01.12.2012 Views

Kinetic and Strain-Induced Self-Organization of SiGe ...

Kinetic and Strain-Induced Self-Organization of SiGe ...

Kinetic and Strain-Induced Self-Organization of SiGe ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

J O H A N N E S K E P L E R<br />

U N I V E R S I T Ä T L I N Z<br />

N e t z w e r k f ü r F o r s c h u n g , L e h r e u n d P r a x i s<br />

<strong>Kinetic</strong> <strong>and</strong> <strong>Strain</strong>-<strong>Induced</strong> <strong>Self</strong>-<strong>Organization</strong> <strong>of</strong> <strong>SiGe</strong><br />

Heterostructures<br />

Dissertation<br />

zur Erlangung des akademischen Grades<br />

Doktor der Technischen Wissenschaften<br />

Angefertigt am Institut für Halbleiterphysik<br />

Betreuung:<br />

Univ. Pr<strong>of</strong>. Dr. Friedrich Schäffler<br />

Eingereicht von:<br />

DI Herbert Lichtenberger<br />

Gutachter:<br />

1. Univ. Pr<strong>of</strong>. Dr. Friedrich Schäffler<br />

2. Dr. Detlev Grützmacher<br />

Linz, August 2006<br />

Johannes Kepler Universität<br />

A-4040 Linz · Altenbergerstraße 69 · Internet: http://www.jku.at · DVR 0093696


Preface<br />

Eidesstattliche Erklärung<br />

Ich erkläre an Eides statt, dass ich die vorliegende Doktorarbeit selbstständig und ohne<br />

fremde Hilfe verfasst, <strong>and</strong>ere als die angegebenen Quellen und Hilfsmittel nicht benutzt bzw.<br />

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

i<br />

Herbert Lichtenberger


ii PREFACE<br />

Kurzfassung<br />

Eine kinetische Wachstumsinstabilität in der Silizium-Germanium Molekularstrahlepitaxie<br />

(<strong>SiGe</strong>-MBE) wurde dazu verwendet, Si-Substrate mit periodisch modulierter Oberflächen-<br />

struktur zu erzeugen. Diese kinetische ”Step-Bunching” Instabilität führt für die verwendeten<br />

vizinalen Si(001) Substrate mit 4 ◦ Verkippung entlang [110] zu einer Welligkeit mit einer<br />

Periode von ungefähr 100 nm und einer Amplitude von etwa 4 nm.<br />

Im ersten Teil dieser Arbeit werden Aspekte selbstorganisierten Wachstums im Si/<strong>SiGe</strong><br />

System beh<strong>and</strong>elt. Die spezielle ”Step-Bunching”-Oberflächenstruktur wurde benutzt, um<br />

das Zusammenspiel zwischen Kinetik, Oberflächenenergie und Verspannung zu untersuchen.<br />

Diese ”Step-Bunching”-Morphologie ist eine ein-dimensionale Oberflächenstruktur, die be-<br />

vorzugte Keimzentren für <strong>SiGe</strong>-Inseln entlang der asymmetrischen Wellenflanken bietet. Dies<br />

ermöglicht uns, kinetische und Verspannungs-induzierte Selbstorganisations-Phänomene im<br />

Si/<strong>SiGe</strong> Heterosystem zu kombinieren. Bei gemäßigten Wachstumstemperaturen um 425 ◦ C<br />

wurden Verspannungs-induzierte Hügelketten beobachtet, die die Flanken senkrecht zur ”Step-<br />

Bunching”-Struktur dekorieren. Die gesamte sich ergebende Morphologie wird dann nur<br />

von (001)-orientierten Oberflächen und {105}-Facetten gebildet. Diese {105}-facettierten<br />

Hügelketten sind anscheinend energetisch bevorzugt und zeigen sich geometriebedingt an<br />

geneigten Substratflächen, deren Ausrichtung in etwa der einer {1 1 10}-Ebene entspricht.<br />

Diese Strukturen weisen eine relativ gute Ordnung auf, die gänzlich auf Selbstorganisati-<br />

on beruht. Derartige morphologische Strukturen bilden die Basis für das Verständnis des<br />

Ge-Insel Keimbildungsmechanismus auf zwei-dimensional vorstrukturierten Si-Substraten.<br />

Im zweiten Teil werden Magneto-Transport Messungen an p-modulationsdotierten Si/<strong>SiGe</strong><br />

Heterostrukturen diskutiert, die auf ”Step-Bunching” Si-Puffer gewachsen wurden. Die kurz-<br />

periodische Oberflächenstruktur des Si-Puffers führt zu einer wohldefinierten Modulation im<br />

<strong>SiGe</strong>-Kanal, was die Anzahl der Streuprozesse im p-dotiertem Quanten-Topf (p-MODQW)<br />

steigern müsste. Daraus resultiert eine Asymmetrie in der Ladungsträgerbeweglichkeit senk-<br />

recht und parallel zur wellenartigen Modulation des elektrisch leitfähigen Kanals, die dazu<br />

beitragen könnte, die verschiedenen Streumechanismen, insbesondere Legierungs- und Grenz-<br />

flächenrauhigkeitsstreuung, zu unterscheiden. Obwohl erst wenige Messungen hierfür durch-<br />

geführt wurden, zeigen erste Ergebnisse eine verminderte Tief-Temperatur Beweglichkeit quer<br />

zur Oberflächenmodulation, und zwar um einen nennenswerten Faktor zwei. Weitere die-<br />

ser Messungen in Kombination mit zusätzlicher Modellierung könnten einen neuen Zugang<br />

eröffnen, um die lang anhaltende Debatte beizulegen und den Mechanismus zu eruieren, der<br />

die Ladungsträgerbeweglichkeit der Löcher in p-MODQW Strukturen limitiert.


PREFACE iii<br />

Abstract<br />

A kinetic growth instability <strong>of</strong> silicon-germanium molecular beam epitaxy (<strong>SiGe</strong>-MBE) was<br />

used to generate periodic Si-ripple-templates. By increasing the substrate miscut to 4 ◦ the<br />

ripple period could be tuned to smaller periods towards the nanometer scale. At Si growth<br />

rates <strong>of</strong> 0.2 ˚A/s a ripple pattern with few defects develops within a small temperature window<br />

around 425 ◦ C. For our vicinal Si(001) substrates with 4 ◦ miscut along [110] this step-bunching<br />

instability provides undulations <strong>of</strong> about 100 nm periodicity <strong>and</strong> 4 nm amplitude.<br />

In the first part <strong>of</strong> this work several aspects <strong>of</strong> self-organized growth in the Si/<strong>SiGe</strong> system<br />

could be addressed. The special morphology <strong>of</strong> step-bunching was used to investigate the<br />

interplay between kinetics, surface energy <strong>and</strong> strain. The step-bunching templates provide<br />

a one-dimensional pattern with preferable nucleation sites for <strong>SiGe</strong>-isl<strong>and</strong>s along the ripple<br />

flanks, <strong>and</strong> thus allow us to combine kinetic <strong>and</strong> strain-driven self-organization phenomena<br />

in the Si/<strong>SiGe</strong> heterosystem. At moderate temperatures around 425 ◦ C strain-driven ridges,<br />

which decorate the ripple flanks perpendicular to the step bunches, were observed. The<br />

whole morphology is only made up with (001)-surfaces <strong>and</strong> {105}-facets. These {105}-faceted<br />

ridges appear to be energetically preferable for geometrical reasons at slopes, which are close<br />

to {1 1 10}-planes. These structures show a fair degree <strong>of</strong> ordering entirely based on self-<br />

organization. Furthermore, such morphological features are the basis for underst<strong>and</strong>ing the<br />

mechanism <strong>of</strong> Ge-dot nucleation on 2D pit-patterned Si-templates.<br />

In the second part magnetotransport measurements on p-modulation doped Si/<strong>SiGe</strong> het-<br />

erostructures grown on top <strong>of</strong> a step-bunched Si-buffer are discussed. The short-scale peri-<br />

odic height fluctuations <strong>of</strong> the Si-buffer form well-defined undulations in the <strong>SiGe</strong>-channel.<br />

This should increase scattering in the remotely p-doped quantum well (p-MODQW). Thus<br />

an asymmetry in mobility perpendicular <strong>and</strong> parallel to the undulations is expected, which<br />

might help to uncouple the different scattering mechanisms. These, namely alloy scattering<br />

<strong>and</strong> interface-roughness related scattering, are conversely discussed as predominant hole-<br />

mobility limiting factors for p-modulation doped structures. Although still at the beginning,<br />

the first measurements confirm a decreased low-temperature mobility across the undulations<br />

by a remarkable factor <strong>of</strong> two. Further measurements combined with additional modeling<br />

are expected to provide a new approach toward settling the long-lasting dispute on the hole-<br />

mobility-limiting scattering mechanisms in p-MODQW structures.


iv PREFACE


Contents<br />

Preface i<br />

Eidesstattliche Erklärung . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i<br />

Kurzfassung . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii<br />

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii<br />

I Basics 1<br />

1 Silicon-Germanium Material System 3<br />

1.1 Structural Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.1.1 Crystal Orientation <strong>and</strong> Miller Indices . . . . . . . . . . . . . . . . . . 5<br />

1.2 Vicinal Si(001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.2.1 Step Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.2.2 Step Types in Different Miscut Angle Regimes . . . . . . . . . . . . . 8<br />

1.3 Electronic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.3.1 B<strong>and</strong> Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.4 Silicon Germanium Alloys (Si1−xGex) . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.4.1 B<strong>and</strong>structure <strong>of</strong> <strong>SiGe</strong>-Heterostructures . . . . . . . . . . . . . . . . . 12<br />

2 Molecular Beam Epitaxy (MBE) 15<br />

2.1 MBE-Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.1.1 Fundamentals <strong>and</strong> Prerequisites . . . . . . . . . . . . . . . . . . . . . 15<br />

2.1.2 Physical Processes in the Growth-Chamber . . . . . . . . . . . . . . . 17<br />

v


vi CONTENTS<br />

2.1.3 Growth-Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1.4 Heteroepitaxy versus Homoepitaxy . . . . . . . . . . . . . . . . . . . 20<br />

2.1.5 Relevant Temperatures in MBE . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2 Cleaning Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.1 Pre-Cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.2.2 RCA-Cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.2.3 HF-Dip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.2.4 Oxide-Desorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.3 MBE-System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.3.1 Vacuum System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.3.2 Evaporation <strong>and</strong> Rate Measurement . . . . . . . . . . . . . . . . . . . 26<br />

2.3.3 Heating <strong>and</strong> Temperature Measurement . . . . . . . . . . . . . . . . . 26<br />

2.3.4 Controlling <strong>and</strong> Monitoring . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

3 Methods <strong>of</strong> Investigation 29<br />

3.1 Transmission Electron Microscopy (TEM) . . . . . . . . . . . . . . . . . . . . 30<br />

3.2 Scanning Electron Microscopy (SEM) . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.3 Atomic Force Microscopy (AFM) . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.3.1 Scanning Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.3.2 AFM-Probes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.3.3 Imaging Artifacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.3.4 Data Types <strong>and</strong> Data Representation . . . . . . . . . . . . . . . . . . 36<br />

II Main Research 41<br />

4 <strong>Self</strong>-Organized Growth – Step-Bunching 43<br />

4.1 Introduction to Step-Bunching . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

4.1.1 Dependence on Temperature, Rate, Thickness, Ge-content <strong>and</strong> Miscut 45<br />

4.1.2 <strong>Kinetic</strong> vs. <strong>Strain</strong>-<strong>Induced</strong> Step-Bunching . . . . . . . . . . . . . . . . 47<br />

4.2 Optimization <strong>of</strong> Step-Bunching for Ripple-Patterns . . . . . . . . . . . . . . . 48


CONTENTS vii<br />

5 <strong>Self</strong>-Organized Growth 2 – <strong>Kinetic</strong>s <strong>and</strong> <strong>Strain</strong> 55<br />

5.1 Introduction to State-<strong>of</strong>-the-Art Ge-Dots . . . . . . . . . . . . . . . . . . . . . 55<br />

5.2 <strong>Kinetic</strong> Step-Bunching <strong>and</strong> <strong>Strain</strong>-Driven Isl<strong>and</strong> Growth . . . . . . . . . . . . 57<br />

5.2.1 Influence <strong>of</strong> Ripple-Template Annealing . . . . . . . . . . . . . . . . . 57<br />

5.2.2 <strong>SiGe</strong>-Overgrowth <strong>of</strong> Step-Bunching Template – <strong>Strain</strong>-Effects . . . . . 57<br />

5.3 Ordering <strong>and</strong> Size <strong>of</strong> <strong>SiGe</strong>-Isl<strong>and</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

5.4 Closer Look on Surface Energy Effects – Facetting . . . . . . . . . . . . . . . 70<br />

5.4.1 Surface Energy Minimization via Excessive Facetting . . . . . . . . . . 77<br />

5.4.2 Step-Bunching Templates for p-Modulation Doped <strong>SiGe</strong>-Structures . . 83<br />

6 p-Modulation Doping <strong>and</strong> Mobility Analysis 85<br />

6.1 Introduction to p-Modulation Doping . . . . . . . . . . . . . . . . . . . . . . 86<br />

6.2 Experimental Aspects <strong>and</strong> Mobility Analysis . . . . . . . . . . . . . . . . . . 90<br />

6.2.1 Processing <strong>of</strong> Hall-Bars . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

6.2.2 Cryo-Measurements <strong>and</strong> Data Evaluation . . . . . . . . . . . . . . . . 93<br />

6.2.3 Data Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

6.3 Outlook <strong>and</strong> Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

III Additional Work 113<br />

7 Transient-Enhanced Si Diffusion 115<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

7.2 Template Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

7.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

7.4 Detailed Characterization <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . 121<br />

8 Germanium Source Reconstruction 125<br />

A Calibration <strong>and</strong> Characterization <strong>of</strong> Sources 133<br />

A.1 Calibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

A.2 Photoluminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139


viii CONTENTS<br />

B General Physical Data 145<br />

B.1 Stereographic Projection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145<br />

B.2 Physical Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146<br />

Bibliography 147<br />

Postface 159<br />

Curriculum Vitae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163


Part I<br />

Basics<br />

1


Chapter 1<br />

Silicon-Germanium Material<br />

System<br />

Within this chapter some basic facts <strong>of</strong> the <strong>SiGe</strong>-material system are covered which are<br />

essential for the underst<strong>and</strong>ing <strong>of</strong> the present work. This introductory part is reassembled<br />

from the diploma-thesis [1] <strong>and</strong> based on st<strong>and</strong>ard text books [2] <strong>and</strong> data collection volumes<br />

[3, 4, 5]. To get a more detailed insight into <strong>SiGe</strong>-heterostructures <strong>and</strong> for further reading<br />

overview articles such as Ref. [6, 7] may be considered valuable.<br />

The beginning <strong>of</strong> semiconductor studies reaches back to the early nineteenth century. One<br />

<strong>of</strong> the most studied elemental semiconductors, that are composed <strong>of</strong> single species <strong>of</strong> atoms,<br />

is silicon (Si). Like germanium (Ge) <strong>and</strong> carbon (C), silicon can be found in column IV <strong>of</strong><br />

the periodic table.<br />

Prior to the invention <strong>of</strong> the bipolar transistor in 1947, semiconductors have been used only<br />

as two-terminal devices, such as rectifiers <strong>and</strong> photodiodes. In the 1950s germanium has been<br />

the most used semiconductor material. But due to various disadvantages <strong>of</strong> germanium, such<br />

as high leakage currents at moderately elevated temperatures, <strong>and</strong> its water soluble germa-<br />

nium oxide, germanium has widely been substituted by silicon in the 1960s.<br />

There are several convincing reasons for the use <strong>of</strong> Si. Silicon <strong>of</strong>fers – compared with ger-<br />

manium – much lower leakage currents <strong>and</strong> a stable natural oxide. Thermally grown silicon<br />

dioxide can be achieved with high quality. Another reason is that device-grade silicon is much<br />

cheaper than any other semiconductor material. Silicon in the form <strong>of</strong> silica <strong>and</strong> silicates is<br />

next to oxygen the most widespread element in the earth crust.<br />

Many <strong>of</strong> the compound semiconductors have electrical <strong>and</strong> optical properties that surpass the<br />

properties <strong>of</strong> silicon or are even absent in silicon. These semiconductors, especially gallium ar-<br />

3


4 CHAPTER 1. SILICON-GERMANIUM MATERIAL SYSTEM<br />

senide (GaAs) <strong>and</strong> heterostructures based on it, are used mainly for microwave <strong>and</strong> photonic<br />

applications. Nevertheless silicon is one <strong>of</strong> the most investigated elements in the periodic<br />

table <strong>and</strong> the whole silicon technology is by far the most advanced among all semiconductor<br />

technologies. [2]<br />

1.1 Structural Properties<br />

Silicon <strong>and</strong> germanium crystallize in a diamond lattice structure. This structure belongs to<br />

the cubic-crystal family <strong>and</strong> can be seen as two interpenetrating fcc sublattices with one sub-<br />

lattice displaced from the other by one quarter <strong>of</strong> the distance along the diagonal <strong>of</strong> the cube<br />

(i.e. a displacement <strong>of</strong> √ 3/4 along [111]). In a diamond lattice all atoms are identical, <strong>and</strong><br />

each atom in the diamond lattice is surrounded by four equidistant nearest neighbors that<br />

lie at the corners <strong>of</strong> a tetrahedron (see Fig. 1.1, refer to the spheres connected with darkened<br />

bars [2]).<br />

In a hard sphere model only 34% <strong>of</strong> the available space is filled, <strong>and</strong> therefore the diamond<br />

lattice structure is not very compact. Si <strong>and</strong> Ge have, like any other group IV element,<br />

four electrons in the outer orbit, <strong>and</strong> each atom shares these valence electrons with its four<br />

neighbors. This sharing <strong>of</strong> electrons is called covalent bonding <strong>and</strong> occurs between atoms <strong>of</strong><br />

either the same element or between atoms <strong>of</strong> different elements that have similar outer-shell<br />

electron configurations (in our case: Si, Ge, C). [2].<br />

Figure 1.1: Diamond lattice structure <strong>of</strong> Si <strong>and</strong> Ge [2].<br />

� Source: <strong>SiGe</strong> crystalstructure.jpg


1.1. STRUCTURAL PROPERTIES 5<br />

Figure 1.2: Schematical picture <strong>of</strong> a (211)-crystal plane [2].<br />

