05.12.2012 Views

Diplomarbeit Diplom-Ingenieur - Institut für Halbleiter

Diplomarbeit Diplom-Ingenieur - Institut für Halbleiter

Diplomarbeit Diplom-Ingenieur - Institut für Halbleiter

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Transmission Electron Microscopy of self-organised<br />

PbTe/CdTe Nanocrystals<br />

<strong><strong>Diplom</strong>arbeit</strong><br />

zur Erlangung des akademischen Grades<br />

<strong>Diplom</strong>-<strong>Ingenieur</strong><br />

im <strong>Diplom</strong>studium<br />

Technische Physik<br />

Angefertigt am <strong>Institut</strong> <strong>für</strong> <strong>Halbleiter</strong>physik<br />

Eingereicht von:<br />

Heiko Groiss<br />

Betreuung:<br />

Univ. Prof. Dr. Friedrich Schäffler<br />

Linz, Oktober 2006<br />

1


Eidesstattliche Erklärung<br />

Ich erkläre an Eides statt, dass ich die vorliegende <strong><strong>Diplom</strong>arbeit</strong> selbständig und ohne fremde<br />

Hilfe verfasst, andere als die angegebenen Quellen und Hilfsmittel nicht benutzt bzw. die<br />

wörtlich oder sinngemäß entnommenen Stellen als solche kenntlich gemacht habe.<br />

i<br />

Heiko Groiss


Kurzfassung<br />

Nanostrukturierung erlaubt die Beeinflussung der elektronischen und optischen Eigenschaften<br />

von <strong>Halbleiter</strong>strukturen <strong>für</strong> vielfältige Anwendungen. Sogenannte Quantenpunkte (QD)<br />

zeigen ein mit Atomen vergleichbares optisches Verhalten und sind eine vielversprechende<br />

Methode zur Herstellung von Materialsystemen <strong>für</strong> optoelektronische Bauelemente <strong>für</strong><br />

Frequenzbereiche, die mit den Standardmaterialien nicht zugänglich sind.<br />

Ein neuartiger Ansatz der QD Produktion durch Entmischung in der Festkörperphase<br />

zweier <strong>Halbleiter</strong> mit unterschiedlichen Gitterstrukturen wurde mit CdTe/PbTe<br />

Heterostrukturen verwirklicht. PbTe besitzt die Struktur von Steinsalz (rs), CdTe hat<br />

Zinkblende-(zb)-Struktur. Eine in CdTe eingebettete PbTe Schicht zerfällt durch Tempern in<br />

hochgradig symmetrische Nanokristalle mit atomistisch scharfen Grenzflächen zu dem<br />

Umgebungsmaterial. Die entstandenen QDs zeigen starke Photolumineszenz im mittleren<br />

Infrarot, einem Energiebereich, in dem viele Molekülanregungsenergien zu finden sind, <strong>für</strong><br />

den aber keine effizienten <strong>Halbleiter</strong>laser existieren. Hauptaufgabe dieser Arbeit war die<br />

Charakterisierung dieser QDs mit dem Transmissionselektronenmikroskop (TEM), wobei<br />

verschieden Mikroskopiertechniken verwendet wurden.<br />

Wie gezeigt wird, besitzen die PbTe Nanokristalle im thermodynamische Gleichgewicht<br />

die geometrische Form von kleinen Rhomboedrischen–Kubo-Oktaeder mit {001}, {110} und<br />

{111} PbTe/CdTe Grenzflächen, wobei auch Einblicke in die Prozesse während des<br />

Temperschrittes gegeben werden. Die Größe der QDs kann mittels Epischichtdicke gesteuert<br />

werden, wobei die Dichte der Quantenpunkte stark mit ihrer mittleren Größe zusammenhängt.<br />

Der Hauptteil dieser <strong><strong>Diplom</strong>arbeit</strong> beschäftigt sich dann mit der atomaren<br />

Charakterisierung der PbTe/CdTe Grenzflächen und den, <strong>für</strong> die Interpretation notwendigen<br />

TEM Simulationen. Durch die unterschiedliche Gitterstruktur der beiden beteiligten<br />

Materialien kommt es an den Grenzflächen zu starken Verschiebungen einzelner Atome oder<br />

ganzer Atomgruppen, um die Bindungkonfigurationen <strong>für</strong> das zb oder rs Gitter besser zu<br />

erfüllen. Abhängig von der Terminierung der CdTe Kristallhälfte konnte die Existenz von<br />

zwei unterschiedlichen {001} Grenzflächen am selben QD nachgewiesen werden. Die<br />

bemerkenswertesten Effekte treten aber an den {110} Grenzflächen auf, wo die<br />

Rekonstruktion zu einem Versatz der beiden Kristallhälften von mehr als 10% der<br />

Gitterkonstante führt. Das Grenzflächenproblem wurde auch von R. Leitsman et al. mit ab-<br />

initio-Rechnungen behandelt. Die theoretischen Verschiebungen und die aus hochauflösenden<br />

TEM-Aufnahmen gewonnenen zeigen eine hervorragende Übereinstimmung.<br />

iii


Abstract<br />

Nano-structuring allows the manipulation of electronic and optical properties of<br />

semiconductors, which is useful for various applications. So-called quantumdots (QD) reveal<br />

optical properties comparable with atoms. They are a promising method for the production of<br />

material systems for opto-electronic devices in frequency ranges, which are not accessible<br />

with the standard materials.<br />

A novel approach of QD formation was realized by decomposition of two immiscible<br />

semiconductors with different lattice structures. The experiments were carried out at<br />

CdTe/PbTe heterostructures. PbTe possesses the structure of rock salt (rs), CdTe has<br />

zincblende (zb) structure. A PbTe layer, embedded in CdTe, disintegrates into highly<br />

symmetrical nano-crystals with atomically sharp interfaces during an annealing step. The QDs<br />

show intense photoluminescence in the mid-infrared, an energy region, where the excitation<br />

energies of molecules are found. No efficient semiconductor lasers exist for this frequency<br />

range. The aim of this work was a concise characterisation of these QDs with transmission<br />

electron microscopy (TEM), for which different techniques were used.<br />

As will be shown, the PbTe nanocrystals have a thermodynamic equilibrium shape of<br />

small rhombi-cubo-octahedrons with {001}, {110} and {111} PbTe/CdTe interfaces. An<br />

insight into the processes during the annealing step is given. The size of the QDs can be<br />

controlled by the epi-layer thickness. The density of the dots depends strongly on their<br />

average size.<br />

The main part of this thesis is the atomic characterisation of the PbTe/CdTe interfaces and<br />

TEM image simulations, which are necessary for the interpretation. The different lattice<br />

structures of the two materials induce strong atomic displacement of single atoms or groups of<br />

atoms at the interfaces. The displacements end up in bonding configurations for the zb or rs<br />

lattice close to their respective bulk configuration. It was shown that two different {001}<br />

interfaces exist at the same QD, depending on the termination of the CdTe crystal half. The<br />

most remarkable effects can be found at the {110} interfaces, where the displacements lead to<br />

a shift of more than 10% of the two crystal halves against each other. The interface problem<br />

was also treated by R. Leitsman et al. with ab-initio calculations. The theoretical<br />

displacements and those obtained by high resolution TEM images show excellent agreement.<br />

v


Contents<br />

Preface i<br />

Eidesstattliche Erklärung i<br />

Kurzfassung iii<br />

Abstract v<br />

Contents vii<br />

1. Introduction 1<br />

1.1. Materials for mid-infrared devices and applications 1<br />

1.2. Quantum Dot precipitation in solids 3<br />

2. The material system and its fabrication 5<br />

2.1. PbTe and CdTe –Two immiscible Materials 5<br />

2.2. Sample growth with Molecular Beam Epitaxy 7<br />

2.2.1. Growth of the CdTe/PbTe Heterostructure 9<br />

2.3. Annealing 11<br />

3. The Transmission Electron Microscope 13<br />

3.1. The JEOL JEM-2011 FastTem 13<br />

3.2. Scattering of Electrons 17<br />

3.3. Contrast Enhancement with Bright and Dark Field Imaging 21<br />

3.3.1. Contrast Mechanism 21<br />

3.3.2. Dark Field and Bright Field imaging 22<br />

3.4. High Resolution Transmission Electron Microscopy 25<br />

3.4.1. Requirements for HRTEM 25<br />

3.4.2. Origin of Lattice Fringes 26<br />

3.4.3. Transferfunction and Scherzer Defocus 27<br />

3.5. Sample Preparation 30<br />

vii


viii<br />

4. Shape and size of the PbTe Dots 31<br />

4.1. Available Samples 31<br />

4.2. Precipitation steps 33<br />

4.3. Precipitation from different layer thicknesses 36<br />

4.3.1. Annealed 5nm layer 36<br />

4.3.2. Annealed 1nm layer 38<br />

4.4. Equilibrium shape 40<br />

5. Interpretation of HRTEM Images 44<br />

5.1. JEMS Simulation Package 44<br />

5.1.1. Multi-Slice Approach 46<br />

5.1.2. Simulation Procedure 47<br />

5.2. Image simulation of the CdTe/PbTe interfaces 50<br />

6. Interface characterization 52<br />

6.1. Available Samples 52<br />

6.1.1. Specimens for HRTEM imaging 52<br />

6.1.2. Beam Damage 53<br />

6.2. HRTEM investigation of PbTe nanocrystals 54<br />

6.2.1. HRTEM images of single QDs 54<br />

6.2.2. Total energy calculations 56<br />

6.2.3. The (001) interface 58<br />

6.2.4. The (110) interface 65<br />

6.2.5. The (111) interface 73<br />

7. Conclusions and Outlook 77<br />

Appendix A Structure factors of CdTe and PbTe 82<br />

Appendix B Simulated HR-maps of CdTe/PbTe interfaces 84<br />

References 91<br />

Curriculum vitae 95<br />

List of Publications 96<br />

Acknowledgment 98


Chapter 1<br />

Introduction<br />

1.1 Materials for mid-infrared devices and applications<br />

Optical devices have a broad field of applications. The physical principle behind<br />

most applications is the recombination of an electron-hole pair over a direct band gap<br />

in semiconductor materials. Band gap design is a key to produce materials with fitting<br />

band gaps. Further tuning of the optical properties is done by micro- and nano-<br />

structuring of the optical active materials. The approach of low dimensional systems<br />

like quantum dots (QDs) or quantum wires leads to confined electrons. Electrons<br />

confined in a certain potential show quantized properties and behave like the electrons<br />

in atoms or molecules with certain excitation levels [1, 2]. The confinement yields to<br />

a blueshift and is controlled by the size of the structures used for confinement.<br />

Optoelectronic device based on QDs are realized for different application, e.g QDs<br />

lasers [3, 4] or single photon sources [5, 6]. But the band gaps of typical direct band<br />

gap semiconductors are not perfectly suitable for the mid-infrared frequency range (3-<br />

30μm). Lead salts are used for applications in this frequency range and further nano-<br />

structuring would improve the performance of these devices. The mid infrared<br />

frequency range can be found e.g. at the absorption energies of CO2 or other<br />

molecules important for environmental monitoring or medical diagnosis. Improved<br />

semiconductor-lasers and detectors would allow efficient application in these fields.<br />

QDs can be produced by using self-organisation schemes during epitaxial growth<br />

of strained semiconductor heterostructures. The best-known self-organisation scheme<br />

is the Stranski-Krastanov (SK) growth mode [7-10]. This growth mode is governed by<br />

surface energy minimisation and strain relaxation. SK growth is always starting with<br />

the formation of a wetting layer to minimise the surface, which covers the entire<br />

substrate. If the stored strain energy gets larger than the energy benefit of layer-by-<br />

layer growth, the formation of QDs starts. A capping layer is needed for an<br />

implementation of these dots, so that the dots are surrounded by a host material.<br />

During the overgrowth, alloying and shape transitions appear [11-13]. An often<br />

1


2<br />

encountered shape of SK dots is that of a pyramid, which during capping usually<br />

changes to the shape of a truncated pyramid. Capping of the dots causes strong<br />

alloying of the host material with the dot material, which is a disadvantage of SK dots.<br />

A further disadvantage is the existing 2D wetting layer. Examples for Stranski-<br />

Krastanov systems are InGaAs dots on GaAs or Ga dots on Si (see Fig. 1.1a).<br />

Another approach is the synthesis of core-shell nanoparticles [14, 15]. During a<br />

chemical reaction of liquid precursors the QDs precipitate out of the liquid. FeO<br />

quantum dots produced by this method can be seen in Fig.1.1b. These dots are highly<br />

symmetric with sharp surfaces and the size distribution can be controlled. With this<br />

technique it is possible to produce dots in the nanometer scale. A disadvantage of this<br />

approach is an organic shell surrounding the dots due to the chemicals that are used to<br />

stabilize the reaction. This is an essential part of the synthesis, otherwise the<br />

nanocrystals would agglomerate. The organic shell is not conductive and thus<br />

hampering electronic applications. The dots can be sorted by size and produced as<br />

powder, but the nanocrystals are not surrounded by a host material and thus difficult<br />

to apply in optoelectronic devices.<br />

Figure 1.1: (a) TEM image of stacked Ge Stranski-Krastanov dots on Si. The dots have<br />

the shapes of truncated pyramids. The wetting layer can be seen on the left and right<br />

handed side of the dots. (b) TEM image of FeO dots fabricated by core-shell synthesis.


1.2 Quantum Dot precipitation in solids<br />

The aim of QD fabrication by precipitation in solids is the combination of the<br />

advantages of the Stranski-Krastanov grown dots and the chemical synthesised ones.<br />

The nanocrystals should be embedded with high shape symmetry and sharp interfaces<br />

in a host material. There should be no alloying between the host and the dot materials.<br />

The size of the dots should be well controllable.<br />

Our material system consists of a CdTe/PbTe heterostructure. Earlier studies of this<br />

material system attempted to produce a PbxCd1-xTe ternary alloy by the Bridgman<br />

method [16-18]. The characterisation of their samples takes place in the μm range<br />

where Cd and Pb rich grains are found. No imaging techniques were used to reveal<br />

structures with the dimensions of nanocrystals, which are treated within this thesis.<br />

Our heterostructures consist of a thin PbTe layer in the nanometer range, clad by<br />

CdTe [16]. These two materials have almost the same lattice constant and<br />

Figure 1.2: PL measurements at room temperature of CdTe/PbTe heterostructures<br />

annealed at different temperatures. Different solid curves in (a) represent different<br />

annealing temperatures. The intensity increases for higher annealing temperatures and<br />

the maximum is shifted to longer wavelength of 3.2 μm for QDs. The sample was excited<br />

below the CdTe barrier with a 1480nm pump laser with 245mW. The dotted lines show<br />

the PL spectrum of bulk PbTe. (b) Shows the luminescent of a PbTe/CdTe<br />

heterostructure which was excited above the CdTe barrier with a 730nm Ti:sapphire laser<br />

with 400mW. [16]<br />

3


4<br />

the heterostructures can be grown almost strain-free. PbTe and CdTe differ<br />

fundamentally in their lattice structure. PbTe has rock salt (rs) structure, whereas<br />

CdTe has zinc blend (zb) structure. The precipitation of the QDs in solids takes place<br />

during an annealing step and is driven by lattice type mismatch. The novel approach<br />

of precipitation in the solid phase should also be applicable to other semiconductor<br />

heterosystems and was shown for the semimetals ErAs and ErSb (rs lattice) on GaAs<br />

or GaSb substrate (zb lattice) respectively [20-23].<br />

Our PbTe/CdTe semiconductor system was at first investigated with photo<br />

luminescent (PL) measurements (see Fig.1.2) on CdTe/PbTe heterostructures. The<br />

measurements were performed at room temperature with excitation below and above<br />

the CdTe barrier for samples which were annealed at different temperatures. The PL<br />

signal of the heterostructures increases significantly after the annealing step. First<br />

TEM investigation showed that the 2-dimensional PbTe layer breaks up into highly<br />

symmetric QDs (see Fig. 1.3) with sharp interface between the dot and host materials<br />

during the annealing step. The resulting dots are higher than the original 2<br />

dimensional quantum well (QW), which leads to a shift of the increased luminescent<br />

peak to larger wave lengths. An in-depth characterisation by TEM of the shape, size<br />

and interfaces of these dots was the aim of this diploma thesis.<br />

Figure 1.3: (a) PbTe quantum dots that precipitated from a 1nm PbTe epi-layer in plan<br />

view. (b) Two dots that precipitated from a 5nm epi-layer in a cross section view. Their<br />

shape is highly symmetric and sharp interfaces can be seen between the PbTe and the<br />

CdTe host material.


Chapter 2<br />

The material system and its fabrication<br />

Lead telluride and cadmium telluride are well-know semiconductor materials. The<br />

narrow band gap material PbTe has rock salt (rs) structure and is used as material for<br />

infrared detectors or as thermoelectric material. CdTe is a wide band gap material<br />

with zinc blend (zb) structure and is used as material for thin film solar cells. Due to<br />

its properties in the infrared, it is often used for optical windows or lenses in this<br />

range. The combination of these two immiscible materials for QD precipitation is a<br />

novel approach for mid infrared materials.<br />

2.1 PbTe and CdTe –Two immiscible Materials<br />

CdTe consists of elements from the II´s and VI´s group of the periodic table. It has a<br />

cubic lattice structure of the zincblend type. A zb - lattice consists of two face centred<br />

cubic (fcc) lattices, shifted by one fourth of the unit cell diagonal. Each sublattice<br />

Figure 2.1: CdTe unit cell; each atom has four nearest neighbours which are from the<br />

other atomic species.<br />

5


6<br />

is occupied with one kind of atoms, in this case either with Cd or Te (see Fig. 2.1). In<br />

this lattice, each atom has four nearest neighbours. There are covalent and ionic<br />

contributions to the chemical bonding. The bonds of the atoms are four-fold<br />

coordinated sp 3 hybridised binding orbitals. Surfaces perpendicular to the 〈100〉 or<br />

〈111〉 crystal directions are polar surfaces and can be either Cd or Te terminated. The<br />

surfaces perpendicular to the 〈110〉 direction is a non-polar one.<br />

In contrast to CdTe, PbTe is an ionic crystal. The element Pb is from the IV´s<br />

group of the periodic table, thus PbTe is an IV-VI semiconductor. The ionic crystal<br />

has also a cubic lattice but with rs structure. This lattice consists of two fcc sublattices<br />

which are displaced by one half of the unit cell diagonal. This configuration leads to<br />

six nearest neighbours for each atom (see Fig. 2.2). The bonds in this crystal are p 3<br />

binding orbital with a strong ionic character. Interfaces of a PbTe crystal<br />

perpendicular to the 〈100〉, 〈110〉, 〈111〉 direction are non polar surfaces, where the<br />

first atomic layer contains the same number of each atom species.<br />

Figure 2.2: PbTe unit cell; each atom has six nearest neighbours which are from the<br />

other atomic species.<br />

CdTe and PbTe have a large miscibility gap [24]. The reason for this immiscibility<br />

is the fundamental difference in the lattice structure and not a difference in lattice<br />

constant as in the case of the miscibility gap of In1-xGaxN [25]. CdTe and PbTe have<br />

almost the same lattice constant, the difference being smaller than 0.3% at 300K. The<br />

lattice constant of CdTe is slightly larger but the thermal expansion coefficient is<br />

smaller. Both materials have the same lattice constant at about 425 K. Both materials<br />

are direct band gap materials which means that the minimum of the conduction band<br />

and the maximum of the valence band are at the same point in the bandstructure.<br />

CdTe is a wide band gap material with a direct band gap at the Γ point. The band gap


of PbTe is one order of magnitude smaller and its direct band gap is at the L point.<br />

