s d 13(100)4nm 11ML_287 11ML_288 QWell_291 QDs_296 QWell_297 PLE sample SC596 Last intensity measurement PLE

V
r
.
i
H'JO

84 85 86 87 88 89 90 91 92

SC596 SC596 second time PLE analysis PLE intensity and temperature analysis PLE QD QWwrth shifted QW analysis PLE strain induced QDs analysis PL+Electric Field analysis Lifetimes Lifetime vs Temperature

t
i

'

tessi
'lilMMBM

ID S3 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128

Task Name Q3 Trainings AFM Training PLE training Time Resolved PL training MBE growth training MBE growth training Dissertation Creation, Paperwork and Defense Read UARK Thesis Guidelines Introduction Lrterature Review Experimental part Results and Discussion 1 Results and Discussion 2 Conclusion Appendixes Turn in complete thesis for review Review by major professor Correct thesis Send abstract, t w o meeting dateAmeflocation to Renee Final review by Major Professor Signoff by Major Professor as complete ready for submission Review by Program Director Signoff by Program Director as complete and ready for commrttee Committee review Public Presentation Defense wrth committee Get committee signatures after defense Correct thesis Final approval by Major Professor Final approval by Program Director Print thesis copies, deliver all paperwork All paperwork to grad school/registrar Conferences Attended VC-NST Fall 2007, Fayetteville, AR MRS Fall Meeting 2008 Boston, MA APS March Meeting 2011 Dallas, TX Seminars I Q4

I2007 I Q1

I Q2

I Q3

I Q4

12008 I 31 I Q2

I Q3

I Q4

2009 Q1 I Q2

I Q3

I Q4

2010 Q1 I Q2

I Q3

I Q4

2011 Q1 I Q2

I Q3

I Q4

2012 Q1

| Baolai Liang | Baolai Liang g Baolai Liang

1,,,,

^ ^

Dr. Eudydes Marega gSS Dr.ZlnmingWang

fcDr.

Salnmo

,+ mm
(-Or. Salanio fli>r. Salanio H<en Vichers r-|4Ken Vichers

fftr. Satanic uKen Vichers

1
i
I

w
Lydie Louis Baolai Liang

130 131 132 133 134 135

All microEP Presentations Public Presentation Dissertation Defense

1 1 1

Grad Student Professional Learning Series Physics Colloquium Nobel laureate lecture Publication and Presentation of Results

!
i
w

137

Dr Salamo's Groop meeting

ID 138 139 140 141 142 143

Task Name Q3 Paper t o JAP Write a paper Submit to Journal Answer to Referee's questions Get published THE END Q4

2007 Q1

Q2

Q3

Q4

2008 Q1 I Q2

Q3

Q4

2009 Q1

Q2

Q3

Q4

2010 Q1

Q2

Q3

Q4

2011 Q1

2012

Q2

I Q3

I Q4

Q1

"S 1

*<>8<01

to

Appendix G: All Publications Published, Submitted and Planned 1. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr, M. Benamara, Z. Ya. Zhuchenko, G.G. Tarasov, C. Lienau, G.J. Salamo, State filling dependent tunneling in hybrid dot-well structures, submitted to Nano Research (2011). 2. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr, M. Benamara, G.G. Tarasov, G.J. Salamo, Spectroscopic signature of strain-induced quantum dots created by buried InAs quantum dots in InGaAs QW, submitted to Journal of Applied Physics (2011). 3. J. Lee, Zh.M. Wang, Y. Hirono, V.G. Dorogan, Yu.I. Mazur, E. Kim, S. Koo, S. Park, S. Song, G.J. Salamo, Low-density quantum dot molecules by selective etching using In droplet as a mask, IEEE Transactions on Nanotechnology 10, 600 (2011). 4. D.F. Cesar, M.D. Teodoro, V. Lopez-Richard, G.E. Marques, E. Marega Jr., V.G. Dorogan, Yu.I. Mazur, G.J. Salamo, Carrier transfer in the optical recombination of quantum dots, Physical Review B 83, 195307 (2011). 5. J. Wu, Zh.M. Wang, V.G. Dorogan, Yu.I. Mazur, S. Li, G.J. Salamo, Insight into optical properties of strain-free quantum dot pairs, Journal of Nanoparticle Research 13, 947 (2011). 6. J. Wu, Zh.M. Wang, V.G. Dorogan, S. Li, Yu.I. Mazur, G.J. Salamo, Near infrared broadband emission of Ino 35Gao 6sAs quantum dots on high index GaAs surfaces, NanoscaleX 1485(2011). 7. V.G. Dorogan, Zh.M. Wang, Vas.P. Kunets, M. Schmidbauer, Y.Z. Xie, M.D. Teodoro, P.M. Lytvyn, Yu.I. Mazur, G.J. Salamo, Alignment and optical polarization of InGaAs quantum wires on GaAs high index surfaces, Materials Letters 65, 1427 (2011). 8. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr, M. Benamara, Z.Ya. Zhuchenko, G.G. Tarasov, C. Lienau, G.J. Salamo, Excited state coherent resonant electronic tunneling in quantum well - quantum dot hybrid structure, Applied Physics Letters 98,083118(2011). 9. J. Lee, Zh. Wang, Y. Hirono, E.-S. Kim, S.-M. Koo, V.G. Dorogan, Yu.I. Mazur, S. Song, G. Park, G.J. Salamo, InGaAs quantum dot molecules during selective etching using an In droplet mask, Journal of Physics D: Applied Physics 44, 025102 (2011). 10. Vas.P. Kunets, M.D. Teodoro, V.G. Dorogan, P.M. Lytvyn, G.G. Tarasov, R. Sleezer, M.E. Ware, Yu.I. Mazur, J.S. Krasinski, G.J. Salamo, Interface roughness scattering in laterally coupled InGaAs quantum wires, Applied Physics Letters 97, 262103 (2010). ll.ZhJVl. Wang, Vas.P. Kunets, Y.Z. Xie, M. Schmidbauer, V.G. Dorogan, Yu.I. Mazur, G.J. Salamo, Multilayer self-organization of InGaAs quantum wires on GaAs surfaces, Physics Letters A 375, 170 (2010). 143

12. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr, Z.Ya. Zhuchenko, M.E.Ware, M. Benamara, G.G. Tarasov, P. Vasa, C. Lienau, G.J. Salamo, Tunneling-barrier controlled excitation transfer in hybrid quantum dot - quantum well nanostructures, Journal ofApplied Physics 108, 074316 (2010). 13. K.A. Sablon, J.W. Little, K.A. Olver, Zh.M. Wang, V.G. Dorogan, Yu.I. Mazur, G.J. Salamo, F.J. Towner, Effects of AlGaAs energy barriers on InAs/GaAs quantum dot solar cells, Journal ofApplied Physics 108, 074305 (2010). 14. Yu.I. Mazur, V.G. Dorogan, D. Guzun, E. Marega Jr., G.J. Salamo, G.G. Tarasov, A.O. Govorov, P. Vasa, C. Lienau, Measurement of coherent tunneling between InGaAs quantum wells and InAs quantum dots using photoluminescence spectroscopy, Physical Review 2? 82, 155413 (2010). 15. J. He, C.J. Reyner, B.L.Liang, K. Nunna, D.L. Huffaker, N. Pavarelli, K. Gradkowski, T.J. Ochalski, G. Huyet, V.G. Dorogan, Yu.I. Mazur, G.J. Salamo, Band alignment tailoring of InAsi.xSbx/GaAs quantum dots: Control of type I to type II transition, Nano Letters 10, 3052 (2010). 16. Zh.M. Wang, Y.Z. Xie, Vas.P. Kunets, V.G. Dorogan, Yu.I. Mazur, G.J. Salamo, Multilayers of InGaAs nanostructures grown on GaAs (210) substrates, Nanoscale Research Letters 5, 1320 (2010). 17. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr., D.F.Cesar, V.Lopez-Richard, G.E. Marques, Z.Ya. Zhuchenko, G.G. Tarasov, G.J. Salamo, Cooperative effects in the photoluminescence of (In,Ga)As/GaAs quantum dot chains structures, Nanoscale Research Letters 5, 991 (2010). 18. J. Wu, D. Shao, V.G. Dorogan, A.Z.Li, S.Li, E.A. DeCuir Jr., M.O. Manasreh, Z.M. Wang, Yu.I. Mazur, G.J. Salamo, Intersublevel infrared photodetector with strain-free GaAs quantum dot pairs grown by high-temperature droplet epitaxy, Nano Letters 10, 1512 (2010). 19. J. Lee, Zh.M. Wang, V.G. Dorogan, Yu.I. Mazur, G.J. Salamo, Evolution of various nanostructures and preservation of self-assembled InAs quantum dots during GaAs capping, IEEE Transactions on Nanotechnology 9, 149 (2010). 20. M.D. Teodoro, V.L. Campo Jr., V. Lopez-Richard, E. Marega Jr., G.E. Marques, Y. Galvao Gobato, F. Iikawa, M.J.S.P. Brasil, Z.Y. AbuWaar, V.G. Dorogan, Yu.I. Mazur, M. Benamara, G.J. Salamo, Aharonov-Bohm interference in neutral excitons: effects of built-in electric fields, Physical Review Letters 104, 086401 (2010). 21.Y.Z. Xie, V.P. Kunets, Z.M.Wang, V. Dorogan, Y.I. Mazur, J. Wu and G.J. Salamo, Multiple stacking of InGaAs/GaAs(731) nanostructures, Nano-Micro Letters 1, 1 (2009). 22. J.H. Lee, Zh.M. Wang, V.G. Dorogan, Yu.I. Mazur, M.E. Ware, G.J. Salamo, Tuning the emission profiles of various self-assembled InGaAs nanostructures by rapid thermal annealing, Journal of Applied Physics 106, 073106 (2009). 23. V.G. Dorogan, Yu.I. Mazur, E. Marega Jr., G.G. Tarasov, M.E. Ware, G.J. Salamo, Hybridized quantum dot-wetting layer states in photoluminescence of In(Ga)As/GaAs dot chain samples, Journal of Applied Physics 105, 124304 (2009). 144

24. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr., P.M. Lytvyn, Z.Ya. Zhuchenko, G.G. Tarasov, G.J. Salamo, One-dimensional features of In(Ga)As/GaAs dot chain structures with changeable interdot coupling, New Journal of Physics 11, 043022 (2009). 25. Yu.I. Mazur, V.G. Dorogan, E. Marega Jr., G.G. Tarasov, D.F.Cesar, V.LopezRichard, G.E. Marques, G.J. Salamo, Mechanisms of interdot coupling in (In,Ga)As/GaAs quatum dot arrays, Applied Physics Letters 94, 123112 (2009). 26. Yu.I. Mazur, V.G. Dorogan, O. Bierwagen, G.G. Tarasov, E.A. DeCuir Jr., S. Noda, Z.Ya. Zhuchenko, M.O. Manasreh, W.T. Masselink, G.J. Salamo, Spectroscopy of shallow InAs/InP quantum wire nanostructures, Nanotechnology 20, 065401 (2009). 27. B.L. Liang, V.G. Dorogan, Yu.I. Mazur, N.W. Strom, J.H. Lee, K.A. Sablon, Zh.M. Wang, G.J. Salamo, InAs quantum dot clusters grown on GaAs droplet templates: surface morphologies and optical properties, Journal of Nanoscience and Nanotechnology 9, 3320 (2009). 28. Vas.P. Kunets, T.Al. Morgan, Yu.I. Mazur, V.G. Dorogan, P.M. Lytvyn, M.E. Ware, D. Guzun, J.L. Shultz, G.J. Salamo, Deep traps in GaAs/InGaAs quantum wells and quantum dots studied by noise spectroscopy, Journal of Applied Physics 104, 103709 (2008). 29. P.M. Lytvyn, Yu.I. Mazur, E. Marega Jr., V.G. Dorogan, V.P. Kladko, M.V. Slobodian, V.V. Strelchuk, M.L. Hussein, M.E. Ware, G.J. Salamo, Engineering of 3D self-directed quantum dot ordering in multilayer InGaAs/GaAs nanostructures by means of flux gas composition, Nanotechnology 19, 505605 (2008). 30. V.G. Dorogan, Yu.I. Mazur, J.H. Lee, Zh.M. Wang, M.E. Ware, G.J. Salamo, Thermal peculiarity of AlAs-capped InAs quantum dots in a GaAs matrix, Journal of Applied Physics 104, 104303 (2008). 31. Yu.I. Mazur, Z.Y. AbuWaar, T.D. Mishima, J.H.Lee, G.G. Tarasov, B.L.Liang, V.G. Dorogan, M.E. Ware, Zh.M. Wang, M.B. Johnson, G.J. Salamo, Spectroscopic observation of developing InAs quantum dots on GaAs ringlike-nanostructured templates, Journal ofApplied Physics 104, 044310 (2008). 32. P.S. Wong, B.L. Liang, V.G. Dorogan, A.R. Albrecht, J. Tatebayashi, X. He, N. Nuntawong, Yu.I. Mazur, G.J. Salamo, S.R.J. Brueck D.L. Huffaker, Improved photoluminescence efficiency of patterned quantum dots incorporating a dots-in-thewell structure, Nanotechnology 19, 435710 (2008). 33. Yu.I. Mazur, S. Noda, G.G. Tarasov, V.G. Dorogan, G.J. Salamo, O. Bierwagen, W.T. Masselink, E.A. DeCuir Jr., M.O. Manasreh, Excitonic band edges and optical anisotrogy of InAs/InP quantum dot structures, Journal of Applied Physics 103, 054315(2008).

145

Das könnte Ihnen auch gefallen