� Source: <strong>SiGe</strong> crystalplane.jpg<br />

1.1.1 Crystal Orientation <strong>and</strong> Miller Indices<br />

The usual method <strong>of</strong> defining the different crystal planes is to use Miller indices. This is<br />

achieved in the following way, demonstrated with an example: As one can see in Fig. 1.2<br />

[2] the depicted plane intercepts the Cartesian coordinates (x, y, z) at a, 2a, 2a. Taking<br />

the reciprocals <strong>of</strong> the intercepting numbers (in terms <strong>of</strong> the lattice constant) yields 1, 1 1<br />

2 , 2 .<br />

Multiplying with 2 gives the smallest three integers <strong>and</strong> the plane is written in Miller indices<br />

notation as (211)-plane.<br />

Some conventions for the Miller indices notation are:<br />

(hkl) : For the plane that intercepts the x-axis on the negative side <strong>of</strong> the origin, such as<br />

(100), (110).<br />

Figure 1.3: Miller indices <strong>of</strong> most important planes in a cubic crystal [2].<br />

� Source: <strong>SiGe</strong> Miller-indices.jpg


6 CHAPTER 1. SILICON-GERMANIUM MATERIAL SYSTEM<br />

{hkl} : For planes <strong>of</strong> equivalent symmetry – such as {100} st<strong>and</strong>s for (100), (010), (001),<br />

(100), (010), (001) in cubic symmetry.<br />

[hkl] : For a crystal direction, such as [100] for the x-axis-direction. Therefore the [hkl]-<br />

direction is perpendicular to the (hkl)-plane.<br />

〈hkl〉 : As for the planes for a full set <strong>of</strong> equivalent directions – such as 〈100〉 st<strong>and</strong>s for [100],<br />

[010], [001], [100], [010], [001].<br />

In Fig. 1.3 [2] there are some important planes in cubic crystals depicted. In a Si-crystal<br />

the {111}-surface is the energetically favored surface-plane.<br />

1.2 Vicinal Si(001)<br />

Generally Si-surfaces are never flat on a long scale. There are always steps even at nearly<br />

singular or low-index surfaces. These tilted surfaces are named ”vicinal” as the average<br />

surface orientation is usually close to the respective low-index surface <strong>and</strong> deviates only<br />

by a small miscut angles from the specific crystal-direction (Fig. 1.4). This is not always<br />

unwanted as a finite miscut provides several steps which enables ”step-flow” growth even<br />

at lower temperatures due to the large number <strong>of</strong> favored nucleation sites (”kinks”): the<br />

steps proceed by incorporation <strong>of</strong> adatoms at these kink sites along the step edges (red<br />

circle in Fig. 1.4) whereas on-terrace nucleation events are suppressed (blue circle in Fig. 1.4).<br />

Figure 1.4: Schematic visualization <strong>of</strong> vicinal Si(001) surface with a miscut angle<br />

between the nominal (001)-crystal direction <strong>and</strong> the average surface orientation.<br />

The high step-edge density <strong>of</strong>fers favored nucleation sites in form <strong>of</strong> steps <strong>and</strong><br />

”kinks” on steps.<br />

� Source: <strong>SiGe</strong> miscut.jpg


1.2. VICINAL SI(001) 7<br />

Vicinal substrates are in principal available with any miscut angle α <strong>and</strong> arbitrary azimuthal<br />

orientation φ. Commercial use is actually only made <strong>of</strong> Si(001) with a miscut <strong>of</strong> up to<br />

± 4 ◦ . [8]<br />

1.2.1 Step Structure<br />

Terraces on vicinal Si(001) which are separated by an odd number <strong>of</strong> mono-atomic steps show<br />

dimer rows which are oriented perpendicular to each other. These dimer rows are (2×1) <strong>and</strong><br />

(1×2) reconstructed domains which are oriented along the [110] <strong>and</strong> [110] direction. This<br />

reconstruction reduces the number <strong>of</strong> dangling bonds <strong>and</strong> thus lowers the surface energy.<br />

There are two different types <strong>of</strong> step segments with mono-atomic height: the upper-terrace<br />

is either reconstructed with dimer rows parallel to the step edge (SA) or perpendicular to it<br />

(SB) [9]. Step edges with arbitrary direction according to their different azimuthal miscut<br />

orientations are assembled from such segments. For SB-steps two kinds have to be distin-<br />

guished depending on the registry <strong>of</strong> the step edge with the dimer row reconstruction on the<br />

lower-terrace. For rebonded SB step edges the number <strong>of</strong> dangling bonds is reduced which<br />

makes them energetically favorable compared to non-rebonded SB-steps. Double steps are<br />

observed primarily in form <strong>of</strong> DB-steps <strong>and</strong> here again mainly in the rebonded type [10, 11].<br />

They separate terraces which expose dimer rows running perpendicular to the DB step edges.<br />

Figure 1.5: Schematic model <strong>of</strong> the most common step types on Si(001) showing<br />

SB <strong>and</strong> DB steps in their rebonded configuration [8].<br />

� Source: <strong>SiGe</strong> vicinal-surface dimers.jpg


8 CHAPTER 1. SILICON-GERMANIUM MATERIAL SYSTEM<br />

DA double steps have a by far subordinate abundance. Fig. 1.5 (after [8]) depicts a schematic<br />

model <strong>of</strong> the most common step types on Si(001) showing SB <strong>and</strong> DB steps in their rebonded<br />

configuration. [8].<br />

1.2.2 Step Types in Different Miscut Angle Regimes<br />

For small miscuts below 1 ◦ only mono-atomic steps with a height <strong>of</strong> 0.136 nm (i.e. aSi/4)<br />

are found. Both single-step types are present but reveal a different nature. Whereas the<br />

SA-steps are rather straight <strong>and</strong> smooth, the SB-steps appear rugged. There are many SA-<br />

kinks in SB-steps <strong>and</strong> few SB-kinks in SA-steps as it costs less energy to create an SA step<br />

segment [11].<br />

Fig. 1.6 [12] shows an STM image (∼ 350 ˚A × 350 ˚A) <strong>of</strong> such a typical vicinal Si(001) surface<br />

with a low miscut <strong>of</strong> 0.66 ◦ along [110]. Equally spaced A- (orange) <strong>and</strong> B-terraces (violet) can<br />

be clearly discerned. The regular (2×1)-reconstruction is revealed as dimer-rows spanning<br />

along the 〈110〉-directions over the terraces with an altered orientation parallel (A-terrace)<br />

<strong>and</strong> perpendicular (B-terrace) to the step-edges. Thus the B-terraces are defined by rugged<br />

SB-steps at the lower side <strong>and</strong> rather straight SA-steps at the upper boundary.<br />

Figure 1.6: STM image (∼ 350 ˚A × 350 ˚A) <strong>of</strong> a vicinal Si(001) surface with a<br />

miscut <strong>of</strong> 0.66 ◦ along [110] revealing equally spaced A- (orange) <strong>and</strong> B-terraces<br />

(violet). The regular (2×1)-reconstruction is revealed as dimer-rows spanning along<br />

the 〈110〉-directions over the terraces with an altered orientation parallel (A-terrace)<br />

<strong>and</strong> perpendicular (B-terrace) to the step-edges. Thus the B-terraces are defined<br />

by rugged SB-steps at the lower side <strong>and</strong> rather straight SA-steps at the upper<br />

boundary. [12]<br />

� Source: <strong>SiGe</strong> STM stepped-surface.jpg


1.3. ELECTRONIC PROPERTIES 9<br />

Monoatomic height steps separate surface stress domains due to the anisotropic in-plane<br />

stress linked to the dimerization <strong>of</strong> the terraces [13, 14]. Especially for low miscuts (< 0.03 ◦ )<br />

a slightly wavy surface morphology is established which shrinks the size <strong>of</strong> stress domains<br />

by introducing reverse steps <strong>and</strong> increasing the step density over the necessary value to<br />

accommodate the miscut [15, 16, 17]. Therefore the local miscut partially exceeds the overall<br />

”averaged” substrate miscut.<br />

For miscut angles exceeding 1 ◦ along [110] the evolution <strong>of</strong> double-atomic steps sets in<br />

<strong>and</strong> the fraction <strong>of</strong> these mainly DB-steps is increased with increasing miscut (between 1 ◦<br />

<strong>and</strong> 10 ◦ ) [18]. The formation <strong>of</strong> the surface structure is a complex interplay <strong>of</strong> strain fields<br />

associated with step rebonding, step-step interactions, <strong>and</strong> energy differences between an<br />

SA-SB step pair <strong>and</strong> rebonded DB-steps [19].<br />

In the current work the main emphasis was directed to vicinal Si(001) substrates with a<br />

miscut <strong>of</strong> 4 ◦ along [110]. Hence the average terrace lengths can be estimated with 1.94 nm<br />

(3.88 nm) assuming purely mono-atomic (di-atomic) surface steps. [8].<br />

1.3 Electronic Properties<br />

1.3.1 B<strong>and</strong> Structure<br />

Silicon is like Germanium <strong>and</strong> Carbon (Diamond) an indirect semiconductor, which means<br />

that the lowest minima <strong>of</strong> the conduction b<strong>and</strong> <strong>and</strong> the highest maxima in the valence b<strong>and</strong><br />

are not located at the same position in � k-space. In Fig. 1.7 [12] the Brillouin zone <strong>of</strong> the<br />

diamond lattice structure is depicted.<br />

Figure 1.7: Brillouin zone <strong>of</strong> the fcc lattice [12].<br />

� Source: <strong>SiGe</strong> Brillouin-zone.jpg


10 CHAPTER 1. SILICON-GERMANIUM MATERIAL SYSTEM<br />

Silicon (Si)<br />

Figure 1.8: B<strong>and</strong> structures for the most prominent group IV semiconductors<br />

Si (a) <strong>and</strong> Ge (b), both featuring an indirect b<strong>and</strong>gap [12, 20].<br />

� Source: <strong>SiGe</strong> b<strong>and</strong>structure.jpg<br />

In Si the conduction b<strong>and</strong> is characterized by six equivalent minima along the 〈100〉-directions<br />

<strong>of</strong> the Brillouin zone located at about k0 = 0.85 (2π/a). The constant energy surfaces are<br />

ellipsoids <strong>of</strong> revolution with major axes along 〈100〉. The valence b<strong>and</strong> <strong>of</strong> Si has its minimum<br />

at the Γ-point where the warped heavy <strong>and</strong> light hole b<strong>and</strong>s are degenerate. The indirect<br />

gap energy Eg,ind is 1.12 eV (300 K). In Fig. 1.8 [12, 20] the Si b<strong>and</strong> structure <strong>and</strong> the indirect<br />

b<strong>and</strong> gap are shown. [3]<br />

Germanium (Ge)<br />

In Ge the conduction b<strong>and</strong> is characterized by eight equivalent minima at the end points <strong>of</strong><br />

the 〈111〉-directions <strong>of</strong> the Brillouin zone <strong>and</strong> the constant energy surfaces are ellipsoids <strong>of</strong><br />

revolution with major axes along 〈111〉. The valence b<strong>and</strong> <strong>of</strong> Ge has its minimum, like Si, at<br />

the Γ point. The indirect gap energy <strong>of</strong> Ge is with Eg,ind = 0.66 eV (291 K) smaller than in<br />

Si (see also Fig. 1.8 [12, 20] for the Ge b<strong>and</strong> structure).<br />

In Tab. 1.1 some important properties for Group IV elements are listed. For the sake <strong>of</strong><br />

completeness additionally to Si <strong>and</strong> Ge also values for carbon (diamond) are tabulated. [3]


1.3. ELECTRONIC PROPERTIES 11<br />

Silicon (Si)<br />

Germanium<br />

(Ge)<br />

atomic number 14 32 6<br />

Carbon (C)<br />

diamond<br />

relative atomic mass [amu] 28.0855(3) [21] 72.64(1) [21] 12.0107(8) [21]<br />

density d [gcm −3 ] 2.329002 5.3234 3.51525 [22]<br />

(25 ◦ C) [23] (298 K) [24]<br />

lattice parameter a [˚A] 5.43102018(34) 5.6579060 3.56683(1)<br />

(295.7 K) [25] (298.15 K) [26] (298 K) [27]<br />

relative lattice mismatch f [%] 0 +4.2 -34.3<br />

indirect energy gap Eg,ind [eV] 1.1242 0.664 5.50(5)<br />

(300 K) [28] (291 K) [29] (RT) [30]<br />

direct energy gap Eg,dir [eV] 4.135 0.805(1) 6.5 [31]<br />

effective electron masses:<br />

(190 K) [32] (293 K) [33]<br />

me,|| 0.1905(1) 0.0807(8) 1.4<br />

me,⊥ 0.9163(4) 1.57(3) 0.36 [34]<br />

(in units <strong>of</strong> m0)<br />

effective hole masses:<br />

(1.26 K) [35] (30...100 K) [36]<br />

mhh 0.537 0.284(1) † 1.08<br />

mlh 0.153 0.0438(3) † 0.36 [37]<br />

(4.2 K) [38] (4 K) [39]<br />

mso 0.234 0.095(7) 0.15 [37]<br />

(in units <strong>of</strong> m0) (4.2 K) [38] (30 K) [40]<br />

electron mobility µe 1450 3900 ≈ 2000<br />

[cm 2 V −1 s −1 ] (300 K) [41] (300 K) [42] (RT) [34]<br />

hole mobility µh 505 1800 2100<br />

[cm 2 V −1 s −1 ] (300 K) [43] (300 K) [44] (RT) [45]<br />

melting point Tm [K] 1685(2) [46] 1211.4 [21] 4100 [47]<br />

→ Tm [ ◦C] 1412 938.3 3800 (subl.)<br />

thermal conductivity κ 130 [48] 60 [49] 7 × 105 [50]<br />

@ 300 K [Wm −1 K −1 ]<br />

Table 1.1: Properties <strong>of</strong> frequently used group IV elements [3, 21].<br />

† constant energy surfaces for the VB are represented by ”warped” spheres; hence the effective masses mh<br />

depend on the crystal direction – values are listed for B parallel to [100] only


12 CHAPTER 1. SILICON-GERMANIUM MATERIAL SYSTEM<br />

Figure 1.9: Lattice parameters <strong>of</strong> Si1−xGex alloys <strong>and</strong> deviation from Vegard’s<br />

rule [51].<br />

� Source: <strong>SiGe</strong> Vegards-rule.jpg<br />

1.4 Silicon Germanium Alloys (Si1−xGex)<br />

Silicon <strong>and</strong> germanium form a continuous series <strong>of</strong> solid solutions Si1−xGex providing a grad-<br />

ual change in some parameters (e.g. lattice parameter a) with x ranging from 0 to 1 [3].<br />

The lattice parameter a<strong>SiGe</strong> <strong>of</strong> Si1−xGex alloys shows a small deviation from Vegard’s rule,<br />

i.e. from the linearity between the lattice constants <strong>of</strong> pure Si <strong>and</strong> Ge. According to ex-<br />

periments by Dismukes et al. [51] (see Fig. 1.9) <strong>and</strong> theoretical Monte Carlo simulations <strong>of</strong><br />

Si1−xGex alloys the lattice constant for Si1−xGex alloys a<strong>SiGe</strong> lies below Vegard’s rule [5].<br />

1.4.1 B<strong>and</strong>structure <strong>of</strong> <strong>SiGe</strong>-Heterostructures<br />

The lattice mismatch (4.2%) strain between Si <strong>and</strong> Ge enables b<strong>and</strong> engineering for <strong>SiGe</strong>-<br />

heterostructures. Due to the different symmetries in Si- <strong>and</strong> Ge-b<strong>and</strong> structure the <strong>SiGe</strong>-<br />

alloy b<strong>and</strong>structure shows a cross-over in the lowest conduction b<strong>and</strong> edge from Si-like [100]-<br />

symmetry to Ge-like [111]-symmetry at x ∼ 0.85 [3]. Fig. 1.10 [6, 52] shows the compositional<br />

dependence <strong>of</strong> the indirect energy gap for unstrained Si1−xGex alloys with the change from<br />

the conduction b<strong>and</strong> (CB) minimum at the ”∆”-point (Si-like) to the L-point (Ge-like) (upper<br />

curve in Fig. 1.10). The two lower curves show the vast influence <strong>of</strong> strain for pseudomorphic<br />

<strong>SiGe</strong>-layers on a Si-substrate. The in-plane compressively strained <strong>SiGe</strong>-layers show a signif-<br />

icantly lowered b<strong>and</strong>gap with a splitting <strong>of</strong> the valence b<strong>and</strong> (VB) maximum in heavy-hole<br />

(HH) <strong>and</strong> light-hole (LH) b<strong>and</strong>. [6, 52]


1.4. SILICON GERMANIUM ALLOYS (SI1−XGEX) 13<br />

Figure 1.10: Compositional dependence <strong>of</strong> the indirect energy gap for Si1−xGex<br />

alloys. The upper curve reveals the change from the Si-like (CB-minimum at the<br />

”∆”-point) to Ge-like (CB-minimum at the L-point) for unstrained <strong>SiGe</strong>-bulk material<br />

<strong>and</strong> x ∼ 0.85. The in-plane compressively strained <strong>SiGe</strong>-layers on a Si-substrate<br />

show a significantly lowered b<strong>and</strong>gap with a splitting <strong>of</strong> the valence b<strong>and</strong> (VB)<br />

maximum in heavy-hole (HH) <strong>and</strong> light-hole (LH) b<strong>and</strong> [6, 52].<br />

� Source: <strong>SiGe</strong> b<strong>and</strong>gap.jpg<br />

Fig. 1.11 (after [7]) depicts the strain-induced b<strong>and</strong> modification <strong>of</strong> a <strong>SiGe</strong>-epilayer on a<br />

Si-substrate <strong>and</strong> a strained Si-epilayer on a relaxed <strong>SiGe</strong>-substrate. The six-fold degenerate<br />

conduction b<strong>and</strong> is split into two groups <strong>of</strong> two-fold degenerate ∆(2) <strong>and</strong> four-fold degenerate<br />