This band gap difference of about 1.2eV at room temperature allows an efficient<br />

carrier confinement in the PbTe/CdTe material system. Important material properties<br />

of CdTe/PbTe are shown in Tab. 2.1.<br />

CdTe PbTe<br />

Structure Zincblend Rocksalt<br />

Lattice constant (300 K) [nm] 0.648 0.6462<br />

Thermal expansion coefficient [K -1 ] 4.70E-06 2.70E-05<br />

Direct band gap at 0 K eV Γ8-Γ6= 1.61 L6 - -L6 + = 0.187<br />

dEg/dT [[eV/K] -3.0E-04 4.1E-04<br />

Electron effective mass [m0] 0.11 ml= 0.24<br />

mt= 0.024<br />

Hole effective mass [m0] 0.41 ml= 0.31<br />

mt= 0.022<br />

Table 2.1: Comparison of important material constants of CdTe and PbTe [26]<br />

An important property of the materials for the transmission electron microscopy is<br />

the structure factor Fhkl. The structure factor depends on the lattice structure of the<br />

material and determines which diffraction spots are allowed. Both structures consist<br />

of fcc lattices. All diffraction spots with mixed even and odd h, k, l are forbidden for<br />

the fcc lattice thus also for PbTe and CdTe. The structure factors of the other<br />

diffraction spots for CdTe and PbTe can be found in the Appendix A.<br />

2.2 Sample growth with Molecular Beam Epitaxy<br />

All samples of the CdTe/PbTe heterostructures were grown by Molecular Beam<br />

Epitaxy (MBE) on (001)-GaAs substrates. The MBE technique is widely used in<br />

semiconductor research because it allows growing material layers with a controllable<br />

thickness down to an atomic monolayer of atoms. Epitaxy means that the substrate<br />

crystal structure determines the further growth. One disadvantage of the MBE<br />

technique is the low growth rate, therefore semiconductor industry mainly uses other<br />

techniques like Chemical Vapour Deposition or Metal Organic Vapour Deposition.<br />

The essential parts of a MBE system are the molecular effusion cells and substrate<br />

heating. These parts are placed in a vacuum chamber because ultra high vacuum<br />

7


8<br />

conditions are essential for the crystal growth. A common method for in situ<br />

characterisation is Reflection High Energy Electron Diffraction (RHEED).<br />

The basic processes of crystal growth take place in three zones between the<br />

effusion cell and the sample surface. Firstly, there is the molecular beam generation<br />

zone. The vapour elements mix in the mixing zone and crystal growth takes place in<br />

the crystallisation zone on the substrate surface. The growth itself is controlled by the<br />

substrate temperature, the beam flux and the process time. A schematic drawing of a<br />

MBE system with the tree process zones can be seen in Fig. 2.3.<br />

Figure 2.3: Sketch of a MBE System [27]. The essential parts of the system are shown,<br />

and the tree main process zones are indicated.<br />

In general, three epitaxial growth modes driven by energy minimisation can be<br />

distinguished. The energy terms are the surface, interface and strain energy. The<br />

Frank-van der Merwe growth is a monolayer-by-monolayer growth. The substrate and<br />

the epi-material have a fitting lattice constant, thus no strain is established during the<br />

crystal growth. The interface energy is smaller than the surface energy, consequently<br />

the epi-layer minimises its surface and grows in closed monolayer steps. The Volmer-<br />

Weber mode refers to a pure island growth on the surface. The strain energy term<br />

plays again no role. The surface energy is now the smaller term and the crystal grows<br />

with the minimum interface in the shape of islands. The Stranski-Krastanov growth is


an intermediate case with a strain energy term. In the beginning, the interface energy<br />

is smaller than the surface energy and the stored strain energy is small. It is<br />

energetically beneficial to build up a strained epi-layer. As soon as the stored strain<br />

energy gets lager than the difference between interface and surface energy, the growth<br />

characteristic switches, and island growth starts to relax strain energy.<br />

2.2.1 Growth of the CdTe/PbTe Heterostructure<br />

All samples investigated within this thesis were grown at the Osaka <strong>Institut</strong>e of<br />

Technology, Japan [26], [28], [29]. The growth was carried out in a MBE apparatus<br />

with solid sources of CdTe, PbTe, Cd and Pb in Knudsen cells.<br />

Figure 2.4: Sketch of the sample Structure; The HQ buffer is necessary to provide a<br />

good (001) CdTe surface, which is either Te or Cd terminated before the growth of PbTe.<br />

The PbTe layer is grown with different thicknesses and the heterostructure is finalised<br />

with a CdTe capping layer.<br />

The basic structure of the samples consists of a high quality (HQ) buffer of CdTe<br />

grown on (100)-oriented GaAs substrates. On top of this buffer the PbTe layers were<br />

grown with different thicknesses. This Single Quantum Well (SQW) was than capped<br />

again with CdTe. Some samples were also capped additionally with MnTe to prevent<br />

the CdTe from re-evaporation during annealing. A sketch of a heterostructure can be<br />

9


10<br />

seen in Fig. 2.4, where also the structure of the HQ buffer is shown. Fig. 2.5 shows a<br />

TEM image of an annealed sample, where also the buffer structure can be seen.<br />

Essential for the growth of a high quality PbTe SQW is a high-quality CdTe buffer<br />

with low surface roughness and low lattice defects. The lattice constants of GaAs and<br />

CdTe differ by ~15% which lead to CdTe growth with a high dislocation density. To<br />

grow a high-quality CdTe buffer, fist a 100 nm thick low temperature (LT) CdTe film<br />

was grown at 280°C on a sulphur treated GaAs substrate, followed by 300 nm CdTe<br />

at 320°C providing a (100) oriented CdTe surface. This was capped with a 100 nm<br />

thick MnTe layer to suppress the re-evaporation during the following thermal<br />

annealing step at 380°C for 30min. On the CdTe/MnTe buffer a superlattice of twenty<br />

5 nm thick MnTe / 2 nm thick CdTe sublayers were grown at 280°C and then<br />

finalised with a CdTe layer (300nm-900nm) at 320°C [28]. Such high-quality buffers,<br />

or similar ones, were used for all samples investigated within this thesis. The root-<br />

mean-square roughness of such a HQ buffer is less than 0.5nm in an area of<br />

3μm × 3μm (measured by atomic force microscopy).<br />

Figure 2.5: A typical PbTe/CdTe heterostructure recorded with the TEM along the [110]<br />

zone axis. The layers with Mn content are brighter and clearly visible. The HQ buffer<br />

was finalised with 900nm CdTe. The image shows an annealed sample. The PbTe dots<br />

can be seen beneath the surface of the sample at the top of the image.<br />

On such high quality CdTe buffers the PbTe SQWs were grown at temperatures<br />

from 220°C to 280°C with different thickness. The CdTe (100) surface is polar,


therefore it can be Te or Cd terminated. Roughness measurements with atomic force<br />

microscopy showed that Cd termination before PbTe growth causes a smoother<br />

surface [26]. The termination also determines the CdTe/PbTe interface structure as is<br />

shown later in this thesis.<br />

2.3 Annealing<br />

The transformation of the heterostructure into PbTe QDs embedded in a CdTe matrix<br />

happens during an annealing step. Annealing was performed after growth at different<br />

temperatures above 300°C for 10min. CdTe and PbTe have the same lattice constant<br />

at about 150°C. The strain in the PbTe SQW switches from -0.3% tensile strain (room<br />

temperature) to 0.4% compressive strain (temperatures above 300°C). During the<br />

annealing step the SQW breaks up into islands of PbTe with a larger thickness than<br />

the original 2-dim layer. The process is driven by the immiscibility of the two<br />

materials and activated by the thermal load transferred during annealing. The<br />

separated PbTe material tries to reach a thermal equilibrium shape caused by<br />

minimising the total energy. The PbTe islands end up in a highly symmetric shape.<br />

First TEM investigations showed that the dots are terminated by {100}, {110} and<br />

{111} planes (see Fig.2.6).<br />

Figure 2.6: (a) Two PbTe dots precipitated from a 5 nm layer. The dots have a highly<br />

symmetric shape and sharp interfaces to the CdTe host material. (b) High resolution<br />

TEM image of a small dot. The PbTe is identifiable by the (110) and (001) lattice planes.<br />

CdTe shows {111} lattice planes. The PbTe dot shows sharp interface to the host<br />

material. The kinds of the interfaces are indicated.<br />

11


12<br />

The equilibrium shape is constructed with these first tree Miller planes. The shape<br />

built up with these planes is a small cubo-octahedron, or, for the special case that the<br />

{110} and {001} interfaces are quadratic, a rhombi-cubo-octahedron (see Fig.2.7).<br />

The definition of a cubo-octahedron is the combined shape of a cube and an<br />

octahedron which leads to a cube with truncated edges and corners. The {001}<br />

interfaces represent the original cube interfaces. There are manifold appearances of<br />

the cubo-octahedron if the sizes of the chopped parts, which build the {110} and<br />

{111} interface classes, changes. The equilibrium shape has important consequences<br />

for interface characterisation. The interface must be parallel to the electron beam for<br />

interface characterisation with high resolution TEM imaging. The {111} interfaces of<br />

the QDs are of triangular shape. If this interface is parallel to the electron beam and is<br />

focused, the projected length of the interface changes and makes it difficult to image<br />

this interface.<br />

Figure 2.7: A small rhombi-cubo-octahedron. The {100} and {110} surfaces have the<br />

same area and are quadratic. Each facet has a parallel associate surface. The shape is<br />

highly symmetric. The {111} surfaces are triangles.


Chapter 3<br />

The Transmission Electron Microscope<br />

The Transmission Electron Microscope (TEM) is a well-known and often used<br />

instrument for the characterisation of nanostructures. All investigations of the QDs<br />

were done with the TEM, thus, it is important to discuss the used techniques in detail.<br />

This chapter provides only a brief introduction to the basic concepts and the main<br />

components of a TEM considering the JEOL JEM-2011 FastTem that was used within<br />

this thesis. The main part of this chapter is then about the used techniques and the<br />

theory behind them. For more detailed information about the basics and further<br />

information, please see [30], [31], [32] and [33].<br />

3.1 The JEOL JEM-2011 FastTem<br />

The basic principle of a TEM is similar to an optical transmission light microscope,<br />

but it uses electrons for the exploration of the specimen. The manipulation of the<br />

electrons is done by magnetic lenses. An exception is the Wehnelt electrode, which<br />

acts as an electrostatic lens. The main parts of a TEM are the electron gun, the<br />

condenser lens system, the specimen stage and the imaging system, consisting of<br />

objective, intermediate and projector lens system. The task of the electron gun is the<br />

generation of a sufficient amount of electrons with a distinct energy, which are then<br />

formed to a parallel beam in the condenser lens system. After the interaction of the<br />

electrons with the specimen, the imaging system reproduces an enlarged image of the<br />

sample on the luminescent screen or a CCD camera. A schematic cross sectional<br />

image of our TEM setup can be seen in Fig. 3.1 [34]<br />

The mode of the electrons is no longer a single beam after the specimen stage. The<br />

crystal structure creates a multitude of multiple scattered electron beams. One can<br />

either image this diffraction pattern (DP) or a real image of the sample on the<br />

13


14<br />

Figure 3.1: Cross sectional view through the column of the JEOL JEM-2011 [34]


screen. These are the two basic modes of imaging (see Fig. 3.2) [32]. Any of the<br />

scattered electron beams are focused into the focal plane of the objective lens, but<br />

they also build up a real image of the sample in the image plane of the objective lens.<br />

Now the setting of the intermediate lens determines the image on the viewing screen.<br />

The diffraction pattern is reproduced on the viewing screen, when the objective plane<br />

of the intermediate lens coincides with the focal plane of the objective lens. On the<br />

other hand, if the image plane of the objective lens is the objective plane of the<br />

intermediate lens, a real image of the specimen appears on the screen.<br />

Figure 3.2: The two basic operation modes of a TEM are controlled by the strength of<br />

the intermediate lens. (a) The focal plane of objective lens and objective plane of the<br />

intermediate lens are at the same level. The projector lens reproduces the diffraction<br />

pattern on the screen. (b) The objective plane of the intermediate lens is in-plane with the<br />

image plane of the objective lens and a real image of the specimen is formed on the<br />

screen. [32]<br />

15


16<br />

There are two important, adjustable apertures between the objective and<br />

intermediate lens. The first is the objective aperture, which is located at the level of<br />

the focal plane and allows the selection of electrons from distinct scattered beams for<br />

image formation. The other is placed in the image plane of the objective lens and acts<br />

as a virtual aperture to choose the area of interest on the sample. It is called field-<br />

limiting aperture or selected area diffraction aperture (SAD).<br />

The JEOL JEM 2011 FastTem uses a LaB6 filament as electron source and the<br />

electrons can be accelerated up to 200keV between the Wehnelt cylinder and the<br />

anode plates. Electrons with an energy of 200keV are in the range where relativistic<br />

effects are playing a role. This leads to a wavelength of the particles of 2.5·10 -2 Å.<br />

The microscope is an electron system, thus the theoretical diffraction limit is given by<br />

the Rayleigh criterion (eqn. 3.1) [32].<br />

The denominator of equation 3.1 is called the numerical aperture and is the product of<br />

the diffraction-index and sin β, whereas β is the semiangle of collection. λ refers to<br />

the wavelength of the electrons. For an electron energy of 200keV this wavelength is<br />

2.5·10 -2 Å. The maximum resolution (the numerical aperture is set to 1) is then<br />

1.53·10 -2 Å.<br />

0.<br />

61⋅<br />

λ<br />

δ =<br />

n ⋅sin<br />

β<br />

The practical resolution is much worse than the theoretical diffraction limit. The<br />

main reasons are lens aberrations and in particular the spherical aberration of the<br />

magnetic lenses (see Fig. 3.3) [32]. The spherical aberration causes a stronger bend of<br />

off-axis electrons toward the optical axis, and thus a spot is reproduced as a disk in the<br />

image plane. The radius of this disk is given by Csβ 3 . β is again the semiangle of<br />

collection and Cs the spherical aberration coefficient, which is an important<br />

characteristic of a TEM. With this value and the Rayleigh criterion one can estimated<br />

a practical point-to-point resolution (eqn. 3.2) [32] of the microscope.<br />

r<br />

min<br />

≈ s<br />

( ) 4<br />

1<br />

3<br />

C<br />

0. 91 λ<br />

(3.1)<br />

(3.2)


This leads with Cs = 1mm and λ = 2.5·10 -2 Å (parameters of the JEM 2011 FastTEM<br />

at 200 keV) to a resolution of about 2 Å. This is near the specified point resolution of<br />

the JEM 2011 FasTEM for 200keV of 2.3 Å, [35].<br />

Figure 3.3: Ray diagram of the lens aberration. A point P is reproduced on the Gaussian<br />

image plane as a disk by a real lens. Due to the stronger bend of an off-axis electron, a<br />

disk with a minimum radius is found in the plane of least confusion. [32]<br />

3.2 Scattering of Electrons<br />

Scattering of electrons at the atoms are the interaction with the specimen. We can<br />

characterise the scattering depending on how we look at the electron, as particle or as<br />

a wave. The expression inelastic or elastic scattering describes whether the particle<br />

electron transfers energy to the target or not. Coherent and incoherent scattering<br />

distinguish between scattering without a phase shift of the incident and the scattered<br />

electron wave and an event with a phase shift.<br />

17


18<br />

Elastic scattering is described by Rutherford scattering, with the important<br />

parameter of the differential cross section which (eqn. 3.3, eqn. 3.4) [32] defines the<br />

atomic form factor (eqn. 3.5) [32].<br />

an isolated atom.<br />

2<br />

f (θ)<br />

is proportional to the scattered intensity of<br />

Equation (3.3) is the screened relativistic Rutherford cross section, a function of<br />

the scattering angle θ. θ0 (eqn. 3.4) is the so-called screen parameter. λR is the<br />

relativistic corrected electron wavelength and E0 the according electron energy. The<br />

properties of the atom are described by the atomic number Z. a0 denotes the Bohr<br />

radius. For a more detailed explanation of the Rutherford scattering see e.g. [32,33].<br />

The structure of the crystal is described by the unit cell structure factor (eqn. 3.7,<br />

detailed calculation for the zb and rs lattice in appendix A). With the Laue equation<br />

K=kD-kI this leads to the description of a diffraction pattern. For a detailed<br />

introduction of reciprocal space and diffraction from a crystal, see e.g. [36].<br />

The intensities of the diffraction spots contain the information of the contrast of a<br />

real image. Therefore, it is important to understand the mechanism behind it to<br />

interpret the contrast of TEM images. The amplitude of an e-beam scattered at one<br />

unit cell is given by [32]:<br />

4<br />

dσ ( θ )<br />

λR<br />

dΩ<br />

=<br />

4<br />

64π<br />

a<br />

A<br />

θ<br />

0<br />

f ( θ)<br />

cell<br />

F (<br />

θ)<br />

=<br />

=<br />

2<br />

0<br />

2<br />

e<br />

=<br />

∑<br />

i<br />

Z<br />

2<br />

2<br />

⎛<br />

⎞<br />

⎜ 2 θ ⎛ θ0<br />

⎞<br />

sin + ⎟<br />

⎜<br />

⎜ ⎟<br />

2 2 ⎟<br />

⎝ ⎝ ⎠ ⎠<br />

0.<br />

117<br />

E<br />

0<br />

1<br />

2<br />

Z<br />

1<br />

3<br />

dσ(<br />

θ)<br />

=<br />

dΩ<br />

2 i r k⋅ π<br />

r<br />

fie<br />

2πi<br />

F<br />

( θ<br />

)<br />

( hx + ky + lz )<br />

i<br />

i<br />

i<br />

2<br />

(3.3)<br />

(3.4)<br />

(3.5)<br />

(3.6)<br />

(3.7)


F(θ) is the structure factor of the unit cell and k is the wave vector of the scattered<br />

beam. Summation over all unit cells leads to an expression for the amplitude of a<br />

scattered electron beam.<br />

This is a summation over n unit cells in an unit area plane separated by the distance t.<br />

The extinction distance ξg (eqn. 3.9, [32]) is an important quantity to describe a<br />

scattering event. It contains the lattice parameters (via the volume Vc of the unit cell),<br />

the atomic number (via Fg, the structure factor of the unit cell for the Bragg angle θB)<br />

and the wavelength λ of the electrons.<br />

Equation 3.8 is an expression of the kinematical theory of electron diffraction,<br />

which is only valid for a thin crystal slice. The intensity of the scattered electron beam<br />

|φg| 2 is proportional to t 2 . If only two electron beams, the direct and one scattered<br />

electron beam, are considered, |φg| 2 is small and the intensity |φ0| 2 of the direct beam is<br />

approximately one. This is valid if the condition t


20<br />

To derive the change of the beam amplitude the two-beam approximation is<br />

applied. This means that only one diffracted beam is strong. This condition can also<br />

be reached in a real TEM by tilting the sample until only one beam besides the (000)<br />

reflex is strong. The other spots can then be neglected. If one writes now this equation<br />

in terms of the change of the amplitudes when the beam passes through the specimen<br />

along the z direction, one obtains the Howlie - Wehlan equations (eqn. 3.11, eqn.<br />

3.12,) [32].<br />

dφ g πi<br />

−2<br />

πisz<br />

= φ 0e<br />

dz ξ g<br />

πi<br />

+ φ<br />

ξ<br />

dφ0 πi<br />

πi<br />

2πisz<br />

= φ0<br />

+ φge<br />

dz ξ0<br />

ξg<br />

This pair of coupled differential equations can be used to derive the amplitude of the<br />

scattered beam. This gives the intensity of a diffracted beam<br />

Whereas seff is the effective excitation error and is given by<br />

φ<br />

g<br />

2<br />

⎛ t ⎞<br />

⎜<br />

π<br />

= ⎟<br />

⎜ ⎟<br />

⎝ ξg<br />

⎠<br />

s is the excitation error and expresses the deviation from the exact Laue condition and<br />

is also a vector in reciprocal space. Many diffraction spots can be seen even if the<br />

Laue condition is not exactly fulfilled, which is then expressed as K=kD-kI+s. The<br />

intensity of the scattered electron beam is a periodic function of the specimen<br />

thickness t. The period of the oscillations is πseff. seff differs for different diffraction<br />

spots (hkl) and approaches s for very large values of s and becomes ξg -2 if s equals<br />

zero. The periodic behaviour of the electron beams plays an important role in the<br />

interpretability of the lattice fringes in HRTEM images (see section 3.4).<br />

s<br />

eff<br />

2<br />

=<br />

0<br />

g<br />

( )<br />

( ) 2<br />

2<br />

πts<br />

eff<br />

πts<br />

sin<br />

s<br />

2<br />

eff<br />

+<br />

1<br />

ξg<br />

2<br />

(3.11)<br />

(3.12)<br />

(3.13)<br />

(3.14)


3.3 Contrast Enhancement with Bright and Dark Field<br />

Imaging<br />

To get sufficient information from a TEM image it is important to know what the<br />

origin of contrast in the image is and how one can use these mechanisms to enhance<br />

it. The term of contrast itself is defined as a difference in intensity between to areas<br />

(egn. 3.15), [32].<br />

Therefore, it is not possible to increase contrast by increasing the intensity of the<br />

electron beam. Rather one will decrease the contrast and the best condition to record<br />

high contrast images is to do it with low intensity.<br />

3.3.1 Contrast Mechanism<br />

There are three main contrast mechanisms that are important for TEM imaging. The<br />

mass-thickness (ρt, density×thickness) contrast is caused by incoherent elastic<br />

electron scattering. The Rutherford cross section depends strongly on the atomic<br />

number Z (eqn. 3.3). Thicker samples cause more scattering events due to the mean<br />

free path remains constant. In a specimen region with a higher mass-thickness more<br />

electrons are scattered, thus less electrons are collected by the objective lens in the<br />

image plane (see Fig. 3.4), [32]. On the other hand, coherent elastic scattering causes<br />

the diffraction contrast. This contrast is a form of amplitude contrast, because<br />

diffraction peaks appear at the Bragg angles. The intensity of the diffraction peaks<br />

depends on the unit cell structure factor and can be used to produce contrast between<br />

materials with different structure.<br />

The mass thickness effect and the diffraction contrast compete. Given that<br />

Rutherford scattering in a thin specimen is strongly forward peaked, the primary spot<br />

of the DP contains most of the mass thickness contrast. This spot can be used to form<br />

a real image which is called bright field image (BF). On the other hand, collecting the<br />

electrons of any diffraction spot with dominant diffraction contrast for imaging is<br />

called Dark field (DF) imaging.<br />

I<br />

C =<br />

2<br />

− I<br />

I<br />

1<br />

1<br />

ΔI<br />

=<br />

I<br />

1<br />

21<br />

(3.15)


22<br />

The origin of the third contrast mechanism is interference of differently diffracted<br />

e-beams. This contrast allows to resolve lattice fringes in High Resolution<br />

Microscopy which is treated in section 3.4.<br />

Figure 3.4: In regions with low mass thickness more electrons are forward scattered to<br />

small angles and can be collected by the objective aperture. This causes a higher intensity<br />

in the formed image. In higher mass thickness regions more scattering events occur<br />

causing darker regions in the image. [32]<br />

3.3.2 Dark Field and Bright Field imaging<br />

Since the diffraction pattern of the crystal is reproduced in the focal plane of the<br />

objective lens and the objective aperture is located in the same plane, a distinct<br />

diffraction spot can be chosen in the DP mode for real image formation. To do that all<br />

other spots are masked by the aperture and only the electrons of one beam are<br />

collected by the intermediate lens. BF images are formed by collecting only the<br />

electrons of the (000) peak by the objective aperture and show strong mass thickness<br />

contrast.<br />

On the other hand, collecting the electrons of any diffraction spot can show<br />

contrast in the real image, whose origin are different lattice parameters or a different<br />

structure form factor. Both of these properties are a part of the extinction distance ξg<br />

(egn. 3.9) which is a determining factor of the amplitude of a scattered beam. This


kind of imaging is called Dark field (DF) imaging. In Figure 3.5 one can see a<br />

comparison of dark and bright field imaging.<br />

Figure 3.5: Sketch of (A) BF and (B) DF imaging. The electrons of a distinct diffraction<br />

spot can be collected for image formation by alignment of a small enough objective<br />

aperture. [32]<br />

Since the diffraction contrast is correlated with the intensity of the Bragg reflexion,<br />

one can increase the contrast of the real image by optimizing the amplitude of the<br />

scattered electron wave. The intensity of the scattered Bragg reflex is periodic in the<br />

thickness t of the sample and periodic in the effective extinction error seff, which can<br />

be changed by tilting the sample. The best results are obtained for the two beam<br />

condition, where only the spot of interest has a high intensity. Of course, the other<br />

bright spot is the (000) reflex.<br />

The DF technique is best suited for specimens with different crystal structure. For<br />

the CdTe/PbTe heterosystem which is treated within this thesis, DF imaging is really<br />

powerful. CdTe crystallizes in the zb lattice and the formfactor of {200} reflexes is<br />

weak. On the other hand the formfactor F{200} for the PbTe rs structure is strong, thus<br />

this reflex is well suited for DF images. In Figure 3.6 one can see the comparison of a<br />

dark and a bright field image at the same position of a CdTe/PbTe sample. The<br />

diffraction pattern of the specimen is also shown. The images were recorded slightly<br />

away of the [001] zone axis.<br />

23


24<br />

Figure 3.6: (a) BF image of a PbTe/CdTe specimen, recorded slightly away of the [001]<br />

direction. The images show the mass-thickness contrast. The PbTe dots appear darker<br />

than the surrounding CdTe host material (mass contrast). A thickness contrast can be<br />

seen at the top and the bottom of image (a). The diffraction contrast is dominating in the<br />

DF image (b). The (020) spot, which is a forbidden reflex for the zb lattice, was used for<br />

image formation. Thus, the PbTe dots appear bright in a dark background. In addition,<br />

the dots in the thicker areas, which are not visible in image (a) are visible in image (b).<br />

(c) shows a diffraction pattern, that was recorded slightly away from the zone axis, so<br />

that the (020) reflex is the brightest of the {200} spots.<br />

Figure 3.7: Sketch of the diffraction condition for CDF. (A) Common two beam<br />

condition. The spot +g(hkl) and 000 are excited. If now the +g(hkl) spot is moved to the<br />

position of the 000 spot it becomes weak and the +g(hkl) spot is strong (B). If the<br />

opposite –g(hkl) spot is tilted toward the 000 position (C) the reverse two-beam<br />

condition is now applied to this spot. [32]


Centred Dark Field<br />

One disadvantage of the straight-forward DF imaging procedure is increased lens<br />

aberration. Scattered e-beams are not near the optical axis of lenses, so they are<br />

stronger bent towards this axis and it is difficult to form high quality real space<br />

images.<br />

To record DF images in the same quality as BF images it is necessary to bring the<br />

e-beam of interest on the optical axis and into a two beam condition. That can be done<br />

by tilting the incident beam in such a way that the spot of interest replaces the (000)<br />

spot. This kind of DF imaging is called Centered Dark Field (CDF) mode. The tilting<br />

of the incident beam changes the diffraction conditions for the electrons of the hkl-<br />

plane and the applied two-beam condition is destroyed. Both conditions can be<br />

reached if first the symmetric spot of the diffraction spot is tilted to the two-beam<br />

condition and after that the incident beam is tilted. This procedure is sketch in Figure<br />

3.7, [32]. Most TEM are able to save the alignments for the beam tilts, so it is possible<br />

to switch simply between DF and BF imaging, once the objective aperture is well<br />

aligned.<br />

3.4 High Resolution Transmission Electron Microscopy<br />

Another kind of investigation of specimens is High Resolution Transmission Electron<br />

Microscopy (HRTEM). The main difference to DF and BF imaging is the number of<br />

collected beams for image formation. To form a HRTEM image more than one beam<br />

is collected by the objective aperture (see Fig. 3.8). If the resolution of the microscope<br />

is high enough, so-called phase contrast can be obtained. This phase contrast contains<br />

the spatial information of the lattice planes of the sample and can be exploited to<br />

explore the crystal structure.<br />

3.4.1 Requirements for HRTEM<br />

There are some essential requirements for the microscope to obtain high resolution<br />

TEM images. [33]<br />

• High mechanical stability, low vibrations<br />

• A goniometer of high precision<br />

• High enough acceleration voltage to reach the necessary resolution<br />

• High gun brightness<br />

25


26<br />

• Stability of the acceleration and lens currents<br />

• On-axis alignment of the electron beam and the voltage centre of lens<br />

current<br />

• Low spherical aberration coefficient of the objective lens<br />

• Parallel beam for good spatial coherence<br />

• Low energy spread and low chromatic aberration coefficient<br />

• Well aligned stigmators of the intermediate and objective lens to avoid<br />

lens astigmatism<br />

Figure 3.8: (a) Diffraction pattern of a CdTe specimen, recorded along the (110) zone<br />

axis. The most important spots that cause the phase contrast of the lattice planes in the<br />

real image, are indicated. The circle indicates the size of the aperture that was used for<br />

real image recording. (b) Real image that shows phase contrast. The {002} and {220}<br />

diffraction spots are strong for the PbTe rs structure, the {111} are strong for the CdTe<br />

zb structure. The two materials can be distinguished by the corresponding lattice planes<br />

in the real image. The lattice planes are indicated.<br />

3.4.2 Origin of Lattice Fringes<br />

The origin of the lattice fringes are the different phase factor of the scattered beams.<br />

This can be easily shown by the solution of the Howlie - Wehlan equations. The<br />

intensity of the solution is given for the two-beam approximation by [32]<br />

2<br />

2 2<br />

I = ψT<br />

= A + B + 2AB<br />

sin( 2πg′<br />

x − πst)<br />

(3.16)


Whereas A 2 and B 2 are the intensities of the primary and scattered beam and the<br />

term 2AB is the interference part. This equation also implements the assumption that<br />

the specimen is thin enough to set sg equal to s. The intensity of the two beams is a<br />

sinusoidal function of g´ (g´ = g+sg). If the term πst is a constant, one will see this<br />

variations as lattice fringes with a periodicity of x = 1/g. If the excitation error s = 0 it<br />

is the exact lattice plane spacing of the scattering planes. If the excitation error is not<br />

zero and there are no variations of the thickness t it will be still the same case. For a<br />

not entirely flat specimen the lattice fringes will be shifted as a function of πst, but the<br />

periodicity will not change drastically.<br />

3.4.3 Transferfunction and Scherzer Defocus<br />

As image formation in a microscope is described in terms of a specimen function<br />

f(x,y), the image function g(x,y) can be written as the convolution of f(x,y) with a so<br />

called impulse response function h(x,y) (eqn. 3.17), [32]. h(x) describes the properties<br />

of the optical system.<br />

g( r) = ∫ f ( r′<br />

) h(<br />

r − r′<br />

) dr′<br />

= f ( r)<br />

⊗ h(<br />

r)<br />

A Fourier transform of a convolution results in a multiplication of the fourier<br />

transformed terms in the reciprocal space. This terms are the contrast transfer function<br />

H(u), the specimen function F(u) and the image function G(u).<br />

G( u ) = H(<br />

u)<br />

F(<br />

u)<br />

H(u) contains the properties of the optical system of the microscope as the aperture<br />

function A(u), the envelope function E(u) and the aberration function B(u).<br />

H( u ) =<br />

A(<br />

u)<br />

E(<br />

u)<br />

B(<br />

u)<br />

A(u) and E(u) restrict the spatial frequency range, whereas A(u) depends on the<br />

aperture size and E(u) is a fixed property of the lens system. Depending on the chosen<br />

aperture, A(u) can be more or less restrictive. B(u) is usually given by eqn. 3.20 and<br />

eqn. 3.21, the phase-distortion function χ(u). A point like object is imaged as disk due<br />

27<br />

(3.17)<br />

(3.18)<br />

(3.19)


28<br />

to defocus Δf and spherical aberration, expressed by Cs. The phase-distortion function<br />

brings that into account for all rays passing through the lens system with the angles<br />

θ = λu.<br />

The specimen is described as a phase object. It can be shown that the phase change<br />

of the electrons only depend on the crystal potential V(x,y,z) if the specimen is thin<br />

enough. The 3-dim potential is then replaced with the projected potential Vt(x,y) on<br />

the x-y plane perpendicular to the electron beam. The so called weak phase object<br />

approximation (WPOA, eqn. 3.22), [32] applies for very thin specimens (Vt(x,y)


Figure 3.9: Plots of the sin x(u) term for different values of Δf and Cs. The first cross<br />

over defines the change form a positive phase contrast to a negative phase contrast. If the<br />

spatial frequencies of the lattice distances of interest is smaller than this first cross over,<br />

the image may be directly interpreted. The effect of the defocus Δf can be balanced<br />

against that of the spherical aberration to optimize the first cross over. The plots are<br />

calculated for 200keV electrons. The plot for Δf=-60nm, Cs=1mm is the optimized<br />

transferfunction for the Jeol JEM 2011 FasTEM (Scherzer defocus)<br />

But there is a way to improve the transferfunction and shift the first cross over to<br />

large value. Scherzer found in 1949 that one can balance the effects of the spherical<br />

aberration against an underfocus of the sample (see Fig. 3.9). This is known as the<br />

Scherzer defocus (eqn.: 3.24), [32].<br />

This leads to a resolution (eqn. 2.25), [32] that can be defined as the reciprocal<br />

value of the first cross over sin χ(u).<br />

Δf Sch<br />

s<br />

= −1.<br />

2(<br />

C λ)<br />

1<br />

2<br />

29<br />

(3.24)


30<br />

For the JEM 2011 FasTEM with a Cs of 1mm and 200keV electrons with a λ of<br />

2.5·10 -2 Å this equation leads to a resolution of 2.3 Å which is equal to the specified<br />

resolution by JEOL [35].<br />

3.5 Sample Preparation<br />

The aim of sample preparation is the production of thin enough samples. Electrons<br />

can only pass through samples thinner than about 100nm. For HRTEM imaging, the<br />

sample must be even thinner and should be only a few nanometers thick if it should be<br />

simple to interpret the images. The sample holder of the Jeol JEM2011 FastTem<br />

requires samples with a diameter of 3mm. These small disks are thinned to the<br />

necessary thickness. For a detailed description of the Sample Preparation see [40].<br />

One distinguishes between two kinds of preparations concerning the direction of<br />

view with respect to the layer structure of the specimen. With a cross sectional sample<br />

(X-specimen) one looks along a direction in the plane with the layer of interest. With<br />

a plan view (P-specimen) sample one looks along a direction perpendicular to the<br />

layer. The material is cut with a wire saw or an ultrasonic cutting device [37]. After<br />

the fabrication of the blank shape (for our sample holder a disk with a diameter of<br />

3mm), it is thinned by grinding on a rotation wheel to about 150 μm. The next step is<br />

done with the so called dimple grinder [38]. A rotating wheel grinds a dimple in the<br />

middle of the rotating disk such that a supporting ring remains. The thinnest area is in<br />

the middle of the dimple and is about 10 μm thick. The last step of thinning is done by<br />

ion milling with argon ions [39] down to 10 nm or less.<br />

The Ar ions are accelerated with 2-3kV for the sputtering process and thus result in<br />

a heating of the sample during the thinning. The sputtering time for the thinning<br />

process varies strongly because grinding with the dimple grinder is only accurate on<br />

the μm scale. The transferred load of heat can cause additional annealing of the<br />

sample. Specimens of un-annealed samples show in some case the beginning stage of<br />

QD precipitation. But this could also be an effect of the CdTe overgrowth. The<br />

additional transferred heat has no effect for annealed samples with QDs in the<br />

equilibrium shape.<br />

r<br />

Sch = s<br />

1<br />

4<br />

3<br />

4<br />

0. 66C<br />

λ<br />

(3.25)


Chapter 4<br />

Shape and size of the PbTe Dots<br />

For the implementation of the QDs for devices one must be able to control the size,<br />

density, shape and position of the nanocrystals. The size of the dots should be<br />

controlled by the layer thickness and the density dependents on the amount of<br />

deposited material. The shape is determined by a thermodynamic equilibrium shape.<br />

The in-depth location of the dots is well controlled by the position of the epi-layer. To<br />

learn more about the size and shape of the dots in dependence of the annealing<br />

conditions and the layer thickness, DF and BF images of several samples were<br />

statistically evaluated.<br />

4.1 Available Samples<br />

The samples were grown at the Osaka <strong>Institut</strong>e of Technology, as mention in chapter<br />

2, and were originally intended for photoluminescence measurements. The samples<br />

were annealed at our institute in Linz. The annealing parameters were adjusted to<br />

optimise the PL measurements. The series of samples, grown at different conditions,<br />

were thus not perfectly suitable for the statistical analysis.<br />

Table 4.1 shows the series of the investigated samples. The most common<br />

thickness of the SQWs was 5nm PbTe, but there was also one attempt of a double<br />

quantum well (samples with series number #11). The quantum wells were grown at<br />

different substrate temperatures from 280°C to 220°C. All PbTe layers were grown on<br />

a Cd terminated CdTe surface, expect for series number #8. The Cd termination was<br />

preferred, because PbTe grown on this termination show less surface roughness [26].<br />

The annealing temperature was in the range of 320°C to 350°C. To investigate the<br />

influence of the layer thickness on the resulting dots a series was grown with SQW<br />

thickness from 1nm to 50nm. The samples with the highest PL intensity (#28-#31)<br />

were used for TEM investigation. For the specimen fabrication we used in most cases<br />

annealed samples, but also as-grown samples were used. First we prepared a cross<br />

31


32<br />

section specimen and, if the TEM images showed interesting features, we tried to<br />

fabricate an additional plan view sample. Because a limited amount of material was<br />

available, and not all attempts of sample preparation were successful, not always both<br />

preparation kinds became available.<br />

TEM specimen numbers<br />

(epi-number following #)<br />

Kind of<br />

Preparation<br />

Cross section(X)<br />

/ Plan view(P)<br />

Substrate<br />

temperature<br />

during SQW<br />

growth<br />

Annealing<br />

temperature<br />

Annealing<br />

time<br />

Termination<br />

before PbTe<br />

growth<br />

PbTe<br />

well<br />

thickness<br />

[°C] [°C] [min] [nm]<br />

192X_SQW_A_CdTe X 280 as grown Cd 5<br />

188X_SQW_B_CdTe X 280 as grown Cd 5<br />

201X_#5_G_CdTe X 250 as grown Cd 5<br />

227P_#5_C_CdTe P 250 cooled Cd 5<br />

204X_#5_C_CdTe X 250 cooled Cd 5<br />

242P_#8_G_CdTe P 280 as grown Te 5<br />

235X_#8_CdTe_G X 280 as grown Te 5<br />

243X_#8_C_CdTe X 280 cooled Te 5<br />

231P_#8_A_CdTe P 280 350 10 Te 5<br />

234X_#8_CdTe_A X 280 350 10 Te 5<br />

223X_#11_G_CdTe X 280 as grown Cd 5<br />

226X_#11_C_CdTe X 280 cooled Cd 5<br />

225X_#11_A_CdTe X 280 350 10 Cd 5<br />

254P_#28_A_CdTe P 220 320 10 Cd 1<br />

258X_#28_A_CdTe X 220 320 10 Cd 1<br />

253P_#29_G_CdTe P 220 as grown Cd 3<br />

245P_#30_A_CdTe P 220 320 10 Cd 3<br />

247X_#30_A_CdTe0 X 220 320 10 Cd 3<br />

248X_#30_A_CdTe90 X 220 320 10 Cd 3<br />

244P_#31_A_CdTe P 220 320 10 Cd 5<br />

246X_#31_A_CdTe0 X 220 320 10 Cd 5<br />

249X_#31_A_CdTe90 X 220 320 10 Cd 5<br />

228X_CdTe_1974_1016 X<br />

229X_CdTe_1974_1015 X implanted<br />

257X_CdTe_Pb_1610 X<br />

Table 4.1: List of all investigated specimens within this thesis. The # in the TEM<br />

specimen numbers marks the epi-layer series, of which cross section and plan view<br />

specimens were fabricated. As-grown and annealed samples were prepared. The last<br />

three samples were an attempt to produce the PbTe/CdTe dots by implanting Pb into bulk<br />

CdTe.