∆(4)-b<strong>and</strong>s. The degenerate heavy-hole <strong>and</strong> light-hole b<strong>and</strong>s are also split with increasing<br />

strain. Whereas a compressively strained <strong>SiGe</strong>-epilayer can be used to confine hole-type<br />

carriers (HH), a tensilely strained Si-channel provides confinement for high mobility ∆(2)<br />

electrons. The second row <strong>of</strong> images in Fig. 1.11 depicts the constant energy surfaces for the<br />

conduction b<strong>and</strong> <strong>of</strong> strained epilayers. There are six ellipsoids <strong>of</strong> revolution with the longi-<br />

tudinal effective mass m e,|| for the symmetry axis along the X-direction <strong>and</strong> the transversal<br />

effective mass me,⊥ perpendicular to that. Along with the strain-induced change in b<strong>and</strong>-<br />

structure also the effective masses are influenced. [7]


14 CHAPTER 1. SILICON-GERMANIUM MATERIAL SYSTEM<br />

Figure 1.11: <strong>Strain</strong>-induced b<strong>and</strong> modification <strong>of</strong> a <strong>SiGe</strong>-epilayer on a Sisubstrate<br />

<strong>and</strong> a strained Si-epilayer on a relaxed <strong>SiGe</strong>-substrate. Whereas a compressively<br />

strained <strong>SiGe</strong>-epilayer can be used to confine hole-type carriers (HH), a<br />

tensilely strained Si-channel provides confinement for high mobility ∆(2) electrons.<br />

(after [7])<br />

� Source: <strong>SiGe</strong> strained layers Si-<strong>SiGe</strong>.jpg


Chapter 2<br />

Molecular Beam Epitaxy (MBE)<br />

This chapter is mainly based <strong>and</strong> transferred from the preceding diploma-thesis [1]. Compa-<br />

rable descriptions <strong>and</strong> contents are also found from the preworkers [12, 53, 54].<br />

The first successful use <strong>of</strong> a molecular beam apparatus for the crystallization <strong>and</strong> inves-<br />

tigation <strong>of</strong> GaAs epilayers by Cho <strong>and</strong> Arthur dates back to the late 1960s [55]. Since then<br />

ultra-high-vacuum epitaxial growth techniques developed rapidly <strong>and</strong> nowadays molecular<br />

beam epitaxy is a powerful means <strong>of</strong> growing layers <strong>and</strong> films with high purity <strong>and</strong> preci-<br />

sion. MBE provides several key advantages over CVD (Chemical Vapour Deposition), LPE<br />

(Liquid Phase Epitaxy) or MOVPE (Metal-Organic Vapour Phase Epitaxy), such as the abil-<br />

ity to control growth reproducibly to atomic monolayer dimensions <strong>and</strong> even the ability to<br />

monitor <strong>and</strong> study the growth process itself via RHEED (Reflection High Energy Electron<br />

Diffraction), XPD (x-ray Photoelectron Diffraction), AES (Auger Electron Spectroscopy)<br />

<strong>and</strong> ellipsometry. It is a far <strong>of</strong>f equilibrium deposition technique <strong>and</strong> features growth with<br />

arbitrary supersaturation. MBE also provides the possibility to control the composition <strong>and</strong><br />

doping <strong>of</strong> the grown structures <strong>and</strong> yields material with impurity levels below ten parts per<br />

billion [56].<br />

2.1 MBE-Growth<br />

2.1.1 Fundamentals <strong>and</strong> Prerequisites<br />

MBE-grown thin films crystallize via reactions between the impinging atomic or molecular<br />

beams <strong>of</strong> the constituent elements <strong>and</strong> the substrate that is kept at an elevated tempera-<br />

15


16 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

Figure 2.1: The relationship between the fundamental units encountered in vacuum<br />

technology [56].<br />

� Source: MBE vacuum.jpg<br />

ture in ultra-high-vacuum (UHV). Total pressures <strong>of</strong> the residual gas in the reactor below<br />

p ≤ 1.33 × 10 −7 Pa (10 −9 Torr) are called UHV. The composition <strong>of</strong> the grown epilayer <strong>and</strong><br />

its doping level is adjusted via the relative arrival rates <strong>of</strong> the different constituents by con-<br />

trolling the beam fluxes <strong>of</strong> the various sources. Usual growth rates are in the range <strong>of</strong> about<br />

1.5 monolayer (ML) per second (i.e. ∼ 1.5 ˚A/s). This moderate growth rate ensures sufficient<br />

surface migration <strong>of</strong> the impinging particles <strong>and</strong> therefore provide a smooth surface (also<br />

depending on the substrate temperature TS; higher TS enhances surface mobility). With the<br />

help <strong>of</strong> simple mechanical shutters in front <strong>of</strong> each <strong>of</strong> the beam sources the whole growth<br />

procedure or just the deposition <strong>of</strong> single constituents or dopants can be started, stopped<br />

or interrupted; that guarantees sharp borders or changes in composition <strong>and</strong> doping on an<br />

atomic scale, at least as long as segregation effects are thermally suppressed.<br />

In order to preserve one <strong>of</strong> the major characteristic features <strong>of</strong> MBE-growth – that is<br />

the beam nature <strong>of</strong> the mass flow towards the substrate – it is an indispensible requirement<br />

that ultra-high vacuum conditions are ensured. There are another two parameters closely<br />

related to the pressure p <strong>of</strong> the residual gas in the growth-chamber, namely the mean free<br />

path lmfp <strong>and</strong> the concentration nconc <strong>of</strong> the gas molecules travelling through the volume<br />

towards the target-substrate. Mean free path is the average distance <strong>of</strong> a molecule between<br />

two successive collisions, whereas the concentration is simply the number <strong>of</strong> species per unit<br />

volume. Large values for lmfp are not only necessary to avoid high scattering rates (that<br />

would destroy the beam-like nature), but also to minimize atoms from the residual gas to


2.1. MBE-GROWTH 17<br />

hit the substrate surface <strong>and</strong> get incorporated into the epilayer. These contaminants <strong>and</strong><br />

impurities would accumulate lattice imperfections or unintentional impurity levels, <strong>and</strong> could<br />

therefore also cause problems in the sensitive crystallization-process. All this would end up in<br />

unpredictable epilayers <strong>of</strong> low quality <strong>and</strong> usefulness. In Fig. 2.1 [56] the relationship between<br />

fundamental units in vacuum technology are compared.<br />

2.1.2 Physical Processes in the Growth-Chamber<br />

The growth process in a MBE-chamber can be divided into three zones, where different<br />

physical reactions take place [57]:<br />

1. In the first zone the molecular beams are generated under UHV conditions from electron-<br />

beam evaporators (for Si, Ge, partly for C) <strong>and</strong> sources <strong>of</strong> the Knudsen-type effusion<br />

cells (mostly used for dopants in <strong>SiGe</strong>C-MBE-technology). The temperatures <strong>of</strong> the<br />

effusion-cells are accurately controlled using proportional-integral-derivative (PID) con-<br />

trollers together with thermocouples for a feedback loop to enable flux-stabilization<br />

better than ± 1% [58]. By choosing appropriate temperatures for the effusion-cells <strong>and</strong><br />

the substrate, the desired structural composition <strong>of</strong> the epilayers can be realized.<br />

2. The second zone in a MBE vacuum reactor is the ”mixing” region in which the different<br />

molecular beams – originated at various sources – interpenetrate each other <strong>and</strong> form<br />

a more or less uniform beam. In case <strong>of</strong> a large enough mean free path there should<br />

not occur many collisions between the different species <strong>of</strong> molecules traversing towards<br />

the substrate, <strong>and</strong> interactions among these particles should be negligible.<br />

3. The actual epitaxial growth process takes place on the substrate surface that forms the<br />

third zone. Here is a series <strong>of</strong> surface processes that can be distinguished:<br />

(a) Adsorption <strong>of</strong> the incoming molecules or constituent atoms impinging on the sub-<br />

strate surface<br />

(b) Surface migration <strong>and</strong> diffusion on the substrate<br />

(c) Incorporation <strong>of</strong> the adsorbed species into the crystal lattice <strong>of</strong> the substrate or<br />

grown epilayers<br />

(d) Thermal desorption <strong>of</strong> atoms that have not been incorporated into the lattice<br />

In Fig. 2.2 [12] the most important surface processes are illustrated.<br />

Atoms or molecules arriving at the substrate surface have an energy corresponding to the


18 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

Figure 2.2: Schematic illustration <strong>of</strong> surface processes occurring during MBEgrowth<br />

[12].<br />

� Source: MBE surface-kinetics.jpg<br />

temperature <strong>of</strong> the region <strong>of</strong> their origin (source temperature). This initial temperature Ti is<br />

usually higher than the substrate temperature TS; hence they have to loose energy via energy<br />

exchange with the substrate until thermodynamic equilibrium at TS is reached. When the<br />

impinging species stays adsorbed due to a lack <strong>of</strong> energy for desorption (re-evaporation into<br />

vacuum) the particles diffuse along the surface in order to find an energetically favored lattice-<br />

site, such as a kink-position, where one half <strong>of</strong> the bonds is occupied. Other possibilities for<br />

reactions on the surface are inter-diffusion (two atoms exchange sites) or nucleation processes<br />

that appear, when migrating atoms aggregate <strong>and</strong> form a new isl<strong>and</strong> on a flat part <strong>of</strong> the<br />

substrate surface. [55]<br />

2.1.3 Growth-Modes<br />

As mentioned at the end <strong>of</strong> the last section, atoms preferentially become incorporated at kink-<br />

sites. Incorporation itself has been experimentally documented [59] <strong>and</strong> can be understood<br />

as a two-step condensation process in which the chemisorbed state is reached via a precursor<br />

physisorbed phase [60]. According to the model in Ref. [61, 62, 63] the atom as physisorbed<br />

species is allowed to diffuse over the surface with a rate constant kdiff to find an energetically<br />

favored site. The interaction potentials as seen by an atom approaching the surface are<br />

schematically depicted in Fig. 2.3 [55, 64]. It is evident from Fig. 2.3 that the physisorbed


2.1. MBE-GROWTH 19<br />

Figure 2.3: The interaction potential due to the surface, as seen by an atom<br />

impinging perpendicularly to the surface for chemisorbed (curve 1) <strong>and</strong> physisorbed<br />

precursor (curve 2) states [55, 64].<br />

� Source: MBE physi-chemisorb.jpg<br />

atom has to overcome a lower barrier to become chemisorbed at the surface, than to evaporate<br />

back into the vacuum because Ea < Edp [55].<br />

There are three growth-modes in crystal-growth that may be distinguished (see Fig. 2.4<br />

[55, 65, 66]). In the isl<strong>and</strong> growth-mode (Vollmer-Weber) small clusters nucleate directly on<br />

the substrate surface <strong>and</strong> build the base to isl<strong>and</strong>s <strong>of</strong> the condensed phase. This growth type<br />

appears when the molecules for deposition are more strongly bound to each other than to<br />

Figure 2.4: Schematic illustration <strong>of</strong> the three crystal-growth-modes. (a) Layerby-layer<br />

or Frank-Van der Merve; (b) layer plus isl<strong>and</strong> or Stranski-Krastanov;<br />

(c) isl<strong>and</strong> or Vollmer-Weber mode. θ represents the coverage in monolayers [55, 65].<br />

� Source: MBE growth-modes.jpg


20 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

Figure 2.5: Pseudomorphic epilayer (left) <strong>and</strong> partially plastically relaxed epilayer<br />

introducing misfit dislocations (right) [12].<br />

� Source: MBE relaxation.jpg<br />

the substrate material (e.g. in growing metals on insulators).<br />

In contrast, the layer-by-layer (Frank-Van der Merve) growth mode happens in the opposite<br />

case, i.e. when the impinging atoms are more strongly bound to the substrate than to each<br />

other. In consequence, the first atoms to condense form a complete monolayer (ML) on the<br />

surface, which is covered with ongoing deposition with a slightly less bound second layer. For<br />

the case in which – during further deposition <strong>and</strong> an increasing number <strong>of</strong> layers – the binding<br />

energy shows a monotonic decrease towards the value <strong>of</strong> the bulk crystal <strong>of</strong> the deposit, this<br />

layer-by-layer mode is obtained (e.g. <strong>of</strong>ten observed in semiconductor on semiconductor<br />

growth).<br />

The Stranski-Krastanov, or layer plus isl<strong>and</strong> growth, is an intermediate case. When the first or<br />

a few monolayers (”wetting-layers”) have been formed subsequent layer growth is energetically<br />

unfavorable <strong>and</strong> isl<strong>and</strong>s form on top <strong>of</strong> these intermediate layers. The accumulation <strong>of</strong> strain<br />

energy with increasing epilayer thickness disturbs the monotonic decrease in binding energy<br />

that is necessary for a layer-by-layer growth. Therefore the epilayer morphology gets three-<br />

dimensional to enhance stress relaxation at expense <strong>of</strong> surface energy. [55]<br />

2.1.4 Heteroepitaxy versus Homoepitaxy<br />

There are two main categories distinguished in epitaxy, namely homoepitaxy <strong>and</strong> hetero-<br />

epitaxy. In homoepitaxy the grown crystal consists <strong>of</strong> one main compound that can be<br />

additionally doped. In heteroepitaxy, which is exploited for b<strong>and</strong>structure engineering, the<br />

crystal, or parts (layers) <strong>of</strong> it, is a mixture <strong>of</strong> different compounds (e.g. Si/<strong>SiGe</strong>).<br />

In heteroepitaxy the lateral lattice parameters <strong>of</strong> the underlying substrate or epilayers are<br />

continued in pseudomorphic overgrowth (see Fig. 2.5 [12]). Due to deviating lattice parame-<br />

ters (e.g. aGe > aSi) the pseudomorphic epilayers contain tensile or compressive strain, which<br />

can have significant influence on growth. For pseudomorpic growth the most relevant pa-


2.2. CLEANING PROCEDURE 21<br />

rameter is the critical thickness tc. This tc is an equilibrium parameter that defines the film<br />

thickness at which the strain relaxation by the generation <strong>of</strong> misfit dislocations begins [67, 68].<br />

Films with thicknesses below tc cannot relax, because the elastic energy stored in the strained<br />

layer is lower than the energy associated with the local distortion around a misfit dislocation.<br />

For films thicker than tc it becomes energetically favorable for the system to form misfit<br />

dislocations in order to provide partial strain relaxation (see Fig. 2.5 [12]) <strong>of</strong> the film. Under<br />

non-equilibrium conditions there is another critical thickness parameter t ∗ c, that defines a<br />

metastable thickness range between tc <strong>and</strong> t ∗ c. In this parameter range the nucleation <strong>and</strong><br />

propagation <strong>of</strong> misfit dislocations is kinetically suppressed [69, 70]. Growth temperature<br />

has a strong influence on the maximum available critical thickness t ∗ c, as strained layers <strong>of</strong><br />

metastable thickness can partly relax during later heat treatment. In Fig. 2.6 [68] the theoret-<br />

ical critical thickness for equilibrium tc, <strong>and</strong> the experimental critical thickness t ∗ c measured<br />

with films grown by MBE at 550 ◦ C (metastable limit) on Si are shown for Si1−xGex [6].<br />

2.1.5 Relevant Temperatures in MBE<br />

Temperature affects all aspects <strong>of</strong> the epitaxial crystal quality. Surface diffusion, incorpora-<br />

tion <strong>and</strong> redistribution <strong>of</strong> impurities <strong>and</strong> lattice defects strongly depend on temperature.<br />

Crystal growth by MBE systems is a non-equilibrium process. Usually growth temperatures<br />

are far below the melting temperatures Tm <strong>of</strong> the different constituents. For Si-MBE sub-<br />

strate temperatures around 550 ◦ C are used. But there is a wide temperature range in which<br />

layer-structures are grown, depending on the intended device properties. To achieve high<br />

crystal quality <strong>and</strong> to keep the defect-density in the epilayers low, higher temperatures up<br />

to ∼ 750 ◦ C are used. To minimize diffusion <strong>and</strong> segregation <strong>of</strong> dopants even lower temper-<br />

atures down to ∼ 300 ◦ C can be <strong>of</strong> practical interest. The substrate temperature influences<br />

the whole surface kinetics <strong>and</strong> controls interface roughness <strong>and</strong> surface morphology [56].<br />

Sometimes high-temperature steps up to 1000 ◦ C <strong>and</strong> more are introduced to growth proce-<br />

dures to heal out crystal defects or to adjust doping pr<strong>of</strong>iles via diffusion. Also, after chemical<br />

cleaning procedures outside the MBE-chamber the substrates are annealed in a short-time<br />

annealing step at 900–1000 ◦ C inside the MBE-system to remove silicon-oxides.<br />

2.2 Cleaning Procedure<br />

It is obvious that a clean substrate surface is essential to make high quality epitaxial growth<br />

possible. A clean <strong>and</strong> smooth surface on an atomic scale is the requirement for perfect


22 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

Figure 2.6: Critical thickness versus Ge-content for Si1−xGex on Si [68].<br />

� Source: MBE Matthews-Blakeslee.jpg<br />

epitaxial crystal growth. In our case the Si-substrates have undergone several precursor<br />

processing steps <strong>and</strong> have been exposed to air. Thus, the Si-surface is covered with metallic<br />

or organic impurities or just the amorphous natural silicon-dioxide SiO2. On a slightly<br />

contaminated surface epitaxial growth would be inhibited or at least the properties <strong>of</strong> the<br />

epilayers – electrical, optical or structural – would be poor. Without oxide desorption the<br />

growth on Si-wafers that are covered by an amorphous SiO2 layer (natural oxide (∼ 20 ˚A) or<br />

chemical oxide caused by the cleaning procedure (RCA: ∼ 60 ˚A)) would produce amorphous<br />

Si epilayers.<br />

In the following sections some relevant cleaning procedures are discussed [53].<br />

2.2.1 Pre-Cleaning<br />

Small substrate-pieces that have been processed before (especially the etched wire substrates,<br />

Ch. 7) are successively cleaned in Trichloretylen, Acetone, Methanol using an ultrasonic<br />

bath (US), 5 min per step. After that the substrates are rinsed with deionized water (DI-<br />

H2O) <strong>and</strong> dried using the flow <strong>of</strong> a nitrogen (N2) nozzle. Additionally, the Si-pieces are<br />

cleaned with sulphuric acid (H2SO4, 96%) <strong>and</strong> hydrogen peroxide (H2O2, 30%) for 15 min at


2.2. CLEANING PROCEDURE 23<br />

H2SO4 : H2O2 = 5:1. In this acidic cleaning step organic impurities on the surface are oxidized<br />

<strong>and</strong> removed. This cleaning procedure closes with rinsing the samples in DI-H2O for another<br />