4.2 Precipitation steps<br />

TEM images of as-grown samples showed that in most cases the 2D layer does not<br />

exist any more after the preparation process. If the 2D layer was not breaking up<br />

during the overgrow process, the start of the disintegration of the layer was induced<br />

by the heat load which was transferred to the sample during the last sample thinning<br />

process, argon ion sputtering. This fact allows some insight into the processes during<br />

the annealing step. The sputtering time varies strongly for different samples.<br />

Therefore, different intermediate steps can be found in these nominally un-annealed<br />

samples, from a nearly closed 2D layer to separated material in off-equilibrium dot<br />

sizes and shapes. The dots of the annealed specimens assume the equilibrium shape,<br />

or they are on the way toward it, depending on the annealing temperature and the<br />

original layer thickness. With this it is possible to demonstrate the precipitation<br />

process from the start of up-breaking of the 2D layer to highly symmetric PbTe dots<br />

in the thermodynamic equilibrium shape.<br />

Figure 4.1: BF TEM image of a nominally non-annealed PbTe layer recorded along the<br />

[001] zone axis of a plane view specimen of series #29. The 2D layer of this series was<br />

3nm thick and disintegrates during the ion milling process. The lateral dimensions of the<br />

wiggly lines are about 10 -15nm. The lines are orientated along the and <br />

directions.<br />

The breaking-up of the PbTe layer starts with a disintegration along extended lines<br />

and building of interconnecting lines of material. Figure 4.1 shows a plan view<br />

specimen of an as-grown sample. The image shows the contracted PbTe material with<br />

the shape of wiggly lines. In this stage most of the PbTe is still interconnected, but<br />

also some separated material in dot size can be found. The material contracts laterally<br />

33


34<br />

and increases the height of the PbTe material lines. The increase of the local layer<br />

thickness can be seen in Fig. 4.2. The image shows a cross-sectional specimen of an<br />

originally 5nm epi-layer sample. The PbTe wires contracts until the thickness of the<br />

Figure 4.2: BF TEM image of an annealed PbTe layer recorded slightly away from the<br />

[110] zone axis on a cross sectional specimen of series #31. Growth direction was along<br />

the [001] axis. The sample was originally a 5 nm SQW and was annealed at 320°C for<br />

10min, which was not sufficient for complete precipitation. The sample contains still<br />

interconnected 2D part with a height form 5 to 20 nm. The lateral dimension shrinks for<br />

thicker layers. Dot-shaped material with sharp interfaces can be seen in the 20 nm layer.<br />

Figure 4.3: Images of several constrictions of the PbTe wires during the dot<br />

precipitation. The images were recorded on a cross section specimen of series #5, as-<br />

grown sample. (a) shows a DF image, (b) and (c) show BF images. The layer was<br />

originally 5nm thick. The dot material already shows sharp interface to the host material.<br />

The shape of the separated material shows larger in-plane dimensions than height.


lines approaches the diameter of the resulting dots. The wires began to constract at<br />

some points to provide sufficient material for this height increase (see Fig. 4.3). At<br />

these constrictions the PbTe wire becomes thinner and thinner and the developing<br />

PbTe wires are separated into PbTe islands. Figure 4.4a shows TEM images of plan<br />

view and cross sectional specimens in this stage. The dots have a elongated shape and<br />

the lateral dimensions are larger than the height of the dots.<br />

Figure 4.4: (a) Cross sectional (DF image) and plan view (BF image) of an annealed<br />

3nm PbTe layer (#30). The dots are elongated and disk-shaped, and show sharp<br />

interfaces (indicated in the plan view image). The height of the dots is about 10nm. The<br />

longer interface in the plan view is either the (010) or the (110) interface, but also highly<br />

symmetric dots can be found. (b) Cross sectional (BF image) and plan view (DF image)<br />

of dots from a 1nm layer, annealed at the same conditions as (a). The dot size is<br />

distributed over some range, but only highly symmetric dots can be found and the sample<br />

seems to be in thermodynamic equilibrium.<br />

The separated dots then try to reach their thermodynamic equilibrium shape (see<br />

fig. 4.4b). The distribution of the constrictions, which separate the PbTe wires into<br />

dots, determines the final size distribution of the QDs. The PbTe wire dimensions are<br />

different for different thicknesses of the original epi-layer. The PbTe wire dimensions<br />

determine the maximum distance between two constrictions which lead to a relatively<br />

sharp upper limit of the resulting dot size. Smaller dots can be found over a larger<br />

range.<br />

35


36<br />

4.3 Precipitation from different layer thicknesses<br />

4.3.1 Annealed 5nm layer<br />

The best specimen of an originally 5nm thick layer resulted from sample #11, which<br />

is the double QW sample. It was possible to record more than 100 dots in one row,<br />

which allows significant dot size statistics. The double quantum well samples were<br />

annealed at 350°C for 10 min. The height (along the growth direction) and the width<br />

of the dots were evaluated with the program package Gatan Digital Micrograph.<br />

Digital Micrograph is a control and analysis software for the Gatan Multi-Scan CCD<br />

camera employed on our TEM [41]. The specimen itself shows dots with almost the<br />

same height, but the width of the dots varies significantly (see fig. 4.5). Smaller dots<br />

with a decreased high show always a high symmetry with comparable height and<br />

widths. Dots with the maximum height have lengths of up to twice their height. This<br />

is a strong indication that not all dots have reached there thermodynamic equilibrium<br />

shape.<br />

Figure 4.5: (a) Cross sectional, dark field image over a large area. The sample contains a<br />

double row of QDs. The bottom row was used to analyse the dimensions of the dots. (b)<br />

Typical QDs found in the sample, recorded as bright field image. The left-hand image<br />

shows two highly symmetric QDs. The crystal orientation is indicated. The QDs show<br />

sharp interfaces. The kind of the interfaces indicated in the right-hand images, which<br />

show two elongated, centro-symmetric dots.


Figure 4.6: The distribution of the measured QDs in sample #11. Cubic dots with<br />

quadratic cross-sections lie on the solid line. Symmetric dots with an aspect ration from<br />

0.8 to 1.25 are bordered by the dashed line. The TEM images in the insets show one dot<br />

at the solid line, one at the dashed line and one dot that is of elongated shape.<br />

Figure 4.6 shows the distribution of the measured dots. The height of the dots has<br />

an upper limit of about 30 nm. Most of the dots show a quadratic shape in the cross<br />

section, but for heights larger than 15 nm also elongated dots can be found. Dots,<br />

which are smaller than 15 nm, have almost the same height and width. Figure 4.6<br />

contains also some example, of how the dots look in detail. The dots were sorted into<br />

classes of 2nm for the height and length to get the distributions of these dimensions.<br />

The number of dots in one class was counted. The accuracy of the measurements with<br />

Digital Micrograph was about 1nm. Figure 4.7 shows these distributions. The heights<br />

of the dots are narrower distributed than the length, and show a relatively sharp upper<br />

limit of 30nm with a maximum at 25nm. Lower limit is at 6nm, but such small dots<br />

are also hard to see in the dark field images with low magnification. But even in HR<br />

images no dots are found with less than 20 atomic layers, which correspond to a dot<br />

with a size of 4.6 nm. The length of the dots is wider spread. The maximum of the<br />

37


38<br />

distribution is at about 30 nm and limited by 8 and 45 nm. If it is also taken into<br />

account that the dot elongation can be parallel to the line of sight, most of the dots in<br />

this sample are elongated. Unfortunately, no plan view specimen of this sample exists,<br />

but it should look like the plan view image of Figure 4.4, which shows an annealed<br />

3nm sample.<br />

Figure 4.7: (a) Height distribution of the dots. The maximum is at 25nm. The upper limit<br />

is at about 30nm. (b) Length distribution of the dots with the maximum at about 30nm.<br />

4.3.2 Annealed 1nm layer<br />

The same measurements were also performed for sample #28, with an originally 1nm<br />

thick PbTe layer. The sample was annealed at 320°C for 10 min. The plan view<br />

specimen of the sample was good enough for statistical evaluation. A cross sectional<br />

specimen was also available and was used to check the highly symmetric shape,<br />

which the dots show in the plan view sample. This can be seen in Figure 4.8. The<br />

appearance of the dots in plan view is an octagon with longer {110} interfaces. The<br />

distances between two parallel {110} interfaces were used to measure the size of the<br />

dots. The cross sectional specimen shows the appearance of these dots in a line of<br />

sight along the [ 110]<br />

direction which confirmed the high symmetry. One interesting<br />

point is the extended (111) interface of these small dots, which will be treated in part<br />

4.4.<br />

The distribution of the dot size can be seen in Figure 4.9. Figure 4.9a shows the<br />

high symmetry of the dots, which seem to become ever better for larger dots. The<br />

ideal case is again indicated with a solid line. The shift of the distribution to one side


Figure 4.8: The image at the top shows the cross sectional specimen of sample #28 as a<br />

BF image. The (111) interface is most pronounced, and for small dots the other interfaces<br />

almost disappear. The DF image at the bottom shows the plan view specimen. The<br />

interfaces of the dots are indicated.<br />

Figure 4.9: (a) Size distribution of the measured dots. Only highly symmetric dots are<br />

found. (b) Classification with respect to size. The most common size is 7 to 11nm. The<br />

size of the smallest dot was 3.5nm<br />

39


40<br />

of this line is due to a slight tilt away from the zone axis to increase the contrast of the<br />

dark field image. The classification of the dots with respect to their size shows a<br />

narrow distribution with a maximum at about 8nm. The upper limit is again sharper<br />

than the lower one, and appears at 11 nm. The smallest dot was 3.5 nm large, but it<br />

was strongly blurred and could also be 5 nm large. The accuracy of this measurement<br />

was about 0.2 nm if the interfaces are sharp for larger dots and get worse for smaller<br />

dots, which can be explained by the shrinking (110) interface down to almost zero, as<br />

can be seen in the cross sectional image of figure 4.8.<br />

The comparison of these two specimens showed that the size of the dots depends<br />

strongly on the original layer thickness. Unfortunately, the cross-sectional specimen<br />

of a 3nm layer was not good enough for statistical evaluation and comparison with the<br />

5nm sample. A rough estimate of the mean volume of the dots from the 3nm layer<br />

resulted in about 6000nm 3 . An accurate estimate of the mean dot volume of the 1nm<br />

samples is 335nm 3 , which corresponds to a dot density of 3 10 11 cm -2 . The largest one<br />

is that from the 5nm layer with a mean volume of about 8000nm 3 , which corresponds<br />

to a dot density of about 6 10 10 cm -2 . Thinner layers break up in dots with smaller<br />

volume. On the other hand, the larger dots of e.g. a 5nm layer, collect much more<br />

material. The volume of the large dots of the 5nm layer is ~25 time larger than for that<br />

ones of the 1nm layer, but the thicker 5nm layer has only 5 times more material per<br />

unit area than a 1nm layer. Thus, one can find 5 times more small dots precipitating<br />

from a 1nm layer than large ones from a 5nm layer.<br />

4.4 Equilibrium shape<br />

As shown in part 4.3, the shape appearance of the dots changes with their size. The<br />

{111} interfaces do not seem to shrink with the same ratio as the others do, when dots<br />

get smaller. As the {111} facets are the interfaces which are hardest to resolve in high<br />

resolution images. The aim of this chapter is to describe the expected shape of the<br />

dots with attention on the {111} interface.<br />

Figure 4.10 shows the length L of the projected (111) interface of dots with<br />

different height H. Only dots with highly symmetric shape of sample series #11 and<br />

#28 were used, which have almost the same height and width. The cross sectional<br />

images of these samples were evaluated. The measurements show that the length of<br />

the {111} interface is almost independent of the size of the dot and varies in a range<br />

of 7 to 10 nm. An exception of this can be found if the interface falls together with a


Figure 4.10: Length of the {111} interface as a function of the height of the dots. Only<br />

dots with the same height and width were used for the measurements. The solid line<br />

indicates a ratio {111} length/height of 0.61.<br />

stacking fault (also along [111] direction) of the CdTe crystal. For the small dots of<br />

sample #28 (originally an 1 nm layer PbTe) this leads to a disappearance of the other<br />

interfaces which can be see on the small dot in the cross sectional image of figure<br />

4.8.The projected length of the {111} interface must be smaller the 7nm for dots with<br />

a height smaller than about 12nm. For the acute angle of the small dots (α=70.5°,<br />

angle between a (111) and ( 1 11)<br />

interface), the ratio of the length to the height is<br />

−1<br />

( 2cos(<br />

α ) = 0.<br />

61<br />

L =<br />

. The solid line in figure 4.10 is a straight line with this<br />

H 2<br />

slope. For dots, which only build up {111} interfaces, the data points of the (111)-<br />

length should lie on the solid line. Also the {111} interface length of dots with all<br />

three interface kinds should have the same behaviour as a function of the height if all<br />

three interface of the dot increase uniformly. The line should be parallel to the solid<br />

line in fig. 4.10. The measurement show this behaviour for small dots with<br />

disappearing {110} and {001} interfaces and shrinking {111} interface. Larger dots<br />

show a relatively constant projection length of the {111} interface.<br />

Figure 4.11 shows two dots with different shapes and sizes. The left dot from series<br />

#28 is a small one with extended {111} interfaces relative to the {001} and {110}<br />

41


42<br />

areas. The sketch on top shows the contours of the dot with indicated interfaces. The<br />

same can be seen for a larger, highly symmetric dot from series #11. For these dots all<br />

projected interface have almost the same length. It can be seen that the (111) interface<br />

for both dots have approximately the same length of about 10 nm.<br />

Figure 4.11: The TEM images shows a small and a bigger QD recorded along the [ 110]<br />

zone axis as bright field images. The top of the image shows the according cubo-<br />

octahedrons with different interface sizes.<br />

Figure 4.12 shows three possible equilibrium shapes of quantum dots. The shapes<br />

were calculated for different interface energies using the Wulff construction. These<br />

calculations and the calculations of the interface energies were performed by R.<br />

Leitsmann et al., <strong>Institut</strong> <strong>für</strong> Festkörpertherorie und -optik, Jena [42]. This group in<br />

Jena investigates the interface problem between PbTe and CdTe theoretically. They<br />

apply density functional theory (DFT) as implemented in the Vienna ab-initio<br />

simulation package (VASP). More detailed description of their work can be found in<br />

the next chapter about the interface characterization. One finding of these calculations<br />

was that the interface energies for the three experimentally observed interfaces classes<br />

are almost the same. They found for the {110}, {100} and {111} interfaces energies<br />

of 0.20, 0.23 and 0.19 J/m². It is important to say that the energies of the polar<br />

interfaces {100} and {111} contain a dipole energy correction term of the same order<br />

as the uncorrected value. Three shapes can be seen in Figure 4.12. The outer right dot<br />

is calculated with the aforementioned corrected energies of the DFT calculation. The<br />

dot in the middle is one with equal interface energies of 0.20 J/m² for all three<br />

interfaces. The dot on the left hand side has again interface energies of 0.20 and 0.23<br />

J/m² for the {110} and {100} interfaces. The interface energy of the {111} interfaces<br />

is 0.22 J/m², which is within the estimated error bar of the calculated value of 0.19<br />

J/m².


Figure 4.12: Different shapes of cubo-octahedrons calculated with the Wulff<br />

construction with different interface energies [42]. The {001}, {110} and {111}<br />

interfaces are indicated as red, green and blue. The directions for the projected shapes are<br />

indicated. A projection along the [ 110]<br />

axis contains parallel interfaces of all three<br />

kinds.<br />

If the different shapes in Fig. 4.12 are scaled to a projected {111} length of 10nm,<br />

then the dot in the middle corresponds to a dot height of about 25 nm. This is the<br />

mean size of the dots in series #11. The right-hand dot represents one with a smaller<br />

height of 20 nm, which can be found in series #28. One important conclusion of the<br />

left dot is the increased difficulty to image the (111) interface. The first difficulty has<br />

its origin in the triangular shape of the interface. This means that the length of the<br />

transmitted interface changes as one focus through this interface. The second point<br />

concerns large dots, such as the one on the left side in Fig. 4.12. In the projection the<br />

{111} interface is too small to connect the {110} with the {001} interface. This edge<br />

consist of the projected {111} interface, and the edge between two {110} interfaces.<br />

The angle of these two parts is almost the same. That only a part of the edge is the<br />

projected {111} interface, is an additional difficulty for imaging.<br />

The interface energies of a real dot are constants and should not change with the<br />

size of the dot. If one calculates the equilibrium shape of the dots one obtains the<br />

same shape for different sizes. The edges and corners between the interfaces are not<br />

considered in the Wulff construction. A system with dominating interface energies<br />

forms an equilibrium shape, which is independent of the dot size. The real system<br />

shows variation of the dot shape depending on the size. This is a hint for the influence<br />

of the edge energies on the equilibrium shape. The behaviour of the real system with<br />

the relatively stable size of the {111} interface leads to a constant edge length<br />

between the {111} and {110} interfaces and shrinking edge between the {110} and<br />

{001} interfaces.<br />

43


44<br />

Chapter 5<br />

Interpretation of HRTEM Images<br />

As shown in chapter 3, HRTEM images are not real images of the structure. The<br />

formed images depend strongly on the transfer function as a function of Δf and the<br />

thickness t via the intensity of a diffracted beam (eqn. 3.13). The first cross-over of<br />

the transfer function changes the sign of the phase contrast. For conditions away from<br />

the Scherzer defocus, the first cross-over moves to smaller values of reciprocal<br />

distances. If the investigated structures are in the range of the resolution limit, small<br />

lattice frequencies (large distances in real space) and large lattice frequencies (small<br />

lattice distances in real space) are mixed up with positive and negative phase contrast.<br />

The intensity of a diffracted beam is proportional sin 2 (πξg -1 t), which leads to a<br />

maximum intensity of a diffraction beam with wave vector g at a multiple of ξg/2. The<br />

same behaviour is then also valid for the interference of diffracted beams, which leads<br />

to the lattice fringes. This strong dependence on Δf and the thickness t makes it<br />

difficult to interpret HRTEM Images.<br />

It is necessary to perform HR map simulations to interpret HRTEM images.<br />

Firstly, one has to make an assumption of the atomic structure of the sample. The next<br />

step is the construction of a three dimensional model of the sample structure which is<br />

then used to simulate HR maps. HR maps are functions of the defocus and thickness<br />

of the samples. Other parameters, which determine the quality of the maps, are, e.g.,<br />

the microscope parameters, astigmatism, noise, etc. These HR maps are compared<br />

with the real HRTEM images, and if they do not fit, the model is changed and the<br />

procedure is repeated until the HR map resembles the real HRTEM image.<br />

5.1 JEMS Simulation Package<br />

The JEMS (Java based Electron Microscopy Software) program is a software<br />

package that allows simulation of the diffraction and imaging of a TEM with a<br />

standard PC. It was developed by Prof. Pierre Stadelmann from the CIME-EPFL


(Centre Interdisciplinare de Microscopie Electronique-Ecole Polytechnique Fédérale<br />

de Lausanne) in Switzerland. A free version for students and information about this<br />

program can be found online at http://cimewww.epfl.ch/people/stadelmann/.<br />

The JEMS program has several helpful functions for electron microscopy. The<br />

crystal menu allows building up the crystal cells necessary for the simulations. The<br />

space and point group of the crystal, its basis and the lattice constant is entered to<br />

build up the unit cells of simple crystals. More complex crystal cells with a low<br />

symmetry are constructed by entering each atom position. The crystal files are saved<br />

in the txt-format as lists of the atomic positions. That allows easy conversion of<br />

calculated atomic positions to JEMS crystal files.<br />

The microscope performance is defined by entering the following parameters:<br />

• Acceleration voltage<br />

• Chromatic aberration coefficient<br />

• Spherical aberration coefficient<br />

• 5 th order spherical aberration coefficient<br />

• Gun energy spread<br />

Also adjustable are the parameters which are changed during the use of the<br />

microscope, such as the beam tilt, size and position of the apertures, the camera<br />

length, etc.<br />

Different functions can be applied to the crystal files and the parameters of the<br />

microscope, such as the generation of a diffraction pattern with or without Kikuchi<br />

lines, Convergent Beam Diffraction (CBED) patterns, or the transferfunctions of the<br />

microscope for a distinct value of the defocus. Also a function for the simulation of<br />

the propagating electron beam is then available. There are two common approaches<br />

for TEM simulations to calculate the propagating electron beam, and the JEMS<br />

program package is able to perform simulations with both methods. The bloch wave<br />

approach is suitable for simulations where the calculation must include High Order<br />

Laue reflections like in the case of CBED patterns. For HRTEM map simulations of<br />

lattice defects, interfaces or large crystal unit cells the multi-slice approach is used.<br />

The simulations concerning the PbTe/CdTe interfaces were done with the multi-slice<br />

approach. A detailed description of all functions of the JEMS program can be found in<br />

the online manual of the program [43].<br />

45


46<br />

5.1.1 Multi-Slice Approach<br />

The calculation of the DP and HR maps is based on the dynamical theory of elastic<br />

electron diffraction at small angles. It uses several approximations, in particular the<br />

small angle approximation. This approximation is satisfied if the energy of the<br />

electrons is lager than about 50 keV, but depends also on the crystal. A full dynamic<br />

calculation would need too much computer power. The multi-slice approach uses the<br />

fact that the crystal potential is not z-dependant and that scattering at these high<br />

energies is strongly forward peaked. The crystal is sliced in layers of thickness Δz,<br />

and the slice potential is replaced by a projected potential (eqn. 5.1).<br />

The calculation starts to calculate the propagating beam through the first slice,<br />

followed by propagation through vacuum to meet the next slice. This step takes also<br />

the microscope behaviour into account. This is repeated until the desired thickness is<br />

reached. In the real space approach, which is used to save computer time, it can be<br />

written as:<br />

Whereas P(r) is the propagator of the electron in free space and describes the<br />

microscope and is also called Fresnel propagator. q(r) is an exponential function of<br />

the projected potential and describes the specimen. It is called the phase grating or the<br />

phase object function. The knowledge that P(r) is strongly forward peaked is used to<br />

decrease the computation time.<br />

The real space approach was used for the JEMS calculations. The program is also<br />

able to display the Fresnel propagator, phase object function and the projected<br />

potential individually. More details of the dynamical theory of elastic electron<br />

diffraction with the approximations that are necessary for the bloch-wave or multi-<br />

slice approach can be found in [44]. For more detailed descriptions of the two<br />

methods see [32] and [33].<br />

V<br />

P<br />

z+<br />

τ<br />

1<br />

=<br />

τ ∫<br />

( r)<br />

V ( r;<br />

z′<br />

) dz<br />

z<br />

[ ψ ( ) ⊗ P<br />

( r)<br />

] q ( r)<br />

ψn+ 1 = n r n+<br />

1 n+<br />

1<br />

′<br />

(5.1)<br />

(5.2)


5.1.2 Simulation Procedure<br />

The standard procedure of image interpretation with HR-map simulations is done with<br />

a HR map which contains simulations for different values of defocus and specimen<br />

thickness. This can be done in the HR multi-slice dialog by choosing the starting<br />

focusing condition, the focus step size and the number of focus steps. The simulated<br />

sample thickness is controlled by the number of slice iterations. The parameters are<br />

the start number of iterations, the number of simulated image and the increment<br />

between these images. The value of the thickness is the number of iteration times the<br />

slice thickness which depends on the crystal model. For instance, the models for the<br />

simulation of bulk CdTe and bulk PbTe are simply the unit cells of these materials<br />

(see Fig. 5.1). The orientation of the cell can be chosen: For the PbTe/CdTe samples<br />

one looks along a direction for plan view specimens and along a <br />

direction for cross section specimens.<br />

Figure 5.1: The unit cells of CdTe and PbTe as models for the JEMS program. The DP<br />

of the materials shows the diffracted beams causing the lattice fringes. The 2D projection<br />

shows 2x2x2 unit cells along the [001] direction. The according HR simulations are<br />

direct maps of the crystal structure and show a positive contrast (dark areas correspond to<br />

the atomic positions). The images were simulated for the Scherzer defocus condition and<br />

a sample thickness of 4.5nm which are 15 PbTe (001) layer for PbTe or 15 Cd-Te bi-<br />

layer for CdTe.<br />

47


48<br />

Firstly, the program calculates the transmitted electron beam for the start number<br />

of slices. With that and the transfer functions of the different values of defocus the<br />

image maps are calculated. This is repeated until the reached iteration number equals<br />

the multiple of the iteration step number. The outcome is then an array of HR images.<br />

Such HR-maps can be seen in Figure 5.2 or 5.4. The image in the left bottom corner<br />

of the HR-map is the simulation for the thinnest sample and the start defocus, the<br />

image in the right top is for the thickest specimen and the end defocus. For the<br />

interpretation of HRTEM images it is also important to know, how the projection of<br />

the lattice is related to the pattern of the lattice fringes. The JEMS program allows<br />

displaying of the atomic positions of the used model in the HR map. The multi-slice<br />

menu point can also display the parameters which are used for the simulations, such<br />

as the projected potential and the phase object function.<br />

Figure 5.2: Comparison of CdTe and PbTe HR maps. The images show either positive<br />

or negative phase contrast. One reason for the strong variation for PbTe is the small<br />

spacing of the (110) lattice planes. This distance is comparable with the resolution limit<br />

of the microscope and therefore strongly sensitive to the defocus. The phase contrast<br />

switches several times. The thicknesses of 3.2, 5.5 and 7.8 nm correspond to 15, 25, 35<br />

layers (PbTe) or bi-layers (CdTe) of (110) lattice planes.


The DP of the materials is very helpful because it is directly correlated to the real<br />

image via the fourier transform. It helps to interpret the real image over the indices of<br />

the diffraction spots. For the plan view (one look along a zone axis) of CdTe<br />

the {002} reflection spots are forbidden, and the {022} spots are strong. The spacing<br />

between the (022) lattice planes are 0.23nm and near the practical resolution of the<br />

JEOL FasTEM 2011. For the rs structure of PbTe the {002} reflexes are the strongest.<br />

The spacing between the (002) planes is 0.46nm. These lattice planes and their<br />

spacing are then the dominant structures in the HR simulation for thin samples at the<br />

Scherzer defocus (see Fig. 5.1).<br />

For the [110] oriented materials the DP looks differently (simulation can be seen in<br />

Fig 5.3). The CdTe DP shows the dominating {111} spots of this material. The first<br />

two strong spots for PbTe are the {002} and the {220} reflexes. The lattice distance of<br />

the {111} planes is 0.37nm, i.e. significantly larger than the resolution limit of<br />

0.23nm. Also far away from the limit is the lattice spacing of the {002} reflex with<br />

0.46nm. However, the spacing of the {220} planes is with 0.23nm again near the<br />

practical resolution limit and in reciprocal space near the first cross over of the<br />

transfer function. This leads to strong variation of the PbTe pattern in the HR-maps as<br />

a function of the defocus, as shown in Fig. 5.2.<br />

Figure 5.3: Simulated diffraction patter for CdTe and PbTe oriented along the [110]<br />

zone axis. The patterns have the same structure but the intensities of the spots, which is<br />

indicated by the diameters of the circles, vary.<br />

Because the PbTe lattice fringes show much faster variation in the sign of the<br />

phase than the CdTe contrast, it could appear for CdTe/PbTe interface that a TEM<br />

49


50<br />

image shows opposite contrast for the two materials. Figure 5.4 shows the small range<br />

where PbTe shows positive phase contrast and can be directly correlated to the lattice<br />

structure. As can be seen in Figure 5.2, the pattern of CdTe is less sensitive to the<br />

defocus value and can be correlated over a larger range directly to the lattice structure.<br />

In order to know during the work at the microscope in which region of defocus one is<br />

working it is necessary to perform focus series to be sure to record HRTEM images at<br />

the correct defocus value.<br />

Figure 5.4: HR map of [110] PbTe at ideal defocus condition (Scherzer defocus -61nm).<br />

The image at Δf = -64nm and t = 1.4nm shows already a differing pattern, which does not<br />

resemble the lattice structure. The thicknesses of 0.5, 0.9 and 1.4 correspond to 3, 5 and 7<br />

(110) PbTe layers<br />

5.2 Image simulation of the CdTe/PbTe interfaces<br />

Different kinds of models were used for the interface simulation, with the largest<br />

model consisting of more than 270 atoms. This large model contains all three relevant<br />

interfaces and was used in the beginning to get a rough overview of how the interfaces<br />

should look in the HRTEM images. The {111} interfaces are hard to image, due to the<br />

size and the shape of these interfaces. A condition for high quality interface imaging<br />

is the orientation of the interface. It must be parallel to the electron beam and it should<br />

be a continuous structure throughout the specimen. It is hard to find a continuous


interface through the whole specimen because the (111) interfaces are small. The<br />

focused length of the interface changes as one changes the defocus because the<br />

interface is triangular shaped. This complicates the situation additionally. Only few<br />

high quality HRTEM images could be recorded and so no further simulations of these<br />

interfaces were performed. For the other two interface classes, {100} and {110},<br />

single models which contain only the interface of interest were used. The program<br />

MATLAB was used to calculate the atomic positions of the largest model. The atomic<br />

positions of the two single interface models were entered manually.<br />

The line-of-sight for cross sectional specimens is a zone axis of the crystal.<br />

This direction is favourable because all three interfaces of the Quantum dots are<br />

parallel to this direction and can be characterised. For this view the atomic rows<br />

parallel to a direction consist for both kinds of material of only one kind of<br />

atoms. This situation is really helpful for interface characterisation. To simulate HR-<br />

maps for these directions with the multi-slice approach it is necessary that the models<br />

have a layer structure where the atomic planes are perpendicular to the <br />

directions. The individual interface models and the according simulation are treated in<br />

the next chapter together with a comparison with the HRTEM images.<br />

51


52<br />

Chapter 6<br />

Interface characterization<br />

6.1 Available Samples<br />

6.1.1 Specimens for HRTEM imaging.<br />

High-resolution interface characterisation requires TEM specimens of very high<br />

quality. The specimen must be a few nanometers thick to get well interpretable<br />

HRTEM images, as shown by the JEMS simulation. Most of the HRTEM imaging<br />

was performed on sample series #11 for a simple reason: Series #11 was originally a<br />

double quantum well, which was transformed by annealing into a double QD-row<br />

sample. Cross sectional specimens were used for the interface characterisation<br />

because in this view all three interfaces can be found parallel to the electron beam.<br />

After preparation, the specimen is a disk with a dimple in the middle. Cross sectional<br />

specimens are constructed of stripes of sample material. In the middle of the dimple is<br />

the adhesive joint of the material stripes which is sputtered away in the thinnest area.<br />

The material is also thinned at this point of the specimen and a curved wedge<br />

develops from the straight edge. Only in a range of about 50 nm away from the edge<br />

the specimen is thin enough for HRTEM imaging. The wedge is steeper if it is farther<br />

away from the former edge, which decreases the limit of 50 nm. As was shows in<br />

chapter 5, this thickness restriction can be even narrower if the image should be easily<br />

interpretable. A double row increases the chance to meet this condition in the sample.<br />

A sketch of a schematic specimen edge is shown in Figure 6.1.


Figure 6.1: A schematic sketch of the specimen edge. At the thinnest area a curved edge<br />

develops upon ion milling. The dotted lines indicated the double QDs row. The hatched<br />

area indicates where good conditions for HRTEM imaging can be found. A double row<br />

doubles the chance for meeting the HRTEM conditions in comparison with a single row.<br />

6.1.2 Beam Damage<br />

Another problem of the HR investigations is beam damage of the sample material. For<br />

HR imaging the TEM is run in a high magnification mode. The electron beam is<br />

strongly focused to get sufficient intensity. The transferred energy to the field of<br />

interest is large enough to produce beam damage of the sample material. The<br />

investigations show that, depending on the thickness of the sample, the CdTe host<br />

material starts to dissolves after about 10 to 20 min of illumination with a 200keV<br />

beam at the intensity needed for the HR TEM images. This makes it necessary to<br />

work fast and to align the microscope sufficiently far away form the area of interest.<br />

Figure 6.2 shows the effect of a strongly focused beam on the CdTe/PbTe material<br />

system. The damage from image (a) to (b) happens during the alignment of the<br />

microscope for HR imaging at the highest magnification, which takes about 20 min.<br />

Image (c) is recorded after further 20 min and 10 min waiting time without<br />

illumination. The thickness of the region near the specimen edge is in the order of the<br />

free mean path of the electrons in CdTe or even thinner, thus less affected by damage.<br />

53


54<br />

Figure 6.2: Different steps of beam damage: (a) shows the original PbTe Dot at the<br />

beginning of illumination. (b) Beam damage is visible after ~20min illumination.<br />

Stacked material with different lattice constants (temperature differences) of the same<br />

lattice structure show so called Moiré Patterns, which originate from the overlap of<br />

different sized periodic structures. (c) QD after further 20min: strong beam damages can<br />

be seen containing even areas where the material was completely evaporated by the<br />

electron beam.<br />

6.2 HRTEM investigation of PbTe nanocrystals<br />

An aim of the HRTEM investigations was to learn more about the correlation between<br />

the PbTe-rs lattice and the CdTe-zb lattice. First use of HRTEM images was the<br />

investigation of the lattice structure of the dots to get a raw overview. The more<br />

sophisticated part was the attempt of detailed interface characterisation.<br />

6.2.1 HRTEM images of single QDs<br />

The dark and bright field images, as shown in the previous chapters, reveal sharp<br />

interfaces between the PbTe and CdTe material. The question was, if this is also the<br />

case on the atomic scale. The analyses of the HRTEM images showed that this is


Figure 6.3: Two HRTEM examples of dots in sample series #11. (a) The dot shows<br />

coherent interface between the PbTe and CdTe lattices, which are almost defect free. (b)<br />

Two dots can be seen to the left (small dot) and the right (large dot) of a stacking fault.<br />

This sample is too thick for high quality HRTEM images and gives rise to distinct mass –<br />

thickness contrast.<br />

Figure 6.4: Dot recorded at an annealed sample of series #11. The image is stitched<br />

together of four HRTEM images. The growth direction is indicated and correlates with<br />

the direction of decreasing wedge thickness. The lattice fringes of PbTe are well oriented<br />

with respect to the CdTe fringes. The number of PbTe lattice planes is constant on both<br />

(110) interfaces, indicating the absence of dislocations.<br />

55


56<br />

indeed the case for the {001} and {110} interfaces. Only few mono-atomic height<br />

steps can be found at the {001} interfaces at the bottom of the PbTe SQW, which is<br />

created during MBE growth. Practically no steps are found at the {110} and {001}<br />

interface, which are created during the annealing process. Figure 6.3 shows some<br />

HRTEM examples of dots with these interfaces. The {001} and {110} interfaces<br />

show perfectly coherent interfaces between CdTe and PbTe. The {111} interfaces<br />

show different appearances. This can be explained by the triangle shape of these<br />

interfaces. If the conditions for HRTEM images are fulfilled no mass-thickness<br />

contrast should be visible. This is partly fulfilled for the top part of Fig. 6.3a.<br />

Another question is the correlation between the two different lattice structures.<br />

Figure 6.4 shows a HRTEM images stitched together of four individual images. The<br />

resulting dot shows well-ordered PbTe lattice planes with almost no lattice defects.<br />

The right hand (110) interface of this dot was used for the interface characterisation<br />

below (part 6.2.4). In particular, counts of the PbTe lattice planes showed that PbTe<br />

and CdTe lattice fringes are well aligned and there are no discrepancies in the number<br />

of lattice planes between the PbTe and CdTe, i.e. no dislocations.<br />

6.2.2 Total energy calculations<br />

Total energy calculations were carried out at the <strong>Institut</strong> <strong>für</strong> Festkörpertheorie und –<br />

optik, Jena, Germany [42] in a collaboration funded by the Spezialforschungsbereich<br />

IRoN of the FWF, Vienna. They applied density functional theory (DFT) in local-<br />

(spin) density approximation (L(S)DA). The calculations were carried out with the<br />

Vienna ab-initio simulation package (VASP). The interface energies and atomic<br />

displacements were calculated using the repeated slab approximation. For the<br />

construction of the slabs an averaged, theoretical lattice constant of 6.41 nm was used.<br />

The slabs, i.e. 3d models of the interface structure, were constructed like the JEMS<br />

crystal files. A continuous Te host matrix was assumed as a starting condition. This<br />

matrix was filled up with a Cd fcc-sublattice displaced by (¼, ¼, ¼) for the CdTe part<br />

and a Pb fcc-sublattice displaced by (½, ½, ½) for PbTe. This leads to the two<br />

different {001} interface spacings. The total free energy for this slab system is smaller<br />

than for {001} interfaces with a displaced Te matrix and equal interface separations.<br />

The slabs should be large enough, electrostatic neutral and stoichiometric. The slab<br />

system for the (111) interface consists of a multiple of tree bi-layers of PbTe and<br />

CdTe respectively. The (001) interface slab consists of four layer (two neutral bi-<br />

layer) CdTe and two neutral layer PbTe. For the (110) interface two neutral layers of


PbTe and CdTe, respectively, are combined. Ionic relaxation is allowed during the<br />

calculation until the interatomic force falls below a pre-defined cut-off.<br />

Displacements occur in two directions, perpendicular and parallel to the respective<br />

interface. For {111} and {001} interfaces, only displacements perpendicular to the<br />

interfaces were found. This is obvious for the polar {001} interface, which is<br />

symmetric around the interface normal. The different terminations lead to different<br />

atomic relaxation. The effects are stronger at the Cd terminated interface, where the<br />

spacing is smaller. The {111} interface shows similar behaviour as the {001}<br />

interface. The polar interfaces show a so-called rumpling effect for the PbTe, which is<br />

also known for bulk PbTe surfaces [45].<br />

The main finding of the {110} simulations is a displacement of the two crystal<br />

halves against each other, i.e. a significant modification of the original starting<br />

condition of a continuous Te matrix. The results lead to a displacement of the Te<br />

matrix of about 0.04nm across the interface. Figure 6.5b shows the calculated atomic<br />

displacement in comparison with the simple starting model in Fig.6.5a of a (110)<br />

interface between bulk PbTe and CdTe with a continuous Te matrix. In Fig 6.5b the<br />

atomic positions of the undisturbed model are indicated as black crosses. The clear<br />

difference is the shift of the two crystal halves against each other. The differences of<br />

Figure 6.5: (a) shows the projected model of the (110) interface, assuming a continuous<br />