15 min.<br />

2.2.2 RCA-Cleaning<br />

The RCA-cleaning procedure is again a multi-step procedure that provides extraordinary<br />

clean surfaces <strong>and</strong> makes good crystalline quality <strong>of</strong> the epilayers possible. Although the<br />

cleaning sequence that has been developed by the Radio Corporation <strong>of</strong> America [71, 72]<br />

is time consuming, the good results (much better than with a short HF-Dip) speak for<br />

themselves.<br />

In the first part <strong>of</strong> the cleaning procedure the Si-substrates are lowered into a boiling alkaline<br />

bath (80 ◦ C) for 15 min. Subsequently the pieces are rinsed in a DI-H2O water-bath for<br />

another 15 min or at least till the resistivity extends values <strong>of</strong> 10–15 MΩcm. In the second<br />

part the substrates are immersed in an acidic mixture (80 ◦ C) for 15 min. Afterwards, in the<br />

final step, the pieces are rinsed again in the DI-H2O water-bath for 15 min to wash away<br />

residual ions <strong>of</strong> the acid. For better h<strong>and</strong>ling the pieces are kept during the whole RCA-<br />

cleaning procedure in a polypropylene basket. Small pieces (17.5 mm × 17.5 mm) are cleaned<br />

in quartz glasses st<strong>and</strong>ing in a temperature regulated water-bath. Whole 4”-Si-wafers are<br />

cleaned in suitable quartz basins <strong>and</strong> are h<strong>and</strong>led with special clamp-holders.<br />

The alkaline mixture consists <strong>of</strong> NH3 (25%), H2O2 (31%) <strong>and</strong> DI-H2O in parts <strong>of</strong> 1:1:5. HCl<br />

(30%), H2O2 (31%) <strong>and</strong> DI-H2O in parts <strong>of</strong> 1:1:5 are the ingredients <strong>of</strong> the acid. Whereas<br />

the acid removes organic impurities, the basic bath should take away metallic contamination.<br />

Although the original recipe for RCA-cleaning includes three HF-Dips (in the beginning, after<br />

the acidic <strong>and</strong> the alkaline treatment) in our ”HF-free RCA-cleaning” procedure no HF-Dip<br />

is implemented. Hence, this variation <strong>of</strong> RCA-cleaning generates a chemical oxide with a<br />

thickness <strong>of</strong> some tens <strong>of</strong> ˚Angstroms <strong>and</strong> a hydrophilic substrate surface. [53]<br />

2.2.3 HF-Dip<br />

An HF-Dip is used to remove the natural SiO2 from the Si-substrate surface immediately<br />

before the Si-pieces or Si-wafers are put into the load-lock chamber. Usually an HF-Dip<br />

is a mixture <strong>of</strong> HF (40%) <strong>and</strong> DI-H2O with a ratio between 1:5 <strong>and</strong> 1:10. The substrates<br />

are put into the HF-solution for about 30 s <strong>and</strong> dried with the N2-nozzle. When the Si has<br />

been ”dipped” long enough the liquid drips down. The so cleaned surface shows hydrophobic<br />

characteristics due to a temporary hydrogen-passivation, <strong>and</strong> inhibits new oxidation for some


24 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

time (several minutes or even up to a few hours).<br />

All that seems to be promising but hydro-carbons from the ambient air pollute the HF-acid<br />

<strong>and</strong> after an HF-Dip organic residuals stay behind. Even after high-temperature annealing<br />

in the MBE-chamber – where the organic hydro-carbons are cracked <strong>and</strong> desorb mainly<br />

from the substrate – some carbon is left, contaminates the surface <strong>and</strong> reacts with silicon to<br />

silicon-carbide (SiC) precipitates that can generate growth defects <strong>and</strong> dislocations.<br />

2.2.4 Oxide-Desorption<br />

Since SiO2 forms an amorphous layer, mono-crystalline epitaxial growth is impossible on<br />

SiO2. In order to desorb the oxide, the substrate is heated up to temperatures above 900 ◦ C.<br />

Due to the temperature dependence <strong>of</strong> the oxide-desorption rate, even higher temperatures<br />

(1035 ◦ C) are preferred <strong>and</strong> used, if possible. Although noteworthy oxide-desorption starts<br />

at about 875 ◦ C, it has been experimentally shown that for higher annealing temperatures<br />

better growth quality can be achieved [53]. It is suggested that some impurities leave the<br />

surface only at higher temperatures.<br />

Not only in the case <strong>of</strong> a precursor RCA-cleaning an oxide-desorption is necessary to remove<br />

the relatively thick chemical oxide, but even after an HF-Dip the time span can be too long<br />

until the substrate is brought into UHV. In usual oxide-desorption cycles the substrates are<br />

kept for ∼ 5 min (depending on the maximum achieved temperature within this cycle) at the<br />

high temperature.<br />

2.3 MBE-System<br />

In this work a Riber SIVA 45-chamber with a base pressure <strong>of</strong> 5 × 10 −11 mbar has been used.<br />

The MBE-system consists <strong>of</strong> the main growth chamber, a load-lock chamber for wafer loading<br />

<strong>and</strong> storage in UHV, <strong>and</strong> an additional chamber for special evaporation processes. In this<br />

<strong>SiGe</strong>C-facility wafers <strong>of</strong> 125 mm diameter as well as 100 mm-wafers <strong>and</strong> smaller pieces can<br />

be used directly or in all-Si adapters (Fig. 2.7), respectively. The latter can adapt substrates<br />

sawed to a size <strong>of</strong> 9.0 mm × 9.0 mm <strong>and</strong>/or 17.5 mm × 17.5 mm. Growth parameters are mon-<br />

itored with a Sentinel flux sensor, <strong>and</strong> controlled by a commercial s<strong>of</strong>tware via PC. [12, 53]<br />

2.3.1 Vacuum System<br />

In addition to clean substrates <strong>and</strong> pure evaporation materials, ultra-high vacuum conditions<br />

are absolutely necessary to obtain epitaxial layers <strong>of</strong> the best quality with a minimum <strong>of</strong>


2.3. MBE-SYSTEM 25<br />

Figure 2.7: All-Si adapter to accommodate two 17.5 mm × 17.5 mm substrates.<br />

� Source: MBE all-Si-adapter.jpg<br />

impurities <strong>and</strong> defects. The two small chambers are pumped with a turbo-molecular-pump<br />

<strong>and</strong> an associated rotary pump. In order to achieve the lowest possible pressures in the<br />

main growth chamber during growth, a 520 l turbo-pump, an ion getter pump <strong>and</strong> a titanium<br />

sublimator cooled with liquid nitrogen (LN2) are installed. Additionally, all sources <strong>and</strong><br />

surfaces inside the chamber becoming hot during growth are water cooled.<br />

Vacuum conditions are checked using st<strong>and</strong>ard Convectron (Pirani) <strong>and</strong> Bayard-Alpert<br />

vacuum gauges; a residual gas quadrupole mass spectrometer (HIDEN) in the growth chamber<br />

completes vacuum monitoring. [12]<br />

Figure 2.8: Schematic principle <strong>of</strong> an e-beam evaporator assembly <strong>and</strong> the flux<br />

measurement [12].<br />

� Source: MBE e-beam-evaporator.jpg


26 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

2.3.2 Evaporation <strong>and</strong> Rate Measurement<br />

In this MBE-system electron beam evaporators for silicon, germanium <strong>and</strong> carbon are used<br />

for the deposition <strong>of</strong> the main epilayer constituents. In the course <strong>of</strong> this thesis the two<br />

formerly small e-beam evaporators for Ge <strong>and</strong> C were replaced by a larger Ge-source (see<br />

Ch. 8). Effusion cells for boron (B, p-type doping) <strong>and</strong> antimony (Sb, n-type doping) are<br />

used <strong>and</strong> also a carbon sublimation cell is installed. The chamber for special processes is<br />

equipped with a C60-effusion cell <strong>and</strong> can be upgraded <strong>and</strong> supplied with additional sources,<br />

such as for manganese (Mn), in the future.<br />

The evaporation rates <strong>of</strong> the resistively heated effusion cells are temperature controlled. The<br />

temperature is measured with tungsten-rhenium thermocouples (Rh 5%/26%) mounted di-<br />

rectly below the radiativly heated crucible, <strong>and</strong> can be stabilized using feed-back loops to<br />

adjust the heating currents.<br />

Since in this work mainly epilayers consisting <strong>of</strong> silicon <strong>and</strong> germanium were grown, in the<br />

following the description is restricted to these sources. The schematic principle <strong>of</strong> an elec-<br />

tron beam evaporator is sketched in Fig. 2.8 [12]. A resistively heated hot filament emits<br />

thermionic electrons that are accelerated to an energy <strong>of</strong> 10 keV. This electron beam is de-<br />

flected by a static magnetic field following a 270 ◦ arc. The beam strikes the target material<br />

which is heated up <strong>and</strong> evaporated locally. With moderately high power (∼ 1 kW) growth<br />

rates <strong>of</strong> typically ∼ 0.4 ˚A/s for Si <strong>and</strong> ∼ 0.025 ˚A/s for Ge can be realized. In the case <strong>of</strong> Si<br />

<strong>and</strong> Ge the evaporation material is located in a crucible <strong>of</strong> pure silicon that is mounted in<br />

a water-cooled copper hearth. A water-cooled ro<strong>of</strong> (stainless steel) that is screened towards<br />

the evaporator by Si shields confines the molecular beam.<br />

The growth rates are measured using a Sentinel III controller (Leybold Inficon). The mea-<br />

suring principle is based on electron impact emission spectroscopy (EIES). This is an optical<br />

technique, where the emission intensities <strong>of</strong> element-specific atomic transitions that are exited<br />

via electron impact, are measured. The signal intensity depends on the density <strong>of</strong> atoms <strong>and</strong><br />

therefore on the flux <strong>of</strong> the evaporated specimen. Growth rates ranging from 0.01 ˚A/s up<br />

to 2 ˚A/s can be conveniently measured, but have to be calibrated by the evaluation <strong>of</strong> thick<br />

Si homoepitaxial <strong>and</strong> <strong>SiGe</strong> heteroepitaxial epilayers using x-ray diffraction (see Ch. A) <strong>and</strong><br />

step-height measuring systems (Alpha-Stepper). [12, 53]<br />

2.3.3 Heating <strong>and</strong> Temperature Measurement<br />

Directly above the substrate holder a resistively heated graphite me<strong>and</strong>er serves as thermal ra-<br />

diation source. The substrate is heated by absorbing the radiation emitted from the graphite


2.3. MBE-SYSTEM 27<br />

Figure 2.9: Schematic picture <strong>of</strong> the Riber SIVA 45 MBE system [12].<br />

� Source: MBE system.jpg<br />

heating system. In order to minimize the radial temperature distribution <strong>and</strong> to improve the<br />

temperature homogeneity over the whole wafer a peripherical silicon ring is mounted in be-<br />

tween. Two thermocouples are mounted close to the middle <strong>of</strong> the substrate-backside (one for<br />

reserve). These thermocouples have been calibrated against an additional thermocouple that<br />

has been cemented into a Si-wafer. To be flexible in measuring the substrate-temperature at<br />

any position a pyrometer has been installed. It has been calibrated against the thermocou-<br />

ple <strong>and</strong> can be used to monitor temperatures in a range between 550 ◦ C to 1050 ◦ C with a<br />

measurement spot <strong>of</strong> about 1 cm 2 .<br />

Within this MBE the substrate can be heated up to a temperature <strong>of</strong> ∼ 1050 ◦ C. The sub-<br />

strate can also be rotated <strong>and</strong> biased with voltages up to -2000 V during growth. Fig. 2.9 [12]<br />

shows schematically the MBE-system used in this work. [12, 53]<br />

2.3.4 Controlling <strong>and</strong> Monitoring<br />

Usually the sources <strong>and</strong> substrate-temperature are controlled <strong>and</strong> monitored via PC using<br />

the ”EPICAD-s<strong>of</strong>tware” (Epis<strong>of</strong>t). Using this s<strong>of</strong>tware process parameters can be changed<br />

manually or automatically according to a programmed growth sequence. Several Eurotherm<br />

controllers read in the actual values that are provided by the thermocouples <strong>and</strong> Sentinel<br />

sensors. The computer passes the programmed ideal values to the Eurotherm units which<br />

calculate the new setpoints <strong>and</strong> control the sources.<br />

The program provides convenient process control <strong>and</strong> graphical parameter visualization.<br />

It provides the possibility <strong>of</strong> setting calibration functions for temperatures <strong>and</strong> fluxes <strong>and</strong><br />

watches vacuum conditions <strong>and</strong> flow rates <strong>of</strong> the water-cooling system for an eventually


28 CHAPTER 2. MOLECULAR BEAM EPITAXY (MBE)<br />

necessary emergency stop. Single <strong>and</strong> complex epilayers can be programmed; the s<strong>of</strong>tware<br />

controls all shutters, <strong>and</strong> regulates the sources to the appropriate temperatures in order to<br />

get the desired atomic fluxes <strong>and</strong> growth rates. Growth control parameters such as actual<br />

<strong>and</strong> nominal values are automatically recorded <strong>and</strong> saved for later re-view.<br />

The growth chamber is equipped with a RHEED-system (Reflection High Energy Electron<br />

Diffraction) for growth monitoring. It consists <strong>of</strong> a RHEED gun <strong>and</strong> a fluorescence screen<br />

that can be used to study or check the surface morphology during growth. [12, 53]


Chapter 3<br />

Methods <strong>of</strong> Investigation<br />

This chapter is thought to give just a brief introduction to TEM <strong>and</strong> AFM, mainly on the<br />

basis <strong>of</strong> Ref. [73] regarding TEM, <strong>and</strong> basically summarizing the contents <strong>of</strong> Ref. [74] for the<br />

AFM related section. A more detailed overview, but still in a compact <strong>and</strong> comprehensive<br />

form are elaborated in the diploma-thesis [1].<br />

In this work various experimental techniques have been used for investigating surface mor-<br />

phology <strong>and</strong> characterizing changes between different preparation steps. As small structures<br />

in the nanometer range will be characterized, st<strong>and</strong>ard optical microscopy faces its limits<br />

<strong>and</strong> therefore other experimental means – providing higher spacial resolution – have to be<br />

applied.<br />

For studying the structures in more detail, electron (optical) microscopy – namely Scanning<br />

Electron Microscopy (SEM) <strong>and</strong> Transmission Electron Microscopy (TEM) – <strong>and</strong> Atomic<br />

Force Microscopy (AFM) are suitable tools. However, each <strong>of</strong> these different methods for<br />

characterization has several advantages <strong>and</strong> disadvantages. With TEM the highest reso-<br />

lution can be obtained <strong>and</strong> TEM images additionally reveal structures below the surface,<br />

however, this technique requires a time-consuming sample preparation <strong>and</strong> provides small<br />

effective regions for analysis (sample-thickness ≤ 100nm!). For SEM the preparative effort is<br />

smaller <strong>and</strong> for AFM measurements almost no preliminary steps are required. Since in AFM<br />

measurements the surface morphology is probed using tips with physical dimensions that<br />

cannot be neglected, AFM images have to be regarded as image <strong>of</strong> the surface in convolution<br />

with the shape <strong>of</strong> the tip.<br />

29


30 CHAPTER 3. METHODS OF INVESTIGATION<br />

3.1 Transmission Electron Microscopy (TEM)<br />

Optical microscopes are limited in image resolution due to the long wavelength <strong>of</strong> visible<br />

light. Historically, this has been the reason for the introduction <strong>of</strong> electrons into microscopy.<br />

After de Broglie’s famous equation (Eq. 3.1), in which particle momentum p <strong>and</strong> its wave-<br />

length λ are related via Planks’s constant h,<br />

λ = h<br />

p<br />

(de Broglie) → λ[nm] ∼ 1.22<br />

� E[eV ]<br />

(3.1)<br />

<strong>and</strong> from a simple transformation, high energy 200 keV-electrons (accelerated by a corre-<br />

sponding voltage <strong>of</strong> 200 kV) have a wavelength λ <strong>of</strong> about 2.5 pm (0.0025 nm), which is sub-<br />

stantially smaller than atomic diameters. This diffraction limit <strong>of</strong> resolution is out <strong>of</strong> reach,<br />

mainly due to electron lenses. These are the crucial <strong>and</strong> limiting point in electron microscopy<br />

since electromagnetic lenses are by far not perfect. Compared to glass lenses that can be pro-<br />

duced with perfect quality the best electromagnetic lenses would correspond to ”the bottom<br />

<strong>of</strong> a Coke bottle as magnifying glass” (cite Williams <strong>and</strong> Carter [73]). Typical values for the<br />

practical resolution <strong>of</strong> a TEM are 0.15–0.3 nm <strong>and</strong> therefore the maximum useful magnifica-<br />

tion in the best high-resolution TEM is about 10 6 (resolution <strong>of</strong> eye approximately 0.1 mm).<br />

For thick samples these limit <strong>of</strong> resolution cannot be obtained due to the increased energy<br />

spread ∆E resulting in chromatic aberration. The mean free path for inelastic scattering<br />

depends on the electron energy, <strong>and</strong> therefore with higher acceleration voltages good high<br />

resolution can even be achieved with relative thick samples (50 nm) [73, 75, 76].<br />

Sample preparation has to be performed carefully <strong>and</strong> is probably the most important step in<br />

order to obtain good high-resolution images. It dem<strong>and</strong>s great skill to prepare thin electron-<br />

transparent (→ 50–100 nm!) specimen over a wide area. Ideally, the volume considered for<br />

TEM-analysis should be free <strong>of</strong> defects <strong>and</strong> artifacts originated by the time consuming prepa-<br />

ration.<br />

In order to image the corrugation <strong>of</strong> buried interfaces (see Ch. 6), or the wire pr<strong>of</strong>ile <strong>of</strong> the pro-<br />

cessed Si-samples (see Ch. 7), cross-sectional specimen have to be prepared. The preparation<br />

cycle for cross-sectional (Si-)samples is outlined in Ref. [1].<br />

The specimen in this work were investigated with a JEOL-2011 FasTEM (HR type) trans-<br />

mission electron microscope [77] with an acceleration voltage <strong>of</strong> 200 kV using a LaB6-cathode.<br />

For this TEM facility a point image resolution <strong>of</strong> 0.23 nm is specified; the magnification can be<br />

chosen from 2000 to 1.5 million. Images are either recorded with a conventional photo-plate<br />

camera, or, nowadays, mainly with a Gatan CCD-camera (1 megapixel).


3.2. SCANNING ELECTRON MICROSCOPY (SEM) 31<br />

3.2 Scanning Electron Microscopy (SEM)<br />

Scanning electron microscopy can be used to study <strong>and</strong> analyze (mainly the surface <strong>of</strong>) bulk<br />

specimen. The information is derived from the interaction between the probe electrons <strong>and</strong><br />

the specimen. The focussed electron beam is deflected by scan coils <strong>and</strong> the probe is scanned<br />

across the specimen in a raster mode. Synchronous to this scanning process, the signals<br />

generated by the probe beam interaction with the specimen are detected. The recorded in-<br />

tensity modulation gives the contrast in SEM images. The scanning process can be applied<br />

to generate a magnified image <strong>of</strong> the sample.<br />

The newly commissioned LEO (ZEISS) SUPRA� 35 FESEM provides access to nanoscale<br />

resolution images without an elaborate sample preparation (specified resolution: 1.7 nm<br />

@ 15 kV). It is well-suited to characterize the surface after critical processing steps, or to<br />

investigate etched pr<strong>of</strong>iles in cross-sectional view (Sec. 7.2). An SEM combines the possibili-<br />

ties to give a fast overview over the whole specimen area <strong>and</strong> to zoom-in for a more-detailed<br />

inspection <strong>of</strong> the surface morphology. It provides a large magnification range <strong>of</strong> typically<br />

20x–500 000x).<br />

For more facts the reader is referred to the diploma-thesis (Ref. [1]) <strong>and</strong> the PhD-thesis<br />

written by T. Berer [78].<br />

3.3 Atomic Force Microscopy (AFM)<br />

Atomic Force Microscopy (AFM) is a further development on the basis <strong>of</strong> scanning probe<br />

microscopy (SPM) implemented in the mid 1980’s. The AFM is an imaging tool with a vast<br />

dynamic range, spanning the realms <strong>of</strong> optical <strong>and</strong> electron microscopes, <strong>and</strong> is operated as a<br />

surface pr<strong>of</strong>iler with unprecedented 3D-resolution [74]. In atomic force microscopy the surface<br />

<strong>of</strong> a sample is scanned with a sharp tip that is several micrometers long <strong>and</strong> has a smallest<br />

diameter <strong>of</strong> typically 10 nm. It is located on the free end <strong>of</strong> a cantilever (∼ 100–200 µm in<br />

length). Forces between the sample <strong>and</strong> the tip cause the cantilever to bend or deflect. As<br />

the tip scans across the surface, these deflections are measured with a detector <strong>and</strong> allow a<br />

computer to generate topographic maps [74].<br />

The force most commonly associated with cantilever deflection in atomic force microscopy is<br />

the interatomic van der Waals force. The dependence <strong>of</strong> this van der Waals force upon the<br />

distance between the tip <strong>and</strong> the sample surface is depicted in Fig. 3.1 [74]. Three distance<br />

regimes are labelled in Fig. 3.1: 1) Contact regime, 2) Non-contact regime (NC), <strong>and</strong>, in<br />

between, 3) Intermittent-contact regime (IC). In the so-called contact mode the tip is held


32 CHAPTER 3. METHODS OF INVESTIGATION<br />

less than a few ˚Angstroms above the sample surface, <strong>and</strong> the van der Waals force between tip<br />

<strong>and</strong> sample is repulsive. In non-contact mode the tip is held tens or hundreds <strong>of</strong> ˚Angstroms<br />

from the surface, <strong>and</strong> therefore the intermittent force is attractive due to long-range van der<br />

Waals interactions. [74]<br />

3.3.1 Scanning Modes<br />

Contact Mode<br />

In this repulsive mode the AFM tip makes s<strong>of</strong>t ”physical contact” with the sample. The<br />

tip is attached to the cantilever with a low spring constant, <strong>and</strong> therefore the contact force<br />

causes the cantilever to bend <strong>and</strong> accommodate the changes in topography, as the scanner<br />

traces the tip across the sample (or the sample under the tip).<br />

Usually, the position <strong>of</strong> the cantilever (degree <strong>of</strong> deflection) is detected with optical tech-<br />

niques. A laser beam is reflected from the back <strong>of</strong> the cantilever onto a position-sensitive<br />

photodetector (PSPD). A change in the bending <strong>of</strong> the cantilever results in a shift <strong>of</strong> the<br />

laser beam on the detector. This system is suited to resolve the vertical movement <strong>of</strong> the<br />

cantilever tip with sub-˚Angstrom resolution.<br />

The AFM can be operated either in constant-height or constant-force mode. In constant-<br />

height mode the height <strong>of</strong> the tip is fixed, <strong>and</strong> the spatial change in cantilever deflection is<br />

Figure 3.1: Dependence <strong>of</strong> interatomic force on tip-to-sample separation [74].<br />

� Source: AFM van-der-Waals.jpg


3.3. ATOMIC FORCE MICROSCOPY (AFM) 33<br />

used to generate the topographic data. In constant-force mode a feed-back loop moves the<br />

scanner up <strong>and</strong> down in z-direction, responding to the local topography, <strong>and</strong> thereby keeping<br />

the force <strong>and</strong> thus the deflection <strong>of</strong> the cantilever constant. In this case the topographic map<br />

can be directly drawn using the z-motion <strong>of</strong> the scanner as height-information.<br />

Due to the ”hard contact” between tip <strong>and</strong> sample, s<strong>of</strong>t surfaces may be deformed, tips may<br />

collect dirt or are rubbed <strong>of</strong>f <strong>and</strong> become blunt.<br />

Non-contact Mode<br />

In NC-mode, the system vibrates a stiff cantilever with an amplitude <strong>of</strong> a few tens to hundreds<br />

<strong>of</strong> ˚Angstroms near its resonant frequency (several 100 kHz). Using a sensitive AC detection<br />

scheme, the changes in resonant frequency <strong>of</strong> the cantilever are measured. Since the resonant<br />

frequency is a measure <strong>of</strong> the force gradient (i.e. derivative <strong>of</strong> the force versus distance<br />

curve), the force gradient reflects the tip-to-sample spacing [74, 79, 80]. Comparable with<br />

the constant-force mode in contact regime, a feed-back system moves the scanner up <strong>and</strong><br />

down in order to keep the resonant frequency or amplitude constant. Again this corresponds<br />

to a fixed tip-to-sample distance, <strong>and</strong> the motion <strong>of</strong> the scanner is used to generate the<br />

topographic data set (see Fig. 3.2 [81]).<br />

NC-AFM does not suffer from tip or sample degradation effects <strong>and</strong> is suited to scan even<br />

s<strong>of</strong>t samples <strong>and</strong> to keep the tips sharp.<br />

Intermittent-contact Mode<br />

IC-AFM is similar to NC-AFM, except that in this mode the vibrating cantilever tip is<br />

brought closer to the sample, so that it barely hits (”taps”) the surface. The changes in<br />

cantilever oscillation amplitude responding to the tip-to-sample separation are monitored to<br />

obtain the surface topography.<br />

This mode is usually preferred as it combines the high resolution <strong>of</strong> contact mode, <strong>and</strong> the<br />

low wear <strong>and</strong> tear <strong>of</strong> the tip in NC-mode. [74]<br />

3.3.2 AFM-Probes<br />

The AFM-tips are the crucial point in AFM as they are in contact with the surface <strong>and</strong> probe<br />

the sample. The sharpest tips that are commercially available have a tip radius as small as<br />

50 ˚A. Due to the fact, that the interaction area <strong>of</strong> tip <strong>and</strong> sample is just a fraction <strong>of</strong> the<br />

tip radius, these tips would result in a lateral resolution <strong>of</strong> at least 20 ˚A. In this case the


34 CHAPTER 3. METHODS OF INVESTIGATION<br />

Figure 3.2: Schematic view <strong>of</strong> hardware components <strong>and</strong> signal pathways in noncontact<br />

(NC) AFM-mode [81].<br />

� Source: AFM schematic-principle.jpg<br />

resolution is usually limited to the step size according to the number <strong>of</strong> image data points<br />

(e.g. 512 × 512) [82]. [74]<br />

3.3.3 Imaging Artifacts<br />

Atomic force microscope images are among the easiest-to-interpret <strong>of</strong> all images generated<br />

by any microscopy technique. With these 3-dimensional AFM images it is easy to determine<br />

whether a feature is protruding from the surface or recessed into it [74].<br />

Tip Convolution<br />

In scanning probe microscopy most imaging artifacts arise from the tip imaging, i.e. the<br />

imaging <strong>of</strong> the convolution <strong>of</strong> the sample surface <strong>and</strong> the tip. As the tips are not ideal,<br />

in case <strong>of</strong> features that are sharper than the tip, the image is dominated by the shape <strong>of</strong><br />

the tip, rather than by the true edge pr<strong>of</strong>ile <strong>of</strong> the structure. In Fig. 3.3 the origin <strong>of</strong> tip<br />

convolution is demonstrated. Many samples have features with steep sides <strong>and</strong> therefore tip<br />

imaging is a common occurrence in images. As a consequence, sidewall angles should be<br />

measured routinely in order to check, whether the imaged slope is limited to the shape <strong>of</strong> the


3.3. ATOMIC FORCE MICROSCOPY (AFM) 35<br />

Figure 3.3: a) Path <strong>of</strong> the AFM-tip proving the topography <strong>and</strong> (b) the effect <strong>of</strong><br />

tip convolution on the resulting image.<br />

� Source: AFM tip-path.jpg<br />

tip, or really represents the topography <strong>of</strong> the sample. At least the height <strong>of</strong> features can be<br />

measured <strong>and</strong> reproduced accurately as long as features are separated wide enough so that<br />

the tip touches the bottom between these features. The lateral dimensions <strong>of</strong> the imaged<br />

features provide just a maximum value, i.e. if one measures a protruding tip-imaged feature<br />

giving a width <strong>of</strong> 250 ˚A, it is clear that the feature is at most 250 ˚A wide.<br />

Feedback Artifacts<br />

SPM images are also affected by feedback artifacts, if the feedback loop is not optimized. In<br />

the case <strong>of</strong> high feedback gains the system may oscillate, generating high frequency periodic<br />

noise in the image, either throughout the image or just localized to structures with steep<br />

slopes.<br />

For feedback gains that are too low, the tip cannot track the surface well <strong>and</strong> the images<br />

loose detail <strong>and</strong> appear smooth. Another effect in the low gain case that is less obvious is<br />

”ghosting”. On sharp slopes, an overshoot can occur in the image as the tip scans up the<br />

slope, <strong>and</strong> an undershoot can occur as the tip scans down the slope. This artifact can usually<br />

be seen on steep features, imaged as bright ridges on the uphill side <strong>and</strong>/or dark shadows on<br />

the downhill side <strong>of</strong> the structure.<br />

Image Processing Capabilities<br />

Commercial SPMs are usually equipped with sophisticated image-processing s<strong>of</strong>tware, pro-<br />

viding curvature enhancement algorithms, filters for environmental noise, or giving the op-<br />

portunity for retouching areas <strong>of</strong> bad data, etc.


36 CHAPTER 3. METHODS OF INVESTIGATION<br />

Figure 3.4: For miscut Si(001) substrates the average orientation usually deviates<br />

from the specific (001)-crystal direction. a) Short terrace segments are hard to<br />

resolve by AFM which impedes accounting for the tilted surface via the miscut angle.<br />

b) The step-bunching morphology exhibits flat <strong>and</strong> wide (001)-oriented terraces<br />

which enables tilt-correction.<br />

� Source: AFM tilt-correction.jpg<br />

Although there are some benefits <strong>of</strong> image-processing s<strong>of</strong>tware, it may not be used care-<br />

lessly, as it could lead to data misrepresentation. Careless flattening could change structure<br />

curvature or could make features even disappear, as well as irresponsible filtering could add<br />

artificial structures to the image. Usually a tilted surface cannot be identified as such directly<br />

from the height-scale AFM-data. Just with additional structure-related knowledge about the<br />

orientation <strong>of</strong> intrinsic morphological features, e.g. facet angles, the correct surface orienta-<br />

tion can be deduced which provides thereafter angle measurements with meaningful values<br />

(see Fig. 3.4). [74]<br />

3.3.4 Data Types <strong>and</strong> Data Representation<br />

In this work a Digital Instruments Veeco Dimension 3100 AFM with Nanoscope IV controller<br />

[83] was employed with Olympus TESP tips [84] in the ”tapping” mode to map highly resolved<br />

surface images.<br />

There are two signals or data types that are collected by the AFM-system which are<br />

exploited for the thesis at h<strong>and</strong>.<br />

The ”height” data, which give directly the topographic information <strong>of</strong> the surface morphol-<br />

ogy, correspond to the change in piezo height (z-direction) needed to keep the amplitude <strong>of</strong><br />

the cantilever vibration constant. Thus for the acquisition <strong>of</strong> the ”height” data the feedback<br />

gains must be high enough, so that the sample surface can be tracked <strong>and</strong> the change in<br />

oscillation amplitude <strong>of</strong> the tip is minimized.<br />

The ”amplitude” data measure the change in amplitude relative to the amplitude setpoint.<br />

Using high feedback gains for collecting ”amplitude” data gives the derivative <strong>of</strong> the to-<br />

pographic height in scan-direction. Hence the ”amplitude” data are <strong>of</strong>ten also referred to


3.3. ATOMIC FORCE MICROSCOPY (AFM) 37<br />

Figure 3.5: Schematic illustration <strong>of</strong> the evaluation <strong>of</strong> AFM data for the surfaceorientation-map<br />

(SOM). a) 3D-model <strong>of</strong> a faceted wire pr<strong>of</strong>ile with surface normal<br />

vectors (”1”, ”2”, ”3”). For each surface point (black spot) the normal orientation<br />

[hkl] (here shown for ”2”) is calculated with the nearest-neighbor points (red spots)<br />

to determine the local plane (rectangle colored orange). b) The intersection points<br />

<strong>of</strong> the local normal vectors <strong>and</strong> a half-sphere seen from top as projection gives a<br />

2D-plot in polar coordinates (ρ, φ). c) Whereas ρ indicates the angle between the<br />

respective local surface normal [hkl] <strong>and</strong> the (001)-surface (growth direction), φ<br />

denotes the in-plane azimuthal angle <strong>of</strong> [hkl] with respect to [100] (seldom [110]).<br />

The intensity at each point (ρ, φ) in the surface orientation histogram is a measure <strong>of</strong><br />

the abundance <strong>of</strong> a respective surface orientation. Marker circles which correspond<br />

to specific surface angles are introduced as guidance for the characteristic facets in<br />

the <strong>SiGe</strong>-system.<br />

� Source: AFM SOM schematic.jpg<br />

as ”derivative” data or data acquired in ”derivative” mode. The ”amplitude” mode pro-<br />

vides a sensitive edge-detection technique which reveals facets with a better signal-to-noise<br />

ratio. [83]<br />

The topographic AFM data (”height” mode) can be also utilized to calculate special data<br />

representation plots to point out special morphological features or arrangements. Espe-<br />

cially useful to evaluate faceted morphologies are the surface-angle-plots (SAP) <strong>and</strong> surface-<br />

orientation-maps (SOM). These are introduced to visualize the local surface inclination in<br />

each AFM data point itself or to depict a histogram <strong>of</strong> the involved surface orientations,<br />

respectively. Fig. 3.5 demonstrates the evaluation <strong>of</strong> AFM ”height” data for the surface-


38 CHAPTER 3. METHODS OF INVESTIGATION<br />

orientation-map (SOM). For each surface point (black spot in Fig. 3.5a) the normal orien-<br />

tation [hkl] (in Fig. 3.5a shown for ”2”) is calculated with the nearest-neighbor points (red<br />

spots) to determine the local plane (rectangle colored orange). The intersection points <strong>of</strong><br />

the local normal vectors <strong>and</strong> a half-sphere seen from top as projection gives a 2D-plot in<br />

polar coordinates (ρ, φ) (Fig. 3.5b). Whereas ρ indicates the angle between the respective<br />

local surface normal [hkl] <strong>and</strong> the (001)-surface (growth direction), φ denotes the in-plane<br />

azimuthal angle <strong>of</strong> [hkl] with respect to [100] (or sporadically [110]). The intensity at each<br />

point (ρ, φ) in the surface orientation histogram is a measure <strong>of</strong> the abundance <strong>of</strong> a respective<br />

surface orientation. Marker circles which correspond to specific surface angles are introduced<br />

as guidance for the characteristic facets in the <strong>SiGe</strong>-system (Fig. 3.5c).<br />

In the SAP representation the local slope is plotted. This is defined as the angle between the<br />

normal orientation vector [hkl] <strong>and</strong> the reference orientation <strong>of</strong> Si(001), namely [001]. It is<br />

extracted again via the nearest-neighbor points in analogy to the SOM evaluation.<br />

Fig. 3.6 gives an overview <strong>of</strong> the different AFM-data types <strong>and</strong> their visualization. In this<br />

exemplary illustration a calculated model <strong>of</strong> a faceted wire-pr<strong>of</strong>ile is utilized. The shape <strong>of</strong><br />

the investigated pr<strong>of</strong>ile can be clearly seen from the 3D-AFM view (Fig. 3.6a). Fig. 3.6c de-<br />

picts a line scan along the fast scan-direction (x-direction). From this cross-sectional view the<br />

faceted nature <strong>of</strong> the wire gets obvious. Surface regions <strong>of</strong> the (001)-, {113}- <strong>and</strong> {111}-type<br />

can be easily distinguished from each other.<br />

The conventional, topographic 2D-AFM representation for which the data are usually ac-<br />

quired in ”height”-mode, deliver only a rough ”feeling” for the morphology. It provides<br />

rather qualitative information about a feature, whether it is recessed into the surrounding<br />

surface or juts out. Usually topographic 2D-AFM data do not reflect the detailed structure<br />

<strong>and</strong> shape (Fig. 3.6b).<br />

In this respect the ”amplitude” data provide a more detailed access compared to the ”height”<br />

data since the ”derivative”-signal shows high contrast <strong>and</strong> reveals sensitively changes in slope<br />

along the scan-direction (Fig. 3.6c). Usually the ”amplitude”-signal features less noise than<br />

the calculated derivative <strong>of</strong> the ”height” data, <strong>and</strong> thus can give meaningful information es-<br />

pecially about micro-facets with small lateral extensions.<br />

Fig. 3.6e shows a surface-orientation-map (SOM) with the relative abundance <strong>of</strong> the promi-<br />

nent surface angles ((001) ≡ 0 ◦ , {113} ≡ 25.2 ◦ , {111} ≡ 54.7 ◦ ). This plotting scheme is well-<br />

suited to quickly reveal special surface orientations even from slightly noisy AFM-data mea-<br />

sured in ”height”-mode. Extracting representative facet angles is <strong>of</strong>ten hardly possible by<br />

analyzing line scans. The marker circles in the SOM images correspond to surface angles ρ<br />

(see Fig. 3.5) which are known for the {113}-, {111}-, <strong>and</strong> {110}-orientation.