Te matrix. (b) shows the atomic positions after total energy minimisation. This model<br />

results in a shift of the two crystal halves against each other of 0.04nm. The black crosses<br />

mark the atomic positions of (a).<br />

57


58<br />

the displacements along the [001] direction and the displacements along the [110]<br />

direction are hardly visible in this picture. The detailed displacements are discussed in<br />

part 6.2.4, where the calculated values are compared with the displacements extracted<br />

from the HRTEM images. The calculated values are also used to construct a JEMS<br />

model for HR map simulations. The results can also be found in part 6.2.4.<br />

6.2.3 The (001) interface<br />

Few is known about the interface between rs and zb crystals. Two important points<br />

were interesting to be investigated at this interface. The first concerns the two<br />

different kinds of interfaces depending on the interface termination. One interface is<br />

determined by the substrate termination before PbTe growth. The opposite interface<br />

above the PbTe SQW should have the other kind of termination, according to atomic<br />

number conservation. It is not obvious that this is the behaviour of the real system, in<br />

particular for annealed samples, where the PbTe crystal part is subjected to strong<br />

shape transitions and the (001) interface between the PbTe and the CdTe capping<br />

layer is newly built. The substrate termination could be destroyed and both (001)<br />

interfaces could contain both kinds of termination. It would be also possible that one<br />

kind of interface termination is preferred and the stoichiometry is fulfilled by excess<br />

atoms being segregated and causing lattice defects. There is one study of a rs-zb (001)<br />

interface in a ErAs(rs)/GaAs(zb) material system, where only one kind of termination<br />

was treated[46]. As will be shown, both kinds of interfaces can be seen in our<br />

experiments, which can be distinguish by the spacing at the interface. SQW and dots<br />

show both interface kinds on the opposite (001) interfaces and no intermixing of the<br />

termination occurs. Both interfaces, the one that is terminated during the MBE growth<br />

and the one that is newly established during the annealing process, show a low defect<br />

density. For the interpretation of the HRTEM images JEMS HR-maps were used.<br />

PbTe and CdTe have different rotational symmetry around the interface normal which<br />

complicates the correlation of the HRTEM image to a lattice structures.<br />

The second point is the rumpling effect at the interface. The DFT calculation<br />

predicted atomic displacements at the interface. This interface rumpling possibly<br />

should also be seen in the HRTEM maps or at least the experiments should show that<br />

there are not other, larger displacements.<br />

(001) interface modelling for the HR-simulations<br />

Both lattices, the rs and the zb lattice, consist of two face centred cubic ( fcc ) lattice<br />

of which one in each lattice consists of Te atoms. We therefore assume for the first


simulation a continuous Te matrix. The two crystal halves are then filled up with the<br />

respective other fcc lattice (Cd or Pb). The difference is only their displacement with<br />

respect to the Te matrix. It is ( ½, ½, ½) for the PbTe rs structure and ( ¼, ¼, ¼) for<br />

the CdTe zb lattice. As will be shown in Chapter 6, it was necessary for the (110)<br />

interface to change the starting condition of a continuous Te matrix. As an average<br />

lattice constant 0.6471 nm was used.<br />

Figure 6.6: (a) Cd terminated CdTe/PbTe (001) interface; (b) Te terminated CdTe/PbTe<br />

(001) interface. The two possible (001) interfaces lead to 4 different images for a<br />

projection along the directions. If one looks along the [ 11<br />

0]<br />

direction the top two<br />

Cd Atoms are in a row, for a view along [ 1 10]<br />

the two atoms at the bottom are in a row.<br />

The appearance of PbTe in the projection does not change under this rotation of 90°. The<br />

used JEMS models have the same structure but are extended along the [001] direction on<br />

the PbTe and CdTe side.<br />

For the (110) interface this leads to only one model, but the polar (001) interfaces<br />

can be either Te or Cd terminated, which leads to two different models (see Fig. 6.6).<br />

Aside of that, the CdTe and the PbTe crystal halves have different rotational<br />

symmetries around the [001] axis. CdTe has a 180° symmetry, PbTe has a 90°<br />

rotational symmetry. This led to two different appearances of the polar interface under<br />

a 90° rotation. These two different directions are not distinguished during the<br />

preparation process and it is necessary to perform two simulations and to allocate this<br />

direction in combination with the grow direction.<br />

59


60<br />

Figure 6.7 shows the four different projections of the Cd and Te terminated<br />

interface. The HR image simulation shows clear differences between the Cd and Te<br />

terminated interfaces. A clear separation can be seen for the Te terminated interface<br />

between the PbTe and the CdTe crystal halves. The orientation determines whether<br />

the CdTe pattern starts with a single or double row of atoms. It is the same behaviour<br />

for the Cd terminated interface, but the distance at the interface is smaller so that the<br />

first CdTe pattern is connected to the PbTe.<br />

Figure 6.7: The four possible appearances of the (001) interface. The projection shows<br />

three unit cells of the JEMS models projected along the directions indicated in Fig. 6.6.<br />

The HR images are simulated for the Scherzer defocus of -61nm and for t = 1.4 nm (7<br />

(110) lattice planes for PbTe).<br />

These four different patterns should be found in a HRTEM image. In an annealed<br />

sample the CdTe substrate is connected with the CdTe capping layer and no phase<br />

discontinuity should appear in the host material of the dots. Two different appearances<br />

of opposite {001} interfaces are possible with the assumption that both kinds of<br />

termination can be found at one dot on opposite {001} interfaces. Figure 6.8 shows<br />

the HR simulation of two opposite (001) interface with respect to the crystal<br />

orientation. Figure 6.8a is a combination of the Cd terminated interface with a<br />

projection axis [ 11<br />

0]<br />

and the Te terminated interface with a projection axis [ 1 10]<br />

with respect to Fig. 6.7. The other two interfaces of Fig.6.7 result in Fig. 6.8b.


Figure 6.8: HR simulation of the (001) interface of both crystal directions. Simulation<br />

parameters: defocus -61nm, thickness 1.4 nm. (a) shows HR simulations of a Cd and Te<br />

terminated (001) interfaces arranged how they should be found at opposite (001)<br />

interfaces at a PbTe SQW or QD. (a) shows the same for a 90° rotated crystal orientation.<br />

HRTEM images of the (001) interface<br />

Figure 6.9 shows a HRTEM image recorded on a series #5, as-grown sample. It<br />

was originally a 5 nm SQW, but it already started breaking apart. The layer thickness,<br />

which can be seen in the left-hand image, is about 10nm and the SQW fragment is<br />

terminated to the left with two {111} interfaces. The zoomed-in sections in this figure<br />

show Fourier filtered images of the two (001) interfaces. The two images are recorded<br />

at different specimen thickness due to the wedged shape of the specimen. There is<br />

also some uncorrected astigmatism left, which can especially be seen in the bottom<br />

section by the preferential direction of the CdTe lattice fringes. Nevertheless, the<br />

section shows the different behaviour of the two kinds of termination. The section at<br />

the top shows the interface created during the annealing step and has larger lattice<br />

spacing. The PbTe and CdTe part are well separated. The section at the bottom shows<br />

the Cd terminated interface, which was defined by growth. The interface spacing is<br />

smaller and the nominal interface is not as clearly visible as for the Te termination.<br />

61


62<br />

Figure 6.9: HRTEM image of an as-grown sample (series #5). Both (001) interfaces can<br />

be seen in the image. This sample was Cd terminated before PbTe growth. The two<br />

zoomed-in images show the Fourier-filtered interfaces. One can see the difference of the<br />

two interfaces, especially at the interface pattern. This pattern depends on the interface<br />

separation, which is determined by the kind of termination. The two patterns of bulk<br />

PbTe and CdTe are better separated for the Te terminated interface where the larger<br />

lattice spacing can be found.<br />

Figure 6.10 shows this behaviour for two images from sample series #5. The two<br />

interfaces were recorded at almost the same position and ideal defocusing condition at<br />

a SWQ. The projected JEMS model of the two different kinds of (001) interfaces on<br />

the left hand side of this figure shows the atomic position at the interface. The<br />

terminations determine the lattice plane distances at the interfaces. The Cd terminated<br />

interface shows a lattice plane spacing of ¼ of the lattice constant. The Te terminated<br />

interface has a spacing of ½ a. The two different terminations can be easily<br />

distinguished, because ¼ a = 0.16nm is below the practical resolution limit of the<br />

microscope. The nominal interface sits in the middle of the ¼ a distance. Thus,<br />

HRTEM images of a Cd terminated interface show dark areas across the nominal<br />

interface. HRTEM images of the Te terminated interface show well separated lattice<br />

fringes of bulk PbTe and CdTe at the interface.


Figure 6.10: The images on the right-hand side show two HRTEM images recorded at<br />

ideal defocus condition along the [ 110]<br />

zone axis. The images were taken on an as-<br />

grown cross sectional specimen of sample #5. They were imaged at almost the same<br />

position on the SQW. (a) shows the (001) interface with Cd termination. On the opposite<br />

side a Te terminated interface is found (b). The MBE growth direction and the supposed<br />

atomic positions are indicated. The left image shows the projected crystal structure with<br />

indications of the interface terminations and the interface spacings. The simulation for<br />

the interface can be seen in fig. 6.8b.<br />

The atomic displacements at the interface are too small to be measured for a<br />

comparison with the calculations. The total energy calculations find a displacement of<br />

the Pb atom of the first lattice plane of PbTe near the interface in direction of the<br />

PbTe, which is comparable to the so called rumpling effect [45]. This was a finding<br />

for both kinds of termination. Additional to this rumpling effect, displacements are<br />

found for the Cd terminated interface with the smaller interface spacing. All atoms of<br />

CdTe are displaced slightly into the CdTe halve and the first PbTe lattice plane is<br />

slightly changed into a zigzag chain like the neutral Cd-Te bi-layer. This can be<br />

interpreted as the domination of the CdTe zb lattice at the Cd terminated interface.<br />

Small variations of the PbTe lattice fringes can be seen in Fig. 6.10a, but this variation<br />

is too small to be evaluated from the HRTEM image. As was shown in Chapter 4, the<br />

PbTe HR image is strongly thickness dependent. The variation resulting from the<br />

displacement can easily be mixed up with the strong thickness variations of the PbTe<br />

63


64<br />

pattern. This is taken into account especially for the (001) interface because the<br />

interface normal is parallel to the direction of increasing thickness of the specimen<br />

wedge. Given that the Cd terminated interface can be seen as being dominated by the<br />

zb-lattice, the Te terminated interface can be interpreted as being dominated by the<br />

PbTe rs structure. The lattice spacing of this interface is ½ a, which is also the lattice<br />

plane spacing of the PbTe lattice planes. HR simulations of this line of sight can be<br />

seen in Fig. 6.8b, or in Appendix B, where HR maps can be found.<br />

The (001) interfaces have two different appearances as shown in Figure 6.8. The<br />

characteristics of the HRTEM lattice fringes of Figure 6.10 are well reproduced with<br />

the HR-simulation of 6.8b. The single row of Te atoms can be seen in the top HR<br />

simulation for the Te termination. This row is followed by a bright region parallel to<br />

the interface representing the interface spacing. This bright, interconnected lattice<br />

fringe parallel to the interface cannot be found at the Cd terminated interface. Figure<br />

6.11 shows the two (001) interfaces recorded along a line-of-sight rotated by 90° with<br />

respect to the HRTEM image of Figure 6.10. The fringes of these images can be<br />

identified as the simulated images of Figure 6.8a. Further HR maps can be found in<br />

Appendix B. The different terminations can again be seen easily by the lattice plane<br />

separation at the interface.<br />

Figure 6.11: The two (001) interface in a line of sight 90° rotated with respect to Figure<br />

6.10. They were recorded on sample #11 at almost ideal defocus conditions at the dot<br />

that can be seen in Figure 6.2. The respective HR simulations are shown in figure 6.8a, or<br />

can be found in Appendix B.


6.2.4 The (110) interface<br />

The main effect to be seen at a {110} interfaces is the shift of the two crystal<br />

halves against each others. A shift of the lattice fringes occurs, if the thickness<br />

changes (see eqn. 3.16). Thus it is important to know that the PbTe part and the CdTe<br />

part of the specimen have the same thickness. Differently thick material regions could<br />

be encountered, if the thinning process by Ar - ion sputtering is differently efficient<br />

for CdTe and PbTe. But no indication for such behaviour can be found in the<br />

specimens. The edge of the specimen is an almost continuous line. This is also the<br />

case, if the edge cuts through a dot, which means that the sputtering rate is widely<br />

independent for the two materials.<br />

(110) interface modelling for the HR-simulations<br />

The (110) interface is from the point of view of the model simpler than the (001)<br />

interface. The (110) interface is a non-polar one and only one model is necessary (see<br />

Fig. 6.12). Also, the rotational symmetry around [110] is the same for PbTe and<br />

CdTe. This leads to only one possible projection of this interface. The interface was<br />

constructed by putting together the (110) surfaces of CdTe and PbTe with a<br />

continuous Te matrix. Fig 6.13 shows a HR simulation of this interface. Under this<br />

Figure 6.12: Starting model of the (110) interface between PbTe and CdTe. The (110)<br />

surface is non-polar for both materials and also the rotational symmetry around the<br />

[ 11<br />

0]<br />

axis is the same. For this assumption of the interface the sp³ hybrid bond of a Cd<br />

atom meets two p³ orbitals of a Te atom. Rearrangement of the bonds leads to atomic<br />

displacements at the interface which were implemented in this model for the comparison<br />

with the HRTEM images.<br />

65


66<br />

assumption the sp³ hybrid bond of a Cd atom meets two p³ orbitals of a Te atom.<br />

Rearrangement of the bonds lead to atomic displacements at the interface, which were<br />

implemented in this model for the interpretation of the HRTEM images in Chapter 6<br />

to reproduce the interface of the experiment.<br />

Figure 6.13: Projection of the (110) interface model with a simulated HR pattern. The<br />

simulation parameters are Δf = -61nm t = 1.8 nm which correspond to 9 (110)–lattice<br />

planes of PbTe. The image shows a positive phase contrast and the atomic positions are<br />

indicated. The assumption of a continuous Te matrix can be clearly seen in the<br />

simulation by the continuous top edge of the dark PbTe rows into the CdTe part.<br />

HRTEM images of the (110) interface<br />

In Figure 6.14 two examples of (110) interfaces can be seen. Both images are<br />

recorded at almost ideal defocus conditions. Image (a) shows a dot at the specimen<br />

edge. The amorphous region is the adhesive from sample preparation. Image (b) was<br />

recorded at the same dot which is shown in Figure 6.2. Both images show no mass-<br />

thickness contrast or, in other words, the phase contrast dominates, which is an<br />

indication that the specimen is thin enough. This is a precondition for a direct<br />

correlation of the lattice fringes with the lattice structure. Especially in image (b) no<br />

substantial change of the lattice fringes can be seen along the [001] direction, which<br />

corresponds to increasing thickness of the specimen wedge. This is an indication of a<br />

wedge with a small angle. The interface normal is perpendicular to the wedge. So, no


Figure 6.14: Two example of the (110) interface recorded with high magnification at<br />

almost ideal defocusing condition. (a) is recorded on sample #30 at the edge of the<br />

specimen wedge. The bright area is the adhesive of sample preparation showing the<br />

pattern of an amorphous material. (b) is recorded on series #11. The interface is part of<br />

the dot shown in Figure 6.2. This image was used for the construction of the JEMS<br />

model. A Fourier filtered section of the interface can be seen in Figure 6.16.<br />

change of the lattice fringes should occur along this direction. This allows a<br />

straightforward interpretation of these images to measure the shift of the crystal<br />

halves.<br />

A straightforward comparison shows a phase shift of the PbTe lattice fringes with<br />

respect to the CdTe lattice fringes. It was not possible to simulate this phase with the<br />

continuous Te matrix model. A shift cannot occur with this model and comparisons<br />

with HRTEM images show pattern discrepancies of the lattice fringes at the interface.<br />

These discrepancies concern especially visible variations of the PbTe lattice fringes<br />

near the nominal interface. It can be seen as alternating bends of the lattice fringes<br />

along the [001] directions. This makes it necessary to adjust the JEMS model. The<br />

first attempt was the use of the calculated atomic displacement from the total energy<br />

calculation. Figure 6.15a and b shows examples of HR simulations and the projected<br />

lattice structures of these two models. The second model leads to HR simulations,<br />

which reproduce the shift of the crystal halves, but do not resemble the characteristics<br />

of the lattice fringes near the interface. Differences can be seen especially for the<br />

alternating bends of the PbTe lattice fringes near the interface.<br />

67


68<br />

Figure 6.15: Examples of HR simulations calculated with different JEMS interface<br />

structure models. (a) HR image of the model with the continuous Te matrix without<br />

atomic displacements. Simulation parameter: Δf=-61 nm, t=0.9nm; (b) Model with the<br />

theoretic atomic displacements of the total energy calculations. Δf=-60 nm, t=0.9nm; (c)<br />

Model constructed with atomic displacements extracted from a HRTEM image. Δf=-60<br />

nm, t=0.9nm;<br />

Figure 6.16: Comparison of a HRTEM image (a) with a HR simulation (b). The used<br />

model for the simulation contains atom displacements extracted from the HRTEM image.<br />

The model reproduces the shift of the crystal halves and the characteristics of the lattice<br />

fringes near the interface. The positions of the atomic species are indicated in the<br />

simulated image. Simulation parameters: Δf=-60 nm, t=0.9nm.


Figure 6.15c and 6.16b show examples of HR simulations with the model which<br />

resembles the real interface best. The approach for this new model was the correlation<br />

of the HRTEM image to the crystal structure. A section of image 6.14b can be seen in<br />

Figure 6.16a. The section is from the middle of the image to minimise the effects of<br />

remaining lens aberration after microscope alignment. The image was also Fourier<br />

filter to get rid of the noise and the mass-thickness contrast. The lattice fringes in both<br />

crystal halves show a positive phase contrast and the patterns of the bulk material at<br />

ideal defocus (Scherzer defocus). HRTEM images were recorded along the whole dot<br />

interface with the same ideal defocusing conditions. The difference of {001}<br />

interfaces regarding the termination and orientation allow a correlation of the lattice<br />

structure and the HRTEM image. The knowledge that the substrate was Cd terminated<br />

before PbTe growth allows the adaptation of the lattice structure. This is important,<br />

because for a non-continuous Te matrix it is important to know which orientation the<br />

Cd-Te dimer has with respect to the dark areas in the CdTe lattice fringes. In case of<br />

Figure 6.16a the tellurium atom rows sit at the top and the Cd atom rows at the bottom<br />

of the dark areas. All lattice fringe structures could be allocated to atomic positions of<br />

the interface structure with the knowledge of the crystal orientation. This makes it<br />

possible to measure the atomic positions in the HRTEM images. This was done with<br />

the program Gatan DigitalMicrograph [41]. Arrays of 8x8 atom positions were<br />

measured at the (110) interface of Figure 6.10b. For the PbTe lattice fringes each<br />

single atom is resolved, thus simply the centre of the dark areas were assigned to the<br />

atomic positions. For the CdTe lattice, where each dark area represents a Cd-Te<br />

dimer, the dark area was fitted by two circles. Each centre of the fitted circles were<br />

assigned to the atomic positions. As reference to the first JEMS model the last Cd-Te<br />

dimer, which is furthest away from the interface, was used. Their displacement was<br />

set to one halve of the 0.04 nm shift. With these values, a new model for the JEMS<br />

simulation was constructed. The most important change to the first JEMS model was<br />

the abandonment of the continuous Te matrix condition. The new model contains a<br />

shift of the CdTe and PbTe crystal halves as predicted by the DFT calculation. A<br />

simulated HRTEM image can be seen in figure 6.16b. The simulated image shows the<br />

same features at the interface as the HRTEM image. The model with indicated shifts<br />

can be seen Figure 6.17.<br />

69


70<br />

Figure 6.17: Projection of the model constructed with the measured displacements. The<br />

shift of 0.04 nm is clearly visible. The horizontal lines indicated the Te positions for a<br />

continuous Te matrix; the dashed line indicates the nominal interface. The origin of the<br />

difference in the HR simulation in comparison to the calculated values is the Te atom in<br />

the first lattice plane parallel to the interface. It has almost no displacement.<br />

The comparison of the calculated values and the measured ones can be seen in<br />

image 6.18. The displacements are sorted with respect to atomic species and<br />

displacement direction. The standard error of the mean of the measured atomic<br />

displacements is in the range of ±0.002 nm to ±0.015 nm for the individual measured<br />

values and is indicated as dashed black lines. The calculated displacements show good<br />

agreement with the measured ones. One difference of the calculated and measured<br />

displacements can be seen in the HR simulations and concerns the alternating bends<br />

of the PbTe lattice fringes near the interface. This visible discrepancy concerns the<br />

atom row along the [110] directions. In this row Cd and Te atoms are alternating. For<br />

the theoretical values the displacements of Te and Pb show the same tendency. In the<br />

simulated HR images these atomic rows show slight variations, but they seem to be<br />

parallel. The measured values for the Te atom in the first lattice plane parallel to the<br />

interface show almost no displacement, which leads to a bending of the rows with this<br />

atom. In the HR simulation this appears as a pattern where the bending of the atomic<br />

rows alter, depending on the atom species in the first PbTe lattice plane near the<br />

interface. The variation along the Te-Pb chains in PbTe can be interpreted as<br />

continuation of the zb structure into the rs structure. One electrostatically neutral<br />

CdTe bi-layer, orientated along [ 110]<br />

, can be described as a Cd-Te zigzag chain.<br />

These chains are continued in the first few lattice planes in the PbTe part. The<br />

theoretical values predict this behaviour, but the calculated displacements are too<br />

small to be visible in the simulated image. The measured displacements show this<br />

behaviour more pronounced. A comparison of the two HR simulation can be seen in


figure 6.15. This displacement can also be interpreted as a domination of the zb<br />

structure at the interface.<br />

This behaviour emphasises that the more important measurement is the comparison<br />

of the simulated HR images with the HRTEM images and thus the used values of the<br />

JEMS model construction. The atomic displacements, used for the JEMS model, are<br />

the mean values of the measured displacements which are shown in Figure 6.18. The<br />

best agreement can be found for the displacements parallel to the [001] direction. The<br />

measured values for the Te and Pb atoms show exactly the same behaviour as the<br />

calculated ones, except for the Te atom of the first PbTe lattice plane, as mentioned in<br />

the previous section. Both, the calculated and measured shift of the two crystal halves<br />

against each other are 0.04 nm. The maximum difference can be seen in the diagram<br />

for the Cd and Te atoms, where the difference to a undisturbed lattice is 0.06 nm for<br />

the calculated and 0.1 nm for the measured one. This is 10% change with respect to<br />

the lattice constant for the calculations and even more for the measured values. The<br />

calculated displacements parallel to the [001] direction for the Cd atoms describes<br />

also the tendency of the measured ones.<br />

The agreement of the displacements parallel to the [110] direction is also really good<br />

for the tellurium atoms, but measurement accuracy is in the range of the largest<br />

feature of the theoretic values, and one has to be careful with the interpretation. The<br />

same is valid for the displacements parallel to [110] for the Cd and Pb atoms. The<br />

calculations find a strong effect, where the atoms of every other lattice plane show a<br />

large displacement away from the interface. This would lead to a pairing of Cd and Pb<br />

atoms, respectively, for the CdTe and PbTe crystal halves. Each pair consists of a not-<br />

displaced and a displaced atom. The measurement shows the behaviour of a common<br />

displacement of the atoms, which decreases with increasing distance to the interface.<br />

In the diagram for the displacements parallel to [110] for Pb and Cd the measured<br />

values show the same tendency as the calculated ones, and represent the average value<br />

of them. The measurements for the displacements parallel to [110] were more difficult<br />

and less accurate, because the lattice plane spacing in this direction is near the<br />

practical resolution limit of the microscope, and the dark disks of the lattice fringes<br />

for PbTe overlap. The same is true for the CdTe lattice spacings parallel to the<br />

interface, even though there is no overlap.<br />

71


72<br />

Figure 6.18: Comparison of the displacements calculated by total energy calculation and<br />

the measured displacements used for the JEMS HR-map simulation. The standard error<br />

of the mean atomic displacements, evaluated from the HRTEM image, are indicated as<br />

dashed black lines in the diagrams. The displacements are ordered with respect to<br />

element and displacement direction. The gridline at zero on the abscissa indicates the<br />

nominal interface. The displacement directions are indicated in the projected crystal<br />

structures for each diagram. The agreement of the data is very good except for the<br />

theoretical predicted rumpling effect for the displacement parallel to [110] for the Pb and<br />

Cd atomic rows. But the measurement shows a mean value of the displacement with the<br />

same shift away from the interface. The shift of the two crystal halves against each other<br />

parallel to [001] can be seen best in the diagram for the Te displacements parallel to<br />

[001].


6.2.5 The (111) interface<br />

It is difficult to image the {111} interfaces. The reason is the triangular shape of<br />

these facets. If this interface is parallel to the electron beam and is focused, the<br />

projected length of the interface changes and makes it difficult to image this interface.<br />

Aside of that, two neighbouring {110} interfaces of a small {111} interface can form<br />

a direct intersection line above the {111} face. If one look along a {110} direction,<br />

this intersection line next to the small {111} interfaces complicates the identification<br />

of the {111} interfaces. This condition can be seen in Fig 4.12. Additionally, the<br />

{111} interfaces of larger dots are the smallest ones and the chance that the interface<br />

penetrates all through the specimen wedge is lower than for larger interfaces. The<br />

penetration through the wedge is an important condition for recording high quality<br />

HRTEM images. Therefore, the imaged {111} interfaces showed no clear interface<br />

lattice fringes.<br />

(111) interface modelling for the HR-simulations<br />

As mentioned before, it was difficult to investigate the (111) interface by TEM due<br />

to the dot geometry. Nevertheless, a model with this interface was used to visualise<br />

the pattern of this interface. The model was constructed with the conditions for the<br />

(001) and (110) interfaces and combines the three interfaces of the PbTe nanocrystals.<br />

This model also shows the corners between the (001), (111) and (110) interfaces (see<br />

Fig. 6.19). It was the largest model used. It consists of about 270 Atoms and was<br />

programmed in MATLAB. The (111) interface is polar for PbTe and CdTe. The<br />

rotational symmetry around a axis is the same for PbTe and CdTe, and one can<br />

build up four different (111) interfaces between CdTe and PbTe under the condition<br />

of atom number conservation. Two of these interfaces are sorted out, because the<br />

bond lengths of the atoms at the interface are too large or small in comparison to bond<br />

lengths in the bulk material and the other interface models. These two interfaces<br />

consist of the same number of atoms like the others, thus the atom number is<br />

conserved. This leads to a corner between the polar (001) interface and the neutral<br />

(110) interface, which contains one of the (111) interfaces defined by the crystal<br />

orientation and not by the (001) interface termination. The corners for both<br />

terminations along the [ 110]<br />

direction contain a (111) interface where the PbTe is Pb<br />

terminated and the CdTe is Te terminated. For the (111) interface along the<br />

[ 11<br />

0]<br />

direction the relations are vice versa. The main difference in the HR simulation<br />

of these two (111) interface is whether the PbTe part is completed by the CdTe or the<br />

PbTe lattice fringes.<br />

73


74<br />

Figure 6.19: Projection of the JEMS atomic model with all tree interfaces for both<br />

terminations and projection axes. The simulations are done for Δf = -61nm t = 1.4 nm (7<br />

(110) lattice planes). As can be seen, the termination of the (001) interface does not<br />

determine the kind of (111) interface. The orientation of the Cd-Te dimer determines the<br />

spacing of the (111) interface. The left image shows the model for the (111) interface<br />

parallel to the [ 1 10]<br />

direction. The right image shows the model for the (111) interface<br />

parallel to the [ 11<br />

0]<br />

direction.<br />

Figure 6.20 shows simulated HR-maps of this model. One can see the change from a<br />

positive phase contrast (dark areas are correlated to atomic positions) to a negative<br />

phase contrast (bright areas are correlated to atomic positions). Images for t = 1.4 nm<br />

and Δf = -56 and -61 nm (Scherzer defocus) show a positive phase contrast. The<br />

images for t = 5.5 nm and a Δf = -66 and -71 show a negative phase contrast. The<br />

simulations in between show images where lattice fringes from different lattice planes<br />

have different phase contrast. These images show strong variations in the appearance<br />

of the pattern.


Figure 6.20: HR-map of the model containing the Te terminated (001) interface with the<br />

projection axis [ 1 10]<br />

with respect to fig. 6.7. The thicknesses of 1.4, 2.7, 4.1, 5.5 nm<br />

correspond to 7, 14, 21, 28 (110)-lattice planes of PbTe. A direct, easy correlation to the<br />

atomic structure is only possible for very thin samples near the Scherzer defocus of -<br />

61nm.<br />

HRTEM images of the (111) interface<br />

The best HRTEM images of the (111) interface were obtained at the dot shown in<br />

figure 6.4. It is know from the analysis of the (110) and (001) interfaces of this dot<br />

that the images were recorded at almost ideal defocus, and that the specimen is thin<br />

enough. This holds also for the images of the {111} interface. The images can be seen<br />

in figure 6.21 and show two (111) interfaces to the left and right of a Te terminated<br />

(001) interface. The interfaces should have the atomic structure as shown on the Te<br />

75


76<br />

terminated model for the [ 1 10]<br />

projection direction (see figure 6.19). The lattice<br />

fringes at the interface (seen in the zoom-ins of figure 6.21) show left and right of the<br />

Te terminated interface the same behaviour. Characteristic is the dark pattern between<br />

the PbTe and the CdTe pattern, which could be either part of the CdTe or PbTe crystal<br />

halves. HR simulations, depicted in figure 6.19 in connection with the used model,<br />

show that this interface behaves like the Cd terminated (001) interface, i.e. the<br />

interface pattern is a mixture of the CdTe and PbTe lattice fringes. This cross-over is<br />

the stair shaped dark area in the small images of figure 6.21 along the diagonals.<br />

Similar patterns can be seen in the HR-map of figure 6.20. Assuming that the JEMS<br />

model of this interface is correct, these differences with respect to the simulation can<br />

only be understood if atomic displacements at this interface occur. Unfortunately,<br />

these two HRTEM images are the only high quality (111) interfaces available, and no<br />

images exist of the other kind with corresponding quality.<br />

Figure 6.21: Two images of the (111) interface. The images were recorded at ideal<br />

defocusing conditions at the dot shown in figure 6.2. The (001) interface between the two<br />

interfaces is Te terminated. The two zoom-ins show a Fourier filtered version of the<br />

interface and they show almost the same lattice fringes. The CdTe and PbTe parts are<br />

indicated left and right of the stair-shaped pattern along diagonal of the small images.


Chapter 7<br />

Conclusions and Outlook<br />

Conclusions<br />

The intense study of as-grown and annealed PbTe/CdTe heterostructures by TEM<br />

reveals many features of this material system. Firstly, it was shown that the PbTe epi-<br />

layer clad by CdTe transform into high quality PbTe QDs embedded in CdTe host<br />

material. The origin of this process is the large solid phase miscibility gap and the QD<br />

formation is driven by the minimisation of the total energy between PbTe and the<br />

CdTe host material. The origin of the immiscibility is the different lattice structure of<br />

the participating materials. Both materials have fcc lattice symmetry. CdTe possesses<br />

zb-lattice structure and has a partly covalent chemical bonding. Each atom in this<br />

crystal type has four next neighbours. The ionic crystal PbTe is of the rs-lattice type,<br />

with six next neighbour surrounding each atom. Dark and bright field investigation of<br />

several samples allowed the construction of the equilibrium shape. All QDs are<br />

defined by {001}, {110} and {111} interfaces. The geometric shape assembled with<br />

these interfaces is a cubo-octahedron. The shape is dependent on the QD size which is<br />

a hint for the influence of the edge energies on the equilibrium shape. A system with<br />

dominating interface energies forms an equilibrium shape, which is independent of<br />

the dot size.<br />

Results of these studies were size distributions of QDs precipitated from different<br />

PbTe epi-layer. The upper size limit for formed QDs is strongly dependent on the<br />

original layer thickness. Typical QDs, developed from a 5 nm epi-layer, are ~25 nm<br />

large. The most common dot size found in an annealed 1nm PbTe layer<br />

heterostructure is in the range of 7 to 10nm. The lower limit is nearly independent of<br />

the original layer thickness. The smallest QDs found in all annealed specimens consist<br />

of about 20 PbTe (002) lattice planes which correspond to a size of 4.6 nm. The<br />

density of the dots depends on the size distribution and the material amount of the<br />

77


78<br />

original SQW. Typical dot densities are e.g. 3 10 11 cm -2 for an annealed 1nm epi-layer<br />

or 6 10 10 cm -2 for an annealed 5nm epi-layer. The SQW thickness is well controllable<br />

and thus allows the control of the size and density distribution.<br />

The HRTEM interface studies, combined with multi-slice HR simulation, highlight<br />

the strong rebonding effects at the CdTe/PbTe interfaces. The {001} PbTe/CdTe<br />

consist of the non-polar PbTe {001} face and the polar {001} faces of CdTe which<br />

can be Cd or Te terminated. This leads to two possible {001} interfaces. The HRTEM<br />

investigations reveal the formation of both interface kinds on opposite {001}<br />

PbTe/CdTe interfaces. One (001) interface of the PbTe QW is terminated during MBE<br />

growth and allow the control of the position of the two kinds of the {001} interfaces.<br />

The main difference of these two interfaces is the lattice plane spacing at the nominal<br />

interface. This distance is a/2 for the Te terminated interface which corresponds to the<br />

lattice plane spacing of the parallel (002) PbTe lattice planes. The Cd terminated<br />

interface has an interface spacing of a/4, the lattice plane spacing of the (004) CdTe<br />

lattice planes. The PbTe crystal halves of both interfaces show a rumpling effect<br />

comparable to free PbTe surfaces. This feature is small for the Te terminated<br />

interface, where practically no other displacements effects occur. The rumpling is<br />

pronounced at the Cd terminated interface and additionally strong displacements can<br />

be found in the CdTe crystal halve of this (001) interface. These displacements allow<br />

the Cd atoms again a fourfold coordinated binding configuration at the interface. This<br />

behaviour allows the classification of the Te terminated interfaces as rs lattice<br />

dominated and the Cd terminated as zb lattice dominated. The displacements are<br />

findings of the ab-initio calculations by R. Leitsman. The two kinds of the (001)<br />

interfaces can be easily distinguished with the TEM and also strong indications of the<br />

displacements can be found.<br />

The {110} interface can also be seen as zb-lattice determined. The strong effects at<br />

this interface also lead to a better fourfold coordinated bonding configuration of the<br />

interfaces near CdTe components. This reconfiguration ends up in a shift of the CdTe<br />

and PbTe crystal halves against each other of 0.04nm. The Cd-Te chains of the zb<br />

lattice are continued in the PbTe part and result in atomic displacements of Pb and Te<br />

atoms in addition to the shift of the crystal halves. The displacements evaluated by<br />

HRTEM were compared with the atomic positions evaluated by the ab-initio<br />

calculations. Both methods reveal the same feature of the CdTe/PbTe interfaces and<br />

also the absolute values of the atomic displacements agree very well. The main effect<br />

of the displacements is again a better fulfilment of the four fold coordination of the<br />

atomic bonds in the zb-lattice near the interface.


Figure 7.1: (a) QDs interfaces stitched together from several HRTEM images. All<br />

interfaces are clearly visible and the main features are indicated. (b) shows a HR<br />

Simulation and (c) shows the according crystal model of a small QD. The orientation and<br />

the (001) terminations of the small dot correspond to the large one.<br />

The {111} face is a polar one for both materials, which results in four possible<br />

{111} interfaces. Two of them contain bondlengths, which are not realistic and no<br />

indications of these two excluded interfaces can be found in the HRTEM images. The<br />

79


80<br />

other two show the same characteristic as the two {001} interfaces. The {111}<br />

interface is difficult to image with the TEM, thus only one kind could identify by<br />

HRTEM images at the right position at the PbTe QD. This is a strong indication that<br />

the former consideration made about the {111} interface is correct and describes the<br />

behaviour of these interface class. A summary of the feature of the QD interfaces can<br />

be seen in Fig.7.1.<br />

Outlook<br />

The {111} interfaces of PbTe QDs are hardly accessible for TEM investigations,<br />

because their triangular shape hamper correct imaging. A possible method for an<br />

accurate characterisation of this interface class would be cross-sectional TEM on<br />

SQWs grown on (111) CdTe substrate with continuous {111} CdTe/PbTe interfaces.<br />

Also these SQWs should have two different kinds of {111} interfaces determined by<br />

the substrate termination as it is the case for the SQWs grown on (001) substrate.<br />

As the investigated samples of this thesis are optimally annealed for PL<br />

measurements and not for obtaining the ideal annealing conditions, further annealing<br />

studies of various samples with different epi-layer thicknesses would be necessary.<br />

An interesting, open question is also the diffusion of the compounds in the bulk CdTe<br />

or PbTe. There are no studies about the diffusion velocity of Pb in CdTe and only one<br />

study of the diffusion velocity of Cd in PbTe [47], which was done without the<br />

knowledge that PbTe forms QDs embedded in CdTe. The knowledge of the exact<br />

diffusions processes would give a deeper insight in the QDs formation processes<br />

during the annealing step and would allow a better controllability of the size and<br />

density distribution.<br />

Another attempt of QD production in the CdTe/PbTe material system was the<br />

implantation of Pb ions into bulk (111) CdTe or CdTe grown on (111) GaAs<br />

substrate. The implantation was done at an acceleration voltage of 200keV with<br />

different doses (10 15 cm -2 , 10 16 cm -2 ) of Pb ions. The high acceleration voltage and the<br />

heavy Pb ions result in an amorphisation of the CdTe bulk material. After an<br />

annealing step and recrystallisation of the CdTe, PbTe QDs can be found. The QDs<br />

show a broad size distribution. The found dots also show the three interfaces of the<br />

epitaxial fabricated QDs but they have more imperfections. An interesting question of<br />

this fabrication method is the controllability of the size and density distribution and<br />

also the in-depth location of the QDs as a function of the dose and the acceleration<br />

voltage. The stoichiometry is not fulfilled in this system. The excess Cd atoms must


segregate at the surface, build lattice defects or could change the Te-terminated<br />

interfaces to Cd-terminated ones. But this would lead to QDs containing a dipole.<br />

Further investigation of this system could give a deeper understanding of these<br />

processes. An example of PbTe QDs, fabricated by ion implantation can be seen in<br />

Fig. 7.2.<br />

Figure 7.2: (a) PbTe dots in CdTe matrix produced by Pb ions implanting in bulk (111)<br />

CdTe with 200keV and a dose of 10 16 cm -2 . The size of the QDs is distributed from about<br />

25nm to 5nm. (b) shows a small QD from a sample with 10 15 cm -2 implanted Pb ions.<br />

The QDs show more imperfection than the epitaxial grown ones.<br />

The goal of all these investigation is of course an application of the QDs produced<br />

by lattice type mismatch. Semiconductor laser in the mid infrared frequency range<br />

would open the possibility of modern molecule spectroscopy for e.g. medical<br />

diagnosis. The CdTe/PbTe material system grown on GaAs substrate is not suitable<br />

for laser application, because the refractive index of CdTe is smaller than the<br />

refractive index of GaAs. This would lead to a deflection of emitted light into the<br />

GaAs substrate. A solution of this problem could be the use of HgxCd1-xTe as bulk<br />

material. The refractive index of this ternary alloy is near four and is larger than the<br />

refractive index of GaAs. HgTe is like CdTe zb structure and has also almost the same<br />

lattice constant as PbTe. But Hg is a really difficult and highly toxic material for<br />

utilisation in a MBE system. The approach of the lattice type mismatch QDs is also<br />

available for other material systems. One example of another material system, which<br />

was investigated, is ErAs on GaAs, mentioned in the introduction. However, ErAs is<br />

not a semiconductor but a semi-metal. Another interesting material is BaF2. BaF2 is<br />

fluorite structure, a cubic lattice structure, and has also a lattice constant near the<br />

lattice constant of PbTe and CdTe. Therefore also the material combinations of<br />