3.3. ATOMIC FORCE MICROSCOPY (AFM) 39<br />

Figure 3.6: Overview <strong>of</strong> different AFM-data types <strong>and</strong> their visualization illustrated<br />

for a faceted wire-pr<strong>of</strong>ile. a) 3D-AFM view <strong>and</strong> d) line scan in the fast<br />

scan-direction (x-direction). b) Compared to the conventional topographic 2D-<br />

AFM image (”height”-mode) the c) ”derivative”-mode signal reveals changes in<br />

slope along the scan-direction. e) The surface-orientation-map (SOM) shows the<br />

relative abundance <strong>of</strong> prominent surface angles <strong>and</strong> the f) surface-angle-plot (SAP)<br />

indicates quantitatively the local inclination angle for the AFM-probed area.<br />

� Source: AFM data-types.jpg<br />

The surface-angle-plot (SAP) indicates quantitatively the local inclination angle in every<br />

point separately, but again for the whole AFM-probed area (Fig. 3.6f). Such plots are again<br />

calculated from the ”height” data <strong>and</strong> evidence hence a noisier slope representation than im-<br />

ages depicting the ”derivative” signal. Nevertheless this evaluation scheme is the only chance<br />

to get quantitative information <strong>of</strong> the local slope in any direction, <strong>and</strong> not only towards the<br />

fast scan-direction (x-direction in the present case). So would a ”derivative”-mode image<br />

with the fast scan-direction exactly along the wire-pr<strong>of</strong>ile (i.e. in y-direction for the discussed<br />

case) exhibit purely uniform data without any contrast.


40 CHAPTER 3. METHODS OF INVESTIGATION


Part II<br />

Main Research<br />

41


Chapter 4<br />

<strong>Self</strong>-Organized Growth –<br />

Step-Bunching<br />

In this <strong>and</strong> the following chapter (see Ch. 5) various sub topics <strong>of</strong> self-organized growth in the<br />

Si/<strong>SiGe</strong> system are discussed. The first introductory part gives a compact summary on the<br />

preparatory work <strong>and</strong> the PhD-theses by Christoph Schelling [12] <strong>and</strong> Michael Mühlberger<br />

[54]. The results on step-bunching presented there serve as starting point for the work at h<strong>and</strong>.<br />

Transferring this know-how to high-miscut samples should yield periodic ripple templates<br />

which are feasible for self-organized growth <strong>of</strong> <strong>SiGe</strong> isl<strong>and</strong>s [85, 86, 87, 88]. Anyway, these<br />

structures on the nanometer scale open up an ample playground to study several aspects <strong>of</strong><br />

growth <strong>and</strong> especially the interplay between kinetics, surface energy <strong>and</strong> strain [88, 89, 90].<br />

The epilayer morphology encountered here is closely related to the features seen for seeded Ge<br />

dot growth on pre-structured Si templates [91, 92, 93]. Thus these experiments may help to<br />

reveal the pathway <strong>of</strong> isl<strong>and</strong> formation in the pits <strong>of</strong> 2D pre-patterned Si substrates [94, 95].<br />

In the following chapters a different notation is used for measured values <strong>and</strong> growth<br />

related parameters. Period <strong>and</strong> height readings especially encountered in AFM data analysis<br />

are listed in nanometers (SI-units), whereas for the MBE-intrinsic parameters, such as layer<br />

thickness <strong>and</strong> growth rate, the more natural Angstrom units are ought to be used consistently.<br />

4.1 Introduction to Step-Bunching<br />

In this section ”step-bunching” is briefly discussed to sketch the actual status <strong>and</strong> under-<br />

st<strong>and</strong>ing <strong>of</strong> the community for this basic growth instability. The results <strong>of</strong> the pre-workers<br />

are summarized here only in a very compact form. For a more comprehensive <strong>and</strong> detailed<br />

43


44 CHAPTER 4. SELF-ORGANIZED GROWTH – STEP-BUNCHING<br />

Figure 4.1: Typical AFM image showing step-bunching <strong>of</strong> a 3300 ˚A thick Si-buffer<br />

(Si @ 0.2 ˚A/s, 490 ◦ C) grown on a Si-substrate with 0.66 ◦ miscut along [110]. The<br />

data represent the ripple-morphology for low-miscut Si-substrates as investigated<br />

by the preworkers Christoph Schelling [12] <strong>and</strong> Michael Mühlberger [54] (data from<br />

[54]).<br />

� Source: Stepbunching AFM low-miscut preworkers.jpg<br />

view on these experimental findings the reader is referred to the PhD-theses written by<br />

Christoph Schelling [12] <strong>and</strong> Michael Mühlberger [54]. Fig. 4.1 shows a typical AFM im-<br />

age demonstrating step-bunching <strong>of</strong> a 3300 ˚A thick Si-buffer (Si @ 0.2 ˚A/s, 490 ◦ C) grown<br />

on a 0.66 ◦ [110] miscut sample. The data represent the ripple-morphology for low-miscut<br />

Si-substrates as investigated by the preworkers (data were taken from [54]).<br />

In general the evolution during growth <strong>and</strong> final morphology <strong>of</strong> an epitaxial film is governed<br />

by the interplay between thermodynamics <strong>and</strong> kinetics. The competition can be described<br />

with the concept <strong>of</strong> local equilibrium [96]. The slowest kinetic rate defines the finite range <strong>of</strong><br />

equilibrium. For extended length scales the kinetics always win over thermodynamic processes<br />

which is thus reflected in the layer morphology. The borderline can be tuned by appropri-<br />

ately choosing the growth rate, substrate temperature <strong>and</strong> miscut. For low growth rates <strong>and</strong><br />

high temperatures the equilibrium properties will dominate over a wide range, whereas for<br />

very rapid growth at low temperatures thermodynamics are spatially restricted by kinetic<br />

limitations. A high miscut provides more step edges <strong>of</strong>fering a large number <strong>of</strong> preferential<br />

nucleation sites <strong>and</strong> therefore corresponds to a higher effective substrate temperature.<br />

<strong>Kinetic</strong> growth instabilities can be explained by adatom currents on the surface during growth.<br />

Especially the behavior <strong>of</strong> an adatom encountering a downward step edge has to be consid-<br />

ered. The adatom can either jump to the lower terrace or can be reflected at the step edge<br />

which is described with the help <strong>of</strong> an Ehrlich-Schwoebel barrier [97, 98, 99]. For growth<br />

on vicinal surfaces the Schwoebel barrier has a stabilizing effect. The migrating adatoms<br />

are mainly incorporated at the upward step edge. The cross-section for impinging atoms is<br />

proportional to the terrace length which helps to shrink the respective terrace by adjusting its


4.1. INTRODUCTION TO STEP-BUNCHING 45<br />

size towards the average. For the inverse behavior <strong>and</strong> incorporation from the upper terrace<br />

the large terraces grow even further on cost <strong>of</strong> the small terraces leading to a destabilized<br />

situation. The step edges <strong>of</strong> the small shrinking terraces are finally collected in bunches<br />

which leads to the name ”step-bunching”. Instabilities can also occur parallel to the miscut<br />

direction where a slight initial waviness <strong>of</strong> the step edges is exaggerated to pronounced un-<br />

dulations due to the incorporation from the lower terrace (Bales-Zangwill instability [100]).<br />

The kinetic stability <strong>of</strong> the surface can be analyzed in terms <strong>of</strong> the surface mass current � j as<br />

a function <strong>of</strong> the local miscut �m <strong>of</strong> the surface [101, 102]. The surface morphology perpen-<br />

dicular to the miscut direction is determined by the sign <strong>of</strong> the derivative <strong>of</strong> the surface mass<br />

current � j after the miscut �m, whereas the surface parallel to the miscut direction is correlated<br />

to the sign <strong>of</strong> the current � j itself. Therefore both variables ∂� j<br />

∂ �m <strong>and</strong> � j( �m) together fully determine<br />

the overall evolution <strong>of</strong> an epitaxial surface. The different scenarios are schematically<br />

visualized in Fig. 4.2 [54]. Step-bunching (with straight step-edges) occurs for a downward<br />

current <strong>and</strong> preferential adatom incorporation from the upper terrace which makes broad<br />

terraces grow faster. [12, 54]<br />

4.1.1 Dependence on Substrate-Temperature, Growth-Rate, Layer-Thick-<br />

ness, Ge-content <strong>and</strong> Miscut<br />

Many experiments were performed to examine the vast parameter space <strong>and</strong> to check the de-<br />

pendence <strong>of</strong> the kinetics on substrate temperature, growth rate, layer thickness, Ge-content<br />

<strong>and</strong> miscut orientation on the surface morphology <strong>of</strong> step-bunching. Manifold results are<br />

reported in the literature by various groups. Not only miscut orientations close to the tech-<br />

nologically most important Si(001) surface were extensively investigated, but also further<br />

surface orientations, such as Si(111), Si(113) <strong>and</strong> Si(118), were examined for their stability<br />

against step-bunching <strong>and</strong> instabilities in general (see e.g. [103, 104, 105]).<br />

Although for vicinal substrates near thermodynamically stable facets the free surface ener-<br />

gies may strongly influence the morphology <strong>of</strong> the grown Si/<strong>SiGe</strong> epilayers, in the following<br />

the general behavior for the nearly singular Si(001) surface (close to the wafer st<strong>and</strong>ard spec-<br />

ification, i.e. < 0.5 ◦ ) is explained. This part is intended to summarize the results gathered<br />

by the preworkers for slightly miscut Si(001) (0.66 ◦ along [110], see PhD-theses [12, 54] <strong>and</strong><br />

related publications [106, 107, 108, 109, 110, 111]).<br />

The above mentioned experiments revealed that step-bunching for a miscut <strong>of</strong> 0.66 ◦ along<br />

[110] occurs only in a very small temperature window <strong>and</strong> shows its maximum around 490 ◦ C<br />

with a flux <strong>of</strong> 0.2 ˚A/s (see Fig. 4.1). For slightly higher or lower substrate temperatures the<br />

regular ripple structure fades away. Post-growth annealing even at the growth temperature


46 CHAPTER 4. SELF-ORGANIZED GROWTH – STEP-BUNCHING<br />

Figure 4.2: Stability analysis in terms <strong>of</strong> surface mass current � j( �m) [101, 102, 54].<br />

� Source: stability analysis 01.jpg<br />

generally leads to a reduction in the amplitude <strong>of</strong> the instability [106]. The steps tend to<br />

rearrange to evenly spaced monoatomic <strong>and</strong> diatomic height steps [18].<br />

The growth-rate influences the evolution <strong>of</strong> the ripples indirectly via the diffusion length.<br />

Higher rates perturb the unhindered adatom current thus giving less pronounced step-bunching.<br />

Increased flux corresponds therefore to a decreased effective substrate temperature.<br />

The third important parameter is the layer-thickness. With increasing layer-thickness t the<br />

period Λ <strong>and</strong> the peak-to-valley height H <strong>of</strong> the ripples is increased according to a power-law<br />

dependence Λ ∝ t α (α ∼ 0.25) <strong>and</strong> H ∝ t β (β ∼ 0.5), respectively. For a Si epilayer thickness<br />

<strong>of</strong> 1000 ˚A deposited at 490 ◦ C <strong>and</strong> a rate <strong>of</strong> 0.2 ˚A/s the measured values read Λ ∼ 450 nm for<br />

the period <strong>and</strong> H ∼ 1.8 nm for the ripple height.<br />

The miscut vector plays a very important role. It can be separated into the inclination angle<br />

θ with respect to the singular surface <strong>and</strong> into φ, which is the azimuthal angle measured rel-<br />

ative to a specific in-plane direction. The miscut angle θ defines the average terrace length,<br />

which is an important quantity considering step-flow growth, <strong>and</strong> thus has to be seen in the<br />

context <strong>of</strong> the diffusion length. Thus, it is not surprising that an increased miscut results in<br />

shorter terraces <strong>and</strong> shifts the maximum <strong>of</strong> pronounced step-bunching to lower temperatures.


4.1. INTRODUCTION TO STEP-BUNCHING 47<br />

The ripple period does not depend on the azimuthal miscut orientation but shows strong in-<br />

fluence on the corrugation <strong>of</strong> the step edges along the bunches. Straight ripples parallel to<br />

the DB double steps were found for a miscut along the [110] direction, whereas a triangular<br />

morphology was observed for a miscut along the [100] direction. In the latter case the DB<br />

double step segments are oriented ± 45 ◦ out <strong>of</strong> the miscut direction, which can be used to<br />

explain the experimental data seen with AFM [12].<br />

The experiments providing the dependence <strong>of</strong> the various parameters listed above were based<br />

on Si homoepitaxy only. Further studies employing heteroepitaxial <strong>SiGe</strong> layers proved that<br />

the amplitude <strong>of</strong> step-bunching is drastically reduced in the presence <strong>of</strong> Ge. Nevertheless,<br />

the influence <strong>of</strong> the Ge-content <strong>and</strong> the resulting strain is discussed controversially in the<br />

literature (see 4.1.2). [12, 54]<br />

4.1.2 <strong>Kinetic</strong> vs. <strong>Strain</strong>-<strong>Induced</strong> Step-Bunching<br />

Although step-bunching is experimentally extensively explored, it is still not fully understood.<br />

There is ongoing interest in step-bunching to learn more about surface step dynamics, which<br />

is approved by the appearance <strong>of</strong> new theoretical models <strong>and</strong> computer simulations.<br />

Various speculations <strong>and</strong> theories on the origin <strong>of</strong> step-bunching were published. In the<br />

early days strain-induced step-bunching was proposed [112, 113]. According to this one-<br />

dimensional theory on a vicinal surface under stress, elastic relaxation at steps produces a<br />

long-range attractive interaction between these steps. As a result, the surface should be<br />

unstable against step bunching, driven by the energetics <strong>of</strong> the system rather than by the<br />

kinetics <strong>of</strong> step flow [112]. <strong>Strain</strong>-induced step-bunching has also be claimed to be experi-<br />

mentally evident [114, 115, 116].<br />

More recent considerations propose kinetic step-bunching as an explanation for the evolv-<br />

ing ripple structure during growth. Several microscopic models evolved describing the step-<br />

bunching instability with the help <strong>of</strong> an inverse Schwoebel barrier [99], step edge diffusion [117]<br />

or the combination <strong>of</strong> attachment <strong>and</strong> detachment at the step edges <strong>and</strong> anisotropic diffu-<br />

sion [118, 119, 120]. Each <strong>of</strong> these processes can be characterized by specific critical exponents<br />

to characterize the correlation between layer-thickness t, ripple-height H <strong>and</strong> period Λ [105]<br />

(see also 4.1.1). Recapitulatory, there is a huge variety <strong>of</strong> different models, e.g. some models<br />

include elastic interactions <strong>of</strong> the terraces [121] whereas other descriptions do not [122].<br />

The results gathered in our group [106, 107, 108, 109, 110, 111, 118, 119, 120] strongly fa-<br />

vor kinetic step-bunching. In these experiments the exotic influences <strong>of</strong> electromigration,<br />

impurities, multispecies coupling by chemical reactions (see [123] <strong>and</strong> references therein)<br />

can be ruled out. Furthermore, the influence <strong>of</strong> strain has to be at least widely neglected


48 CHAPTER 4. SELF-ORGANIZED GROWTH – STEP-BUNCHING<br />

as the most pronounced ripple patterns were achieved by Si homoepitaxy. <strong>SiGe</strong> heteroepi-<br />

taxial layers showed a reduced peak-to-valley height regarding step-bunching <strong>and</strong> even in<br />

Si/<strong>SiGe</strong> superlattice structures the amplitude <strong>of</strong> the resulting ripple morphology appeared<br />

to be rather diminished. With increasing Ge-content the amplitude <strong>of</strong> the step-bunching<br />

instability decayed continuously although strain increases. This behavior can be explained<br />

with Ge segregation [124, 125, 126], which is significant at the employed temperatures, <strong>and</strong><br />

the alteration <strong>of</strong> the adatom dynamics due to the presence <strong>of</strong> even small amounts <strong>of</strong> Ge<br />

on the surface [127, 128]. Another strong indication against strain-driven step-bunching is<br />

that after-growth annealing does not enhance the ripple structure but leads to an attenu-<br />

ation <strong>of</strong> the step-bunching instability. The opposite behavior <strong>and</strong> an enhancement <strong>of</strong> the<br />

instability had to be expected if the effects were ruled by strain since its influence should be<br />

thermodynamically enhanced. [12, 54]<br />

4.2 Optimization <strong>of</strong> Step-Bunching for Ripple-Patterns with<br />

Small Periodicity<br />

It was already outlined in the previous part that the kinetic step-bunching instability is ex-<br />

tremely sensitive to the growth temperature [106, 108, 120]. For the experiments discussed<br />

in the following we used substrates with 4 ◦ miscut in [110] direction (ø 3”, 375 µm thick, CZ,<br />

(001)-orientation, 4.0 ◦ [110] miscut; n-doped, As, 1 – 10 Ωcm) that were cut to a practical<br />

size <strong>of</strong> 17.5 mm × 17.5 mm. All growth procedures started with a high temperature Si-buffer<br />

which should ensure a flat surface <strong>and</strong> reproducible starting point for the study.<br />

With increasing substrate miscut from 0.66 ◦ to 4 ◦ the ripple period could be tuned to<br />

smaller periods towards the nanometer scale. At Si growth rates <strong>of</strong> 0.2 ˚A/s a ripple pattern<br />

with few defects develops within a small temperature window around 425 ◦ C. Although the<br />

layers are grown under UHV-conditions in our MBE-system (base pressure < 10 −10 mbar),<br />

at slightly lower temperatures many defects are incorporated <strong>and</strong> the ripples decompose into<br />

isl<strong>and</strong>s, which are aligned in chains (Fig. 4.3, 400 ◦ C). For marginally higher temperatures the<br />

ripple structure fades away (Fig. 4.3, 450 ◦ C). This behavior is visualized in Fig. 4.4 where the<br />

ripple height <strong>of</strong> the step-bunching structure is plotted as a function <strong>of</strong> substrate temperature<br />

for a 1000 ˚A Si-buffer. Data for growth temperatures below 400 ◦ C were not evaluated since<br />

already for 400 ◦ C the ripple morphology changes gradually to a hillock pattern. Generally,<br />

the transition from hillocks, appearing at low temperatures, to elongated ripples through lat-<br />

eral bunch expansion leads to bifurcations at optimized temperatures . These are caused by<br />

ripples which originate from different parts <strong>of</strong> the sample, <strong>and</strong> thus can be accidentally out-<strong>of</strong>-


4.2. OPTIMIZATION OF STEP-BUNCHING FOR RIPPLE-PATTERNS 49<br />

Figure 4.3: AFM images showing the growth temperature dependence <strong>of</strong> stepbunching.<br />

The series <strong>of</strong> 500 ˚A thick Si-buffers grown on 4 ◦ [110] miscut samples<br />

confirms that only within a small temperature range (here at around 425 ◦ C) a<br />

pronounced ripple structure evolves [89, 106]. (miscut vector points upwards!)<br />

� Source: Stepbunching AFM Temp-series.jpg<br />

phase. This opens up a 2D-pathway for ”transversal” coarsening (proposed as ”ripple-zipper”<br />

mechanism by C. Schelling et al. [12]) which increases the period <strong>of</strong> step-bunching for increas-<br />

Figure 4.4: Evaluation <strong>of</strong> the step-bunching ripple height as a function <strong>of</strong> substrate<br />

temperature for a 1000 ˚A Si-buffer. The maximum <strong>of</strong> the instability for 4 ◦<br />

miscut <strong>and</strong> a Si-rate <strong>of</strong> 0.2 ˚A/s occurs around 425 ◦ C.<br />

� Source: Stepbunching temp eval.jpg


50 CHAPTER 4. SELF-ORGANIZED GROWTH – STEP-BUNCHING<br />

Figure 4.5: AFM images showing the layer thickness dependence <strong>of</strong> step-bunching<br />

at 425 ◦ C. The series <strong>of</strong> 250, 500, 1000 <strong>and</strong> 3000 ˚A thick Si-buffers grown on 4 ◦ [110]<br />

miscut samples demonstrates that the ripple structure gets more pronounced with<br />

increasing layer thickness [89].<br />

� Source: Stepbunching AFM Thick-series.jpg<br />

ing layer thickness. For perfect ripples spanning over several micrometers only ”longitudinal”<br />

coarsening with the dissolution <strong>of</strong> intermediate bunches <strong>and</strong> excessive material transport can<br />

Figure 4.6: Evaluation <strong>of</strong> the step-bunching ripple period (a) <strong>and</strong> height (b) on<br />

the thickness <strong>of</strong> the Si-buffer for 4 ◦ [110] miscut. Both values gradually increase<br />

with layer-thickness but show seemingly a reduced gain for thicker layers.<br />

� Source: Stepbunching period-height eval.jpg


4.2. OPTIMIZATION OF STEP-BUNCHING FOR RIPPLE-PATTERNS 51<br />

Figure 4.7: AFM image (a) <strong>and</strong> Fourier transform (b) <strong>of</strong> optimized 1000 ˚A thick<br />

Si-buffer grown with a Si-rate <strong>of</strong> 0.2 ˚A/s at 425 ◦ C on a 4 ◦ [110] miscut substrate<br />

resulting in a regular ripple template.<br />

� Source: Stepbunching AFM FFT.jpg<br />

lead to continuously growing ripples [112] (see also [12] for a discussion). Therefore also the<br />

influence <strong>of</strong> the Si-buffer layer thickness was explored systematically (Fig. 4.5). In the early<br />

stage <strong>of</strong> step bunching (Fig. 4.5, 250 ˚A) there are many uncorrelated localized step bunches.<br />

With increasing layer thickness (500 ˚A) the individual bunch segments merge <strong>and</strong> form well-<br />

pronounced elongated ripples (500 ˚A), which finally span several micrometers (1000 ˚A). The<br />

low growth temperature <strong>of</strong> 425 ◦ C leads to an increasing number <strong>of</strong> accumulated defects with<br />

increasing layer thickness. These show up occasionally as holes <strong>and</strong> constricted bunches<br />

(Fig. 4.5, 3000 ˚A). For this reason the layer thickness was not further increased within this<br />

series, which impedes an extraction <strong>of</strong> reasonable coefficients for the power-law dependence<br />

<strong>and</strong> scaling <strong>of</strong> step-bunching (see 4.1.1). Nevertheless, the evaluation <strong>of</strong> the ripple period <strong>and</strong><br />

height <strong>of</strong> the step-bunches as a function <strong>of</strong> the Si-buffer thickness for 4 ◦ miscut along [110] is<br />

depicted in Fig. 4.6. Both values gradually increase with layer-thickness but show seemingly<br />

a reduced gain for thicker layers.<br />

Fig. 4.7 shows an extended AFM image (5 µm scan-size) <strong>and</strong> the corresponding Fourier trans-<br />

form <strong>of</strong> the optimized 1000 ˚A thick Si-buffer grown with a Si-rate <strong>of</strong> 0.2 ˚A/s at 425 ◦ C on<br />

typical 4 ◦ [110] miscut substrates resulting in a regular ripple template <strong>and</strong> providing the<br />

base for further experiments. The measured parameters are Λ ∼ 105 nm for the period <strong>and</strong><br />

H ∼ 4 nm for the ripple height. Fig. 4.8 <strong>and</strong> Fig. 4.9 show tilt-corrected AFM data for the<br />

same optimized Si-buffer with a scan-size <strong>of</strong> 500 nm. From the line scan (Fig. 4.8) <strong>and</strong> the<br />

surface-angle-plot (SAP, Fig. 4.9) the typical morphology <strong>of</strong> step-bunching is revealed. In<br />

the corner regions between the lower ends <strong>of</strong> the slope <strong>and</strong> the adjacent terraces seemingly<br />

also retrograding steps become visible. It is not really clear whether this is an artifact <strong>of</strong><br />

the measurement or due to applying tilt-correction. Similar line-scans were also reported


52 CHAPTER 4. SELF-ORGANIZED GROWTH – STEP-BUNCHING<br />

Figure 4.8: AFM image <strong>and</strong> corresponding line-scan in miscut direction for a<br />

nominally 4 ◦ [110] miscut sample with the optimized 1000 ˚A thick Si-buffer. By<br />

correcting the AFM data for the miscut (3.86 ◦ ) the line-scan reveals extended terraces<br />

in (001)-orientation.<br />

� Source: Stepbunching linescan.jpg<br />

Figure 4.9: Analysis <strong>of</strong> local surface inclination from the tilt-corrected AFM<br />

image with a scan-size <strong>of</strong> 500 nm (see Fig. 4.8). The plot illustrates the surface<br />

l<strong>and</strong>scape with the flat terraces <strong>and</strong> the steep ripple flanks which is typical for stepbunching.<br />

� Source: Stepbunching surfangles.jpg


4.2. OPTIMIZATION OF STEP-BUNCHING FOR RIPPLE-PATTERNS 53<br />

Figure 4.10: AFM images giving a comparison between a 4 ◦ [110] miscut substrate<br />

(a) <strong>and</strong> a flat reference sample (b) grown simultaneously. The epilayer consists<br />

<strong>of</strong> a 1000 ˚A thick Si-buffer grown at 425 ◦ C <strong>and</strong> a rate <strong>of</strong> 0.2 ˚A/s.<br />

� Source: Stepbunching miscut-dummy.jpg<br />

by the preworkers [12]. The overall l<strong>and</strong>scape consists <strong>of</strong> flat extended terraces with (001)-<br />

orientation <strong>and</strong> steep ripple flanks with typical slopes slightly above 10 ◦ . In the central part<br />

<strong>of</strong> the surface inclination visualization a region <strong>of</strong> out-<strong>of</strong>-phase bunches can be seen, whereas<br />

in the outer right part a constriction gets clearly visible.<br />

In all sample series growth was performed on a miscut sample <strong>and</strong> a flat reference sample<br />

simultaneously. The morphologies <strong>of</strong> the grown epilayers were compared by AFM (Fig. 4.10).<br />

For the flat reference, Fig. 4.10b reveals that in the low growth temperature regime around<br />

425 ◦ C step-flow growth is suppressed <strong>and</strong> ”isl<strong>and</strong>ing” occurs. The small sharp dot-like<br />

features in the 2 µm AFM data <strong>of</strong> the reference sample can be attributed to the onset <strong>of</strong><br />

oxidation. The samples were taken out from the UHV-system right before the AFM char-<br />

acterization. Still, such artifacts could not be prevented since the time consuming AFM<br />

Figure 4.11: AFM images point up the importance <strong>of</strong> sample cleaning. Scan<br />

sizes <strong>of</strong> 2 µm (a) <strong>and</strong> 20 µm (b) show many holes <strong>and</strong> line defects in the optimized<br />

1000 ˚A thick Si-buffer grown at 425 ◦ C.<br />

� Source: Stepbunching defects.jpg


54 CHAPTER 4. SELF-ORGANIZED GROWTH – STEP-BUNCHING<br />

characterizations were performed on air.<br />

It cannot be pointed out with enough emphasis that a proper sample preparation is <strong>of</strong> great<br />

importance. A failed cleaning procedure due to problems with the DI-water system can<br />

have significant consequences as evidenced in Fig. 4.11. AFM images with different scan-<br />

sizes <strong>of</strong> 2 µm (Fig. 4.11a) <strong>and</strong> 20 µm (Fig. 4.11b) show many holes <strong>and</strong> line defects in the<br />

otherwise – especially regarding growth parameters – optimized 1000 ˚A thick Si-buffer grown<br />

at 425 ◦ C. This again justifies our elaborate cleaning procedures especially applied to the<br />

17.5 mm × 17.5 mm sample pieces. [89, 90]


Chapter 5<br />

<strong>Self</strong>-Organized Growth 2 –<br />

Combination <strong>of</strong> <strong>Strain</strong>-Effects <strong>and</strong><br />

<strong>Kinetic</strong> Phenomena<br />

This chapter deals with several aspects <strong>of</strong> self-organized growth in the Si/<strong>SiGe</strong> system as al-<br />

ready outlined in the preamble <strong>of</strong> the preceding chapter (see Ch. 4). The special morphology<br />

<strong>and</strong> periodic surface modulation <strong>of</strong> step-bunching provide a versatile basis for the investiga-<br />

tion <strong>of</strong> the interplay between kinetics, surface energy <strong>and</strong> strain. Varying the parameters for<br />

the <strong>SiGe</strong>-epilayers grown on top <strong>of</strong> these ripple templates delivers insight <strong>and</strong> better under-<br />

st<strong>and</strong>ing on the nucleation <strong>and</strong> evolution <strong>of</strong> <strong>SiGe</strong>-isl<strong>and</strong>s. In an introductory discussion the<br />

self-organized <strong>SiGe</strong>-isl<strong>and</strong>s found as byproduct here are linked to state-<strong>of</strong>-the-art perfectly<br />

organized <strong>SiGe</strong>-dots on pre-patterned substrates.<br />

5.1 Introduction to State-<strong>of</strong>-the-Art Ge-Dots<br />

In recent years self-organized <strong>SiGe</strong>/Ge dots attracted increased interest <strong>and</strong> are regarded<br />

promising c<strong>and</strong>idates for nano-electronic <strong>and</strong> optoelectronic applications due to quantum size<br />

effects [129, 130, 131, 132]. This is supported by the compatibility <strong>of</strong> these feasible devices<br />

with the sophisticated <strong>and</strong> well-established Si microelectronic technology [133]. Conventional<br />

deposition <strong>of</strong> self-assembled Ge-dots leads to a heterogeneous growth <strong>and</strong> thus fluctuations<br />

in size <strong>and</strong> r<strong>and</strong>om positioning <strong>of</strong> the isl<strong>and</strong>s. Although the exact positioning is not <strong>of</strong> direct<br />

interest for optical properties, the size homogeneity is usually improved along with lateral<br />

55


56 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

ordering. Otherwise, for electronic applications the direct access to individual dots <strong>and</strong> a<br />

controlled spatial arrangement is inevitable for functional devices. [93, 134]<br />

Under optimized growth conditions also for conventional unpatterned Si(001) substrates<br />

uniform self-assembled Ge-dots were achieved [135]. Surfactant-mediated growth turned out<br />

to yield homogeneous high-density Ge-dots (see Sec. 5.3). Furthermore, processes based on<br />

natural substrate pre-patterning via growth instabilities on vicinal substrates [85, 104] also<br />

resulted in the formation <strong>of</strong> purely self-organized Ge-dot growth <strong>and</strong> therefore without special<br />

ex situ substrate preparation steps. Nevertheless the <strong>SiGe</strong>-system does not deliver perfect 3D-<br />

ordering <strong>of</strong> quantum dots based on strain fields alone as found for other material systems [136].<br />

Therefore different approaches were adopted to achieve organized Ge-dot nucleation <strong>and</strong><br />

lateral positioning. Vertical ordering <strong>of</strong> Ge-dots in growth direction gets accessible via residual<br />

strain from the dot-layers beneath, which favors nucleation above the buried isl<strong>and</strong>s, <strong>and</strong> thus<br />

results in stacking for Si/<strong>SiGe</strong> multi-layer structures. In these superlattices strain-filtering<br />

can improve the size homogeneity for the upper Ge-dot epilayers [137]. Various attempts to<br />

achieve 2D ordering include selective epitaxial growth within the windows <strong>of</strong> SiO2-masks [138,<br />

139, 140], buried stress centers from oxygen implantation [141], implementation <strong>of</strong> intended<br />

defects in general [142] or misfit dislocation networks [143]. Also, conventional lithographic<br />

means with subsequent etching are employed to generate templates with modulated surfaces,<br />

such as mesa structures [144, 145, 146, 147] or pits [148], <strong>of</strong>fering preferential nucleation<br />

sites for Ge-isl<strong>and</strong>s [149]. The periodicity <strong>of</strong> 2D pit-patterned substrates is ever shrunk by<br />

optimization <strong>and</strong> introduction <strong>of</strong> new lithographic means such as e-beam lithography [91],<br />

nanoimprinting [150], focused ion beam (FIB) etching [134] <strong>and</strong> x-ray interference lithography<br />

(XIL) [151]. These high-density pit-patterns with a typical pitch well below 100 nm provide<br />

well-defined periodic nucleation sites. For every pit the same amount <strong>of</strong> Ge-atoms contribute<br />

to 3D growth. Thus equal material catchment areas for each single Ge-isl<strong>and</strong> enable state-<br />

<strong>of</strong>-the-art Ge-dot growth with high homogeneity <strong>and</strong> a narrow size distribution. [93, 134]<br />

Novel x-ray techniques are developed to characterize the growth mode, strain state <strong>and</strong><br />

shape <strong>of</strong> Ge isl<strong>and</strong>s during their growth on Si(001) in situ [152]. Up to now real-time Ge-dot<br />

growth analyses were mainly reserved to STM experiments [153], which were especially suited<br />

to explore the strain-driven transition from 2D- to 3D-growth <strong>of</strong> Ge on Si [154] with high<br />

resolution. 2D-pit-patterned substrates are not only promising c<strong>and</strong>idates for the realization<br />

<strong>of</strong> seeded <strong>and</strong> organized growth for future applications, but also open the opportunity for basic<br />

research to get a deeper insight <strong>and</strong> a detailed underst<strong>and</strong>ing <strong>of</strong> the underlying mechanisms<br />

<strong>of</strong> Ge-dot formation.


5.2. KINETIC STEP-BUNCHING AND STRAIN-DRIVEN ISLAND GROWTH 57<br />

Figure 5.1: AFM images <strong>of</strong> the optimized Si-ripple template (Sec. 4.2) as grown<br />

(a) <strong>and</strong> after annealing at 550 ◦ C for 15 min (b). The ripple amplitude was decreased<br />

during the post-growth heat treatment from ∼ 5 nm to less than ∼ 2 nm.<br />

� Source: <strong>Kinetic</strong>s n <strong>Strain</strong> annealing.jpg<br />

5.2 Interplay between <strong>Kinetic</strong> Step-Bunching <strong>and</strong> <strong>Strain</strong>-<br />

Driven Isl<strong>and</strong> Growth<br />

5.2.1 Influence <strong>of</strong> Ripple-Template Annealing<br />

In the following especially the growth-temperature dependence <strong>of</strong> <strong>SiGe</strong>-epilayers is evaluated.<br />

For most <strong>SiGe</strong>-deposition procedures the temperature has to be ramped up to promote strain<br />

effects <strong>and</strong> pronounced facet formation by approaching thermodynamic equilibrium. Hence<br />

the growth temperature for the <strong>SiGe</strong>-epilayer has to be adjusted <strong>and</strong> stabilized within a<br />

short growth interrupt directly after the preparation <strong>of</strong> the periodic Si-ripple template. As<br />

the temperature ramp is known to affect the integrity <strong>of</strong> the ripple template the dependence<br />

<strong>of</strong> annealing has to be checked to ensure a remaining 2D-pattern to be present at least in the<br />

initial stage <strong>of</strong> <strong>SiGe</strong>-deposition. The AFM data in Fig. 5.1 show the decrease in amplitude <strong>of</strong><br />

the optimized Si-ripple template (Sec. 4.2) after annealing at 550 ◦ C for 15 min. The ripple<br />

height was diminished during the post-growth heat treatment from ∼ 5 nm to less than ∼ 2 nm.<br />

This result confirms again the kinetic origin <strong>of</strong> step-bunching but it unambiguously makes<br />

clear that the ramp-up to the <strong>SiGe</strong>-epilayer deposition temperature has to be performed<br />

as fast as possible to avoid loosing the ripple-structure completely. Especially for high-<br />

temperature <strong>SiGe</strong>-epilayers the thermal volatility <strong>of</strong> the template might be a problem.<br />

5.2.2 <strong>SiGe</strong>-Overgrowth <strong>of</strong> Step-Bunching Template – <strong>Strain</strong>-Effects<br />

All further experiments were conducted with the optimized 1000 ˚A thick Si-buffer (at 425 ◦ C,<br />

0.2 ˚A/s Si), which show typical dimensions <strong>of</strong> 100 nm for the ripple period <strong>and</strong> 4 nm for


58 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

the ripple height (miscut 4 ◦ [110]). By increasing the substrate miscut to 4 ◦ the period <strong>of</strong><br />

the evolving step-bunching structure was reduced to such an extent that the spacing <strong>of</strong> the<br />

pronounced ripple morphology complies with the average distance <strong>of</strong> <strong>SiGe</strong>-isl<strong>and</strong>s deposited in<br />

the Stranski-Krastanov growth mode. The step-bunching templates supply a one-dimensional<br />

pattern with preferable nucleation sites for <strong>SiGe</strong>-isl<strong>and</strong>s along the ripple flanks, <strong>and</strong> thus<br />

allows us to combine kinetic <strong>and</strong> strain-driven self-organization phenomena in the Si/<strong>SiGe</strong><br />

heterosystem.<br />

The deposition parameters for the <strong>SiGe</strong>-epilayer were varied to find parameters where regular<br />

strain-driven features appear. Fig. 5.2 shows the comparison <strong>of</strong> two different strained <strong>SiGe</strong>-<br />

epilayers grown on top <strong>of</strong> the rippled Si-buffer at 425 ◦ C. The excessive strain in the 50 ˚A<br />

Si0.55Ge0.45 top-layer (1506LSG) leads to more pronounced <strong>and</strong> regular features compared to<br />

the 150 ˚A Si0.75Ge0.25 epilayer (1497LSG). The morphology <strong>of</strong> the <strong>SiGe</strong>-layer shows the onset<br />

<strong>of</strong> 3D-growth with additional ridges in miscut direction <strong>and</strong> perpendicular to the elongated<br />

bunches. The layer sequence <strong>and</strong> deposition parameters for the rippled Si-buffer <strong>and</strong> <strong>SiGe</strong>-<br />

epilayers are listed in Tab. 5.1.<br />

For the following investigations the thinner highly strained <strong>SiGe</strong>-epilayer (50 ˚A Si0.55Ge0.45)<br />

was used to check the influence <strong>of</strong> the deposition temperature on the evolution <strong>of</strong> the ridge<br />

structure. The strain is limited to < 2% to avoid plastic relaxation but high enough to get<br />

access to the underlying mechanism.<br />

Fig. 5.3 <strong>and</strong> Fig. 5.4 show data based on Atomic Force Microscopy (AFM) for a pure, 1000 ˚A<br />

Sample 1497LSG 1506LSG 1539LSG<br />

High T Buffer – – 240 ˚A Si<br />

Si @ 0.2 ˚A/s<br />

750 ◦ C → 425 ◦ C<br />

Rippled Buffer 1000 ˚A Si 1000 ˚A Si 1000 ˚A Si<br />

Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s<br />

425 ◦ C 425 ◦ C 425 ◦ C<br />

Growth interrupt – – 425 ◦ C → 625 ◦ C<br />

(5 min)<br />

<strong>SiGe</strong> epilayer 150 ˚A Si0.75Ge0.25 50 ˚A Si0.55Ge0.45 50 ˚A Si0.55Ge0.45<br />

Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s<br />

Ge @ 0.0667 ˚A/s Ge @ 0.1636 ˚A/s Ge @ 0.1636 ˚A/s<br />

425 ◦ C 425 ◦ C 625 ◦ C<br />

Table 5.1: Growth sequence for ripple-template preparation <strong>and</strong> <strong>SiGe</strong>-deposition.


5.2. KINETIC STEP-BUNCHING AND STRAIN-DRIVEN ISLAND GROWTH 59<br />

Figure 5.2: AFM data for <strong>SiGe</strong>-epilayers grown on top <strong>of</strong> the rippled Si-buffer<br />

at 425 ◦ C. The increased strain in the 50 ˚A Si0.55Ge0.45 top-layer (b, d, 1506LSG)<br />

leads to more pronounced <strong>and</strong> regular features compared to the 150 ˚A Si0.75Ge0.25<br />

epilayer (a, c, 1497LSG).<br />

� Source: <strong>Kinetic</strong>s n <strong>Strain</strong> Ge-cont var.jpg<br />

thick Si-buffer, <strong>and</strong> the same Si-buffer covered with 50 ˚A Si0.55Ge0.45 at temperatures ranging<br />

from 350 ◦ C to 625 ◦ C. The epilayer morphology for various temperatures is illustrated with<br />

3D-AFM data (Fig. 5.3c <strong>and</strong> Fig. 5.4c).<br />

Only at very low temperatures around 350 ◦ C the Si0.55Ge0.45 film replicates the underlying<br />

ripples <strong>of</strong> the Si-buffer in a conformal manner. At 425 ◦ C the stress in the top-layer leads to the<br />

formation <strong>of</strong> the aforementioned ridges at the ripple flanks decorating the main structure. A<br />

slight increase <strong>of</strong> the substrate temperature leads to the development <strong>of</strong> {105}-faceted isl<strong>and</strong>s,<br />

which are known from the hut-clusters <strong>of</strong> <strong>SiGe</strong>- <strong>and</strong> Ge-films [155]. Due to the high miscut <strong>of</strong><br />

our substrates the isl<strong>and</strong>s are bound by two {105}-facets <strong>and</strong> the (001)-facet on top, while the<br />

underlying ripple pattern is widely conserved. The Fast Fourier Transform (FFT) evaluations<br />

in Fig. 5.3b <strong>and</strong> Fig. 5.4b reveal a period <strong>of</strong> approximately 100 nm for the step bunches on the<br />

Si-buffer under the chosen growth conditions. At medium temperatures around 550 ◦ C the<br />

Si0.55Ge0.45 isl<strong>and</strong>s decorate the kinetic step bunches with a single dot row per step bunching<br />

period, but have a somewhat smaller spacing <strong>of</strong> ∼ 70 nm along the bunches. Nevertheless there<br />

is no clear lateral ordering discernible for 550 ◦ C, as can be seen from the broad halo that


60 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.3: AFM data for a pure Si-buffer, <strong>and</strong> Si-buffers covered with 50 ˚A<br />

Si0.55Ge0.45 at various temperatures. Conventional 2D-AFM images (a), are complemented<br />

with FFTs (b), 3D-AFM (c) <strong>and</strong> SOM representations (d), to illustrate<br />

the transition from pure step bunching to {105}-faceted isl<strong>and</strong>s (see also Fig. 5.4).<br />

� Source: <strong>Kinetic</strong>s n strain <strong>SiGe</strong>45 varT01.jpg


5.2. KINETIC STEP-BUNCHING AND STRAIN-DRIVEN ISLAND GROWTH 61<br />

Figure 5.4: AFM data <strong>and</strong> evaluation in continuation <strong>of</strong> Fig. 5.3 [89, 90].<br />

� Source: <strong>Kinetic</strong>s n strain <strong>SiGe</strong>45 varT02.jpg<br />

appears in the FFT in addition to the well-defined signal <strong>of</strong> the periodic ripples. The ripple<br />

flanks are preferable nucleation sites which can be used to achieve at least one-dimensional<br />

ordering <strong>of</strong> the <strong>SiGe</strong>-isl<strong>and</strong>s [105]. The adatom diffusion length, which depends exponentially


62 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

on the substrate temperature, rules the spacing <strong>of</strong> the <strong>SiGe</strong>-isl<strong>and</strong>s along the ripples, whereas<br />

in an appropriate temperature window the perpendicular dot distance in miscut direction is<br />

given by the ripple period only. At 625 ◦ C (see Tab. 5.1, 1539LSG) the adatom diffusion length<br />

is sufficiently high, so that the <strong>SiGe</strong>-isl<strong>and</strong> spacing matches the period <strong>of</strong> the ripples well <strong>and</strong><br />

an homogeneous ordering, both along <strong>and</strong> perpendicular to the ripple pattern is found. The<br />

respective FFT shows pronounced ordering already for a single <strong>SiGe</strong>-layer. Although hard to<br />

distinguish from a distorted hexagonal ordering, the four-fold symmetry <strong>of</strong> the hut clusters<br />

on Si(001) <strong>and</strong> mutual isl<strong>and</strong> repulsion along the [100]-directions make it more likely that<br />

the isl<strong>and</strong>s order anti-correlated in a face-centered rectangular fashion as already reported<br />

by Zhu et al. [85].<br />

The individual facets <strong>of</strong> the isl<strong>and</strong>s are determined from a surface-orientation-histogram,<br />

which is derived from tilt-corrected AFM-data (Fig. 5.3d <strong>and</strong> Fig. 5.4d). For each surface<br />

point the normal orientation [hkl] is calculated with the nearest-neighbor points to deter-<br />

mine the local plane <strong>and</strong> then plotted in polar coordinates (ρ, φ). Whereas ρ indicates the<br />

angle between the respective local surface normal [hkl] <strong>and</strong> the (001)-surface (growth direc-<br />

tion), φ denotes the in-plane azimuthal angle <strong>of</strong> [hkl] with respect to [100]. The intensity<br />

at each point (ρ, φ) in the surface orientation histogram is a measure <strong>of</strong> the abundance <strong>of</strong> a<br />

respective surface orientation (see also Sec. 3.3.4). The marker circles correspond to surface<br />

angles ρ <strong>of</strong> 11.3 ◦ <strong>and</strong> 25.2 ◦ , which are introduced as guidance for the characteristic {105}-<br />

<strong>and</strong> {113}-facets in the <strong>SiGe</strong>-system [155, 156]. Both the pure Si-buffer <strong>and</strong> the buffer over-<br />

grown with 50 ˚A Si0.55Ge0.45 at 350 ◦ C show just a streak spanning from (001) to (1 1 10)<br />

along the [110]-direction with an average slope <strong>of</strong> 10±2 ◦ . Obviously, the surface consists <strong>of</strong><br />

(001)-oriented stripes <strong>and</strong> more <strong>and</strong> less inclined regions in the flanks <strong>of</strong> the step-bunches<br />

which indicates a rounding <strong>of</strong> the ripple edges. For growth temperatures between 425 ◦ C<br />

<strong>and</strong> 550 ◦ C the surface orientation signal extends from the center – corresponding to (001) –<br />

towards the [105]- <strong>and</strong> [015]-directions. The asymmetry in the appearance <strong>of</strong> the {105}-facets<br />

is based on the high miscut (Fig. 5.5) causing the shape elongation due to the vicinality <strong>of</strong> the<br />

surface [157] <strong>and</strong> resulting in a rhombic base <strong>of</strong> the pyramids [158] (see also Sec. 5.3). The<br />

(105)- <strong>and</strong> (015)-facets can easily grow out <strong>of</strong> the ripple-flanks whereas the formation <strong>of</strong> the<br />

retrograding (105)- <strong>and</strong> (015)-facets occurs only for an extension <strong>of</strong> the ”downhill”-facets <strong>and</strong><br />

elevation beyond the upper terrace, which is kinetically suppressed at lower temperatures.<br />

At 625 ◦ C the height <strong>of</strong> the dots drastically increases forming rather symmetric dots with<br />

{105}- <strong>and</strong> higher-index facets. Although the underlying ripple pattern <strong>of</strong> the Si-buffer is<br />

now hardly visible, since it has already begun to dissolve thermally (see also 5.2.1), still, the<br />

nucleation sites <strong>of</strong> the <strong>SiGe</strong>-dots are obviously influenced by the ripple pattern: reference


5.2. KINETIC STEP-BUNCHING AND STRAIN-DRIVEN ISLAND GROWTH 63<br />

Figure 5.5: 3D-AFM data <strong>and</strong> schematic drawings for a rippled Si-buffer covered<br />

with 50 ˚A Si0.55Ge0.45. The distorted AFM-data representation (b) <strong>of</strong> the 3Ddata<br />

(a) helps visualizing the faceted zigzag pattern <strong>of</strong> the step-bunching areas.<br />

The dominant facets for the <strong>SiGe</strong>-layers deposited at 425 ◦ C <strong>and</strong> above 550 ◦ C are<br />

depicted schematically in (c) <strong>and</strong> (d), respectively [89, 90].<br />

� Source: <strong>Kinetic</strong>s n strain model.jpg<br />

samples on substrates with no miscut show a clear 4-fold symmetry in the FFT, which can<br />

be easily explained with the typical hut-cluster networks (see Sec. 5.3). The <strong>SiGe</strong> top-layer is<br />

decomposed into uniform pyramids <strong>and</strong> most <strong>of</strong> the signal is evenly distributed over all four<br />

{105}-facets. Also, at these high growth temperatures signs <strong>of</strong> high-index facets, which are<br />

known from Ge-domes, are found (see also Sec. 5.4).<br />

Fig. 5.5 illustrates schematically how, <strong>and</strong> where, isl<strong>and</strong> nucleation commences: The <strong>SiGe</strong>-<br />

film does not completely disintegrate into individual isl<strong>and</strong>s. Instead, upon <strong>SiGe</strong>-deposition<br />

the flanks <strong>of</strong> the step bunches are converted into a zigzag train <strong>of</strong> adjacent (105)- <strong>and</strong> (015)-<br />

facets, which is in fact a strain-driven step-me<strong>and</strong>ering instability (Fig. 5.5a-b) that was<br />

already reported by Teichert et al. [115, 116]. The originally smooth flanks with typically<br />

10±2 ◦ inclination with respect to (001) are energetically favorable nucleation sites for the<br />

strained <strong>SiGe</strong>-epilayer leading to the pronounced {105}-faceted ridge structure. These su-<br />

perimposed features are oriented perpendicular to the step bunches <strong>and</strong> mark the transition


64 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

from conformal Si/<strong>SiGe</strong> epilayer growth [88] to strain-driven 3D-isl<strong>and</strong> growth, where the<br />

epilayer completely decomposes into separated isl<strong>and</strong>s (T ∼ 550 ◦ C). For 425 ◦ C the ridge<br />

structure consists only <strong>of</strong> the (105)- <strong>and</strong> (015)-facets that emerge easily out <strong>of</strong> the ripple-<br />

flanks <strong>and</strong> are bound by the upper (001)-terrace to the top (Fig. 5.5c). Higher temperatures<br />

favoring strain are needed to overcome the kinetic limitation <strong>and</strong> finally enable the formation<br />

<strong>of</strong> the retrograding (105)- <strong>and</strong> (015)-facets (Fig. 5.5d). This is realized with the extension <strong>of</strong><br />

the ”downhill”-facets <strong>and</strong> elevation beyond the upper terrace, which is confining the ridges<br />

for lower temperatures. Obviously, the Ge-rich epilayer material avoids nucleation on the<br />

(001)-terraces in favor <strong>of</strong> 3D-growth on the ripple flanks [86]. The distorted AFM-data rep-<br />

resentation (Fig. 5.5b) <strong>of</strong> the 3D-data (Fig. 5.5a) helps visualizing the faceted zigzag pattern<br />

<strong>of</strong> the step-bunching areas.<br />

The measured slope <strong>of</strong> the ripple flanks corresponds well with the 8.05 ◦ inclination with<br />

respect to the (001)-surface <strong>of</strong> the [551] intersection line between two adjacent {105}-facets.<br />

Therefore a special situation is expected for substrate areas with a local miscut <strong>of</strong> ∼ 8–<br />

10 ◦ . There the {105}-faceted ridges can extend over larger distances in [110]-direction.<br />

Such structures were demonstrated with <strong>SiGe</strong>-wires in extended holes with slowly varying<br />

Figure 5.6: Visualization <strong>of</strong> strain relaxation. The compressively strained<br />

Si0.55Ge0.45 epilayer can easily relief stress perpendicular to the {105}-faceted ridges<br />

(broad black arrows). Due to the tilted base plane <strong>of</strong> these ridges (i.e {1 1 10}) there<br />

are densely stepped edges <strong>of</strong> the wire-like structure along the [551]-direction which<br />

enables strain relaxation along the wire direction (broad red arrows).<br />

� Source: <strong>Kinetic</strong>s n strain relaxation.jpg


5.3. ORDERING AND SIZE OF SIGE-ISLANDS 65<br />

slope [159, 160], <strong>and</strong> with Ge-wires where STM-measurements were used to resolve the recon-<br />

struction <strong>of</strong> the {105}-facets [105]. For these extended ridges the surface energy is minimized<br />

with {105}-facets, which is immediately plausible. For strain energy relaxation the wire<br />

structures might not seem that ideal right from the beginning. Clearly the one-dimensional<br />

structure can easily relax strain in perpendicular direction. But also the densely stepped<br />

{105}-facets are not completely inadequate for relieving strain energy along the wire direc-<br />

tion. This is visualized in Fig. 5.6. Due to the tilted base plane <strong>of</strong> the {105}-faceted ridges (i.e<br />

{1 1 10}) there are densely stepped edges in the wire-like structure along the [551]-direction.<br />

This enables strain relaxation also along the wire direction (broad red arrows in Fig. 5.6).<br />

The experiments clearly demonstrate that a reduction <strong>of</strong> the step-bunch spacing to a<br />

value approaching the average spacing <strong>of</strong> the Si0.55Ge0.45 isl<strong>and</strong>s under the chosen growth<br />

conditions couples the two otherwise independent mechanisms <strong>of</strong> kinetic (homoepitaxial) step<br />

bunching [118, 120] <strong>and</strong> <strong>of</strong> strain-induced 3D isl<strong>and</strong> growth [88]. This way long-range ordering<br />

<strong>of</strong> self-organized <strong>SiGe</strong>-dots is achieved that is entirely based on self-organization phenomena.<br />

By optimizing the growth parameters, <strong>and</strong> by introducing strain filtering [137, 161], further<br />

improvements in size-uniformity <strong>and</strong> ordering are possible [85, 105, 110] which will be required<br />

for any potential applications (also see Sec. 5.1). [89, 90]<br />

5.3 Ordering <strong>and</strong> Size <strong>of</strong> <strong>SiGe</strong>-Isl<strong>and</strong>s<br />

In this section the ordering <strong>of</strong> the <strong>SiGe</strong>-isl<strong>and</strong>s found in Sec. 5.2.2 is picked up again. Addi-<br />

tional experiments with different Ge-content, layer-thickness <strong>and</strong> growth temperature docu-<br />

ment the influence <strong>of</strong> strain on ordering <strong>and</strong> size <strong>of</strong> the <strong>SiGe</strong>/Ge isl<strong>and</strong>s. The main intention is<br />

again to clarify the importance <strong>of</strong> the ripple template for the nucleation <strong>of</strong> the <strong>SiGe</strong>-epilayers<br />

on the 4 ◦ miscut substrates in comparison with flat st<strong>and</strong>ard Si(001).<br />

Fig. 5.7 shows the comparison <strong>of</strong> ordering for a 50 ˚A Si0.55Ge0