BaF2/PbTe or CdTe/BaF2 would be an interesting attempt for investigation. As the<br />

refractive index of BaF2 is small (n=~1.4, for the mid infrared frequency range), it is<br />

also a substrate candidate of CdTe/PbTe structures.<br />

81


82<br />

Appendix A<br />

Structure factors of CdTe and PbTe<br />

CdTe and PbTe can be described by a cubic unit cell. CdTe has zinc blend structure,<br />

PbTe has rock salt structure. Both materials can be described by a face centred cubic<br />

(fcc) unit cell with different basis. That why we first treat the structure factor of a<br />

simple fcc lattice. A detailed description can be found in [32].<br />

The expression for Fhkl, the structure factor for the (hkl) spot is given by:<br />

The fi is the atomic form factor, given by the differential cross section of the atoms.<br />

The coordinates xi, yi and zi describe the atomic positions in the unit cell. For the fcc<br />

lattice these positions are given by ( 0, 0, 0), ( ½, ½, 0), ( ½, 0, ½) and ( 0, ½, ½). This<br />

lead to the structure factor of the fcc lattice:<br />

If h, k, l are all even or all odd all three e-function are one. If the values of h, k, l are<br />

mixed even and odd, two of the e-functions are -1. This lead to following structure<br />

factors:<br />

hkl<br />

• h, k, l are all even or all odd →<br />

• h, k, l are mixed even and odd →<br />

F<br />

fcc<br />

F<br />

=<br />

∑<br />

i<br />

f<br />

2πi<br />

ie<br />

F<br />

F<br />

( hx + ky + lz )<br />

πi(<br />

h+<br />

k { ) πi(<br />

h+<br />

l)<br />

πi(<br />

k+<br />

l<br />

=<br />

f 1+<br />

e + e + e<br />

) }<br />

i<br />

= 4f<br />

i<br />

= 0<br />

i<br />

(3.26)<br />

(3.27)


A1 CdTe structure factor<br />

CdTe has zinc blend structure. This can be described by a fcc lattice with a two-atom<br />

base and a base vector of ( ¼, ¼, ¼ ). In the origin sits a tellurium atom, at the end of<br />

the base vector a cadmium atom. fTe and fCd describe the atomic form factor of the<br />

atom species. The structure factor becomes:<br />

This leads to the following rules:<br />

• h, k, l are mixed even and odd →<br />

• h, k, l are all odd →<br />

• h, k, l are all even, h+k+l=2N, N is odd →<br />

F<br />

F<br />

F<br />

• h, k, l are all even, h+k+l=2N, N is even F →<br />

A2 PbTe structure factor<br />

= 0<br />

= 4(fTe±ifCd)<br />

= 4(fTe-fCd)<br />

= 4(fTe+fCd)<br />

PbTe has rock salt structure. This can be described by a fcc lattice with a two-atom<br />

base and a base vector of ( ½, ½, ½ ). The base consists of a Pb-Te dimer. The atomic<br />

form factors are again given by fPb and fTe. The structure factor becomes:<br />

The selection rules are given by:<br />

• h, k, l are all even →<br />

• h, k, l are all odd →<br />

• h, k, l are mixed →<br />

F =<br />

⎧<br />

⎨f<br />

Te + f Cde<br />

⎩<br />

F<br />

F<br />

F<br />

π<br />

2<br />

= 4(fTe+fPb)<br />

= 4(fTe-fPb)<br />

= 0<br />

( h+<br />

k+<br />

l)<br />

⎫<br />

⎬F<br />

⎭<br />

fcc<br />

π { ( h+<br />

k+<br />

l<br />

F =<br />

f<br />

)<br />

Te + fPbe<br />

} Ffcc<br />

83<br />

(3.28)<br />

(3.29)


84<br />

Appendix B<br />

Simulated HR-maps of CdTe/PbTe interfaces<br />

B1 The (001) interface<br />

The four HR-maps (Figure B2 to B5) are for the two different terminations of the<br />

CdTe crystal halve and the two different projections along a zone axes. The<br />

crystal orientation indication relates to figure B1.<br />

Figure B1: The (001) lattice plane through the origin represents the Te termination. The<br />

first (001) lattice plane filled with Cd near the origin represents the Cd termination.


Figure B2: HR-map for the Te terminated (001) interface along the [ 1 10]<br />

direction.<br />

Figure B3: HR-map for the Cd terminated (001) interface along the [ 1 10]<br />

direction.<br />

85


86<br />

Figure B4: HR-map for the Te terminated (001) interface along the [ 1 10]<br />

direction.<br />

Figure B5: HR-map for the Cd terminated (001) interface along the [ 1 10]<br />

direction.


B2 The (110) interface<br />

The tree HR-maps show simulations of the tree used models. Figure B6 is the HR-<br />

map of the JEMS model with the continuous Te matrix. Figure B7 shows the HR-map<br />

of the model constructed with the displacements calculated by DFT. Figure B8 shows<br />

the HR-maps constructed with the displacements extracted from HRTEM images.<br />

Figure B6: HR-map of the (110) interface model with the continuous Te matrix.<br />

87


88<br />

Figure B7: HR-map of the (110) interface model with the calculated displacements.


Figure B8: HR-map of the (110) interface model with measured displacements.<br />

89


90<br />

B3 The (111) interface<br />

For the sake of completeness a HR-map of a (111) interface is shown. Only one kind<br />

of the two different (111) interface was images with the TEM. The HR-map (see<br />

figure B9) is the corresponding simulation to Figure 6.15, shown in chapter 6. The<br />

used JEMS model was the model containing all tree interfaces and the corner between<br />

these interfaces.<br />

Figure B9: HR-map of the (111) interface model with the Te terminated (001) interface<br />

for a projection direction along the [ 1 10]<br />

direction.


References<br />

[1] R.C. Ashoori, Nature 379, 423 (1996)<br />

[2] G. Schedelbeck, W. Wegscheider, M. Bichler, G. Abstreiter, Science 278, 1792 (1997)<br />

[3] S. Fafard, K. Hinzer, S. Raymond, M. Dion, J. McCaffrey, Y. Feng, S. Charbonneau;<br />

Science 274, 1350 (1996)<br />

[4] Y. Arakava, A. Yariv, IEEE J. Quantum Electron. 22, 1887 (1986)<br />

[5] Z. Yuan, B.E. Kardynal, R.M. Stevenson, A.J. Shields, C.J. Lobo, K. Cooper, N.S. Beattie,<br />

D.A. Ritchie, M. Pepper, Science 295, 102 (2002)<br />

[6] P. Michler, A. Kiraz, C. Becher, W.V. Schoenfeld, P.M. Petroff, L. Zhang, E. Hu,<br />

A. Imamogulu, Science 290, 2282 (2000)<br />

[7] D. Leonard, M. Krishnamurthy, C.M. Reaves, S.P. Denbaars, P.M. Petroff,<br />

Appl. Phys. Lett. 63, 3203 (1993)<br />

[8] J.M. Moison, F. Houzay, F. Barthe, L. Leprince, E. André, O. Vatel, Appl. Phys. Lett. 64, 196<br />

(1994)<br />

[9] V.A. Schukin, D. Bimberg, Rev. Mod. Phys. 71, 1125 (1999)<br />

[10] G. Springholz, V. Holy, M. Pinczolits, G. Bauer, Science 282, 734 (1998)<br />

[11] J. Stangl, V.Holy, G. Bauer, Rev. Mod. Phys. 76, 725 (2004)<br />

[12] T. Walther, A.G. Cullis, D.J. Norris, M. Hopkinson, Phys. Rev. Lett. 86, 2381 (2001)<br />

[13] A. Rosenauer, D. Gerthsen, D. Van Dyck, M. Arzberger, G. Böhm, G. Abstreiter,<br />

Phys. Rev. B 64, 245334 (2001)<br />

[14] A.P. Alivisatos, Science 271, 933 (1996)<br />

[15] C.B. Murray, C.R. Kagan, M.G. Bawendi, Science 270, 1335 (1995)<br />

[16] E. Saucedo, L. Fornaro, V. Corregidor, E. Diéguez, Eur. Phys. J. Appl. Phys. 27, 427 (2004)<br />

[17] E. Saucedo, V. Corregidor, L. Fornaro, N.V. Sochinskii, J. Silveira, E. Diéguez,<br />

Eur. Phys. J. Appl. Phys. 27, 207 (2004)<br />

[18] T. Scheidt, E.G. Rohwer, H.M. von Bergmann, E. Saucedo, E. Diéguez, L. Fornaro,<br />

H. Stafast, J. of Appl. Phys. 97, 103104 (2005)<br />

[19] W. Heiss, H. Groiss, E. Kaufmann, G. Hesser, M. Böberl, G. Springholz, F. Schäffler,<br />

K. Koike, H. Harada, M. Yano, Appl. Phys. Lett. 88, 192109 (2006)<br />

[20] C. Kadow, J.A. Johnson, K. Kolstad, J.P. Ibbetson, A.C. Gossard,<br />

J. Vac. Sci. Technol. B 18(4), 2197, (2000)<br />

[21] C. Kadow, J.A. Johnson, K. Kolstad, A.C. Gossard, J. Vac. Sci. Technol. B 21(1), 29, (2003)<br />

[22] D.O. Klenov, D.C. Driscoll, A.C. Gossard, S. Stemmer, Appl. Phys. Lett. 86, 111921, (2005)<br />

91


92<br />

[23] M.P. Hanson, D.C. Driscoll, C. Kadow, A.C. Gossard, Appl. Phys. Lett. 84(2), 221, (2004)<br />

[24] V. Leute, N.J. Haarmann and H.M. Schmidtke; Z. Phys. Chem. 190, 253 (1995)<br />

[25] I-hsiu Ho and G.B. Springfellow, Appl. Phys. Lett. 69, 2702 (1996)<br />

[26] K. Koike, T. Honden, I. Makabe, F.P. Yan, M. Yano, J. Crystal Growth 257, 212 (2003)<br />

[27] M.A. Herman, H. Sitter, Molecular Beam Epitaxy – Fundamentals and Currents Status, 2 th ed.,<br />

Springer, Berlin Heidelberg, (1989)<br />

[28] K. Koike, T. Tanaka, S. Li, M. Yano, J. Crystal Growth 227, 671 (2001)<br />

[29] M. Yano, I. Makabe, K. Koike, Physica E 20, 449 (2004)<br />

[30] W. Schwinger, Transmissionselektronenmikroskopie an <strong>Halbleiter</strong> - Nanostrukturen,<br />

<strong>Diplom</strong>athesis, <strong>Institut</strong> <strong>für</strong> <strong>Halbleiter</strong>physik, Johannes Kepler Universität Linz,<br />

(Februar 2003)<br />

[31] W. Schwinger, Epitaxial Overgrowth of Fullerenes on Si(100), <strong>Diplom</strong>athesis, <strong>Institut</strong> <strong>für</strong><br />

<strong>Halbleiter</strong>physik, Johannes Kepler Universität Linz, (April 2003)<br />

[32] David. B. Williams und C. Barry. Carter, Transmission Electron Microscopy - A Textbook for<br />

Material Science, Plenum Press, New York & London, (1996)<br />

[33] L. Reimer, Transmission Electron Microscopy - Physics of Image Formation and<br />

Microanalysis, 4 th ed., Springer, (1997)<br />

[34] JEOL LTD., 1-2 Musashino 3-chome Akishima Tokyo 196-8558 Japan, JEM-2011, 1999.<br />

[35] JEOL LTD., 1-2 Musashino 3-chome Akishima Tokyo 196-8558 Japan, Ultra-High<br />

Resolution Analytical Electron Microscope, Spezification.<br />

[36] Charles Kittel, Einführung in die Festkörperphysik, Oldenburg, München&Wien, 1993<br />

[37] Gatan, Inc., 5933 Coronado Lane, Pleasanton, CA94588, USA, UltraSonic Disc Cutter,<br />

Model601, User's Guide, November 1998, Revision2.<br />

[38] Gatan, Inc., 5933 Coronado Lane, Pleasanton, CA94588, USA, Dimple Grinder, Model656,<br />

User's Guide, November 1998, Revision2.<br />

[39] Gatan, Inc., 5933 Coronado Lane, Pleasanton, CA94588, USA, Precision Ion Polishing<br />

System, User's Guide, November 1998, Revision3.<br />

[40] Michael Teuchtmann, Einführung in die TEM-Probenpräparation, TSE, Universität Linz,<br />

2001.<br />

[41] DigitalMicrograph 3.10.1 for GMS 1.5.1, Gatan Inc., Pleasanton, CA, USA<br />

[42] R.Leitsmann, L.E.Ramos, F.Bechstedt, Phys.Rev. B 74, 085309, 2006<br />

[43] P. Stadelmann, JEMS manual,<br />

http://cimewww.epfl.ch/people/stadelmann/jemsWebSite/jems.html<br />

[44] Dynamical theory of electron diffraction at small angles,<br />

http://cimewww.epfl.ch/people/stadelmann/<br />

/jemsUpgrade/DynamicalTheoryNov27.html<br />

[45] A.A. Lazarides, C.B. Duke, A. Paton and A. Kahn, Phys. Rev. B 52, 18495, 1995<br />

[46] D.O. Klenov, J.M. Zide, J.D. Zimmerman, A.C. Gossard, S. Stemmer, Appl. Phys. Lett. 86,<br />

241901 (2005)<br />

[47] V.B. Bobruiko, T.A. Kouznetzova, M.P. Belyansky, A.M. Gaskov, Materials Science and<br />

Engineering B32, 7, (1995)


Curriculum Vitae<br />

1 March 1979 born in Linz<br />

Sept. 1985 - July 1989 Primary school in Unterweitersdorf<br />

Sept 1989 - July 1993 Secondary school in Gallneukirchen<br />

Sept. 1993 - July 1998 LiTec in Linz (Secondary Technical and<br />

95<br />

Vocational College), branch of study: mechanical<br />

engineering, Matura (general qualification for<br />

university entrance) July 1998<br />

Oct. 1998 - Oct. 1999 Civilian service at the Lebenshilfe OÖ<br />

(association for mental- and multi-handicapped<br />

people)<br />

Since Oct. 1999 Study of Technical Physics at the Johannes<br />

Kepler University, Linz<br />

Sept. 2004 - Feb. 2005 Study of Physics at the University of Groningen,<br />

Netherlands<br />

Since Aug. 2005 <strong>Diplom</strong>a Thesis for Technical Physics at<br />

the <strong>Institut</strong>e of Semiconductor Physics,<br />

Johannes Kepler Universität, Linz,<br />

Univ. Prof. Dr. Friedrich Schäffler


96<br />

List of Publications<br />

“Centrosymmetric PbTe/CdTe quantum dots coherently embedded by epitaxial precipitation”<br />

W. Heiss, H. Groiss, E. Kaufmann, G. Hesser, M. Böberl, G. Springholz, F. Schäffler,<br />

K. Koike, H. Harada, M. Yano<br />

Appl. Phys. Lett. 88, 192109 (2006)<br />

“The coherent {100} and {110} interfaces between rocksalt-PbTe and zincblende-CdTe”<br />

H. Groiss, W. Heiss, F. Schäffler, R. Leitsmann, F. Bechstedt, K. Koike, H. Harada, M. Yano<br />

Submitted to JCG Proceedings for 14th International Conference on Molecular Beam Epitaxy,<br />

MBE 2006, September3-8 2006, Tokyo, Japan<br />

“Quantum dots with coherent interfaces between rocksalt-PbTe and zincblende-CdTe”<br />

W. Heiss, H. Groiss, E. Kaufmann, G. Hesser, M. Böberl, G. Springholz, F. Schäffler,<br />

R. Leitsmann, F. Bechstedt, K. Koike, H. Harada, M. Yano<br />

Submitted to AIG Proceedings for 28th International Conference on the Physics of<br />

Semiconductors, ICPS 2006, July 24-28 2006, Vienna, Austria<br />

“Photoluminescence Characterization of PbTe/CdTe Quantum Dots Grown by Lattice-Type<br />

Mismatched Epitaxy”<br />

K. Koike, H. Harada, T. Itakura, M. Yano, W. Heiss, H. Groiss, E. Kaufmann, G. Hesser,<br />

M. Böberl, G. Springholz, F. Schäffler<br />

Submitted to JCG Proceedings for 14th International Conference on Molecular Beam Epitaxy,<br />

MBE 2006, September3-8 2006, Tokyo, Japan<br />

“Rebonding at coherent interfaces between rocksalt-PbTe/zinc-blende-CdTe”<br />

R. Leitsmann, L. E. Ramos, F. Bechstedt, H. Groiss, F. Schäffler, W. Heiss, K. Koike,<br />

H. Harada, M. Yano<br />

Submitted to New Journal of Physics


Contributed talks:<br />

“The coherent {100} and {110} interfaces between rocksalt-PbTe and zincblende-CdTe”<br />

H. Groiss, W. Heiss, F. Schäffler, R. Leitsmann, F. Bechstedt, K. Koike, H. Harada, M. Yano<br />

14th International Conference on Molecular Beam Epitaxy, MBE 2006, September3-8 2006,<br />

Tokyo, Japan<br />

Invited talks:<br />

“Quantum dots with coherent interfaces between rocksalt-PbTe and zincblende-CdTe”<br />

W. Heiss, H. Groiss, E. Kaufmann, G. Hesser, M. Böberl, G. Springholz, F. Schäffler,<br />

R. Leitsmann, F. Bechstedt, K. Koike, H. Harada, M. Yano<br />

28th International Conference on the Physics of Semiconductors, ICPS 2006,<br />

July 24-28 2006, Vienna, Austria<br />

97


98<br />

Acknowledgement<br />

I would like to thank those who supported me during my diploma thesis and thus contributed<br />

to the successful completion of this work.<br />

Especially I would like to thank:<br />

The group of Prof. Ph.D. M. Yano at Osaka institute of technology, especially<br />

Ph.D. K. Koike, for the supply of excellent samples and valuable discussions during our<br />

meeting at the MBE2006 conference in Tokyo;<br />

DI G. Hesser, for carrying out excellent sample preparation, giving me a detailed<br />

introduction into the TEM and a amicable support during my work at the TEM;<br />

DI R. Leitsmann and Prof. Dr. F. Bechstedt for the ab-inito calculations and interesting<br />

discussion about the PbTe/CdTe QDs project;<br />

Prof. Dr F. Schäffler for giving me the great opportunity of this diploma thesis, the scientific<br />

support and the suggestions for my diploma thesis;<br />

All member of the institute working on the PbTe/CdTe project, especially Prof. Dr. W. Heiss<br />

and DI E. Kaufmann for helpful support and valuable discussion;<br />

Prof. Dr. G. Bauer for the possibility to write my diploma thesis at the <strong>Institut</strong>e of<br />

Semiconductor Physics;<br />

DI D. Pachinger and Dr. H. Lichtenberger for the introduction and support concerning the<br />

TEM;<br />

DI M. Ratajski for the good cooperation at the TEM within the Department for Technical<br />

Support;


All member of the institute for the friendly atmosphere, especially the members of the office<br />

for the help at all problems concerning the diploma thesis and the creative working<br />

atmosphere;<br />

My parents Gertrud and Franz Groiss for financial support and general support all the years<br />

of my studies and also my brother Stefan for his support;<br />

The PbTe/CdTe project is part of the SFB 25 IR-ON. I would like to thank the Fond zur<br />

Förderung der Wissenschaftlichen Forschung (Austria) for their financial support.<br />

I would like to thank the Marubun Research Promotion Foundation for the travel grant for<br />

the MBE2006 conference in Tokyo, Japan.<br />

99

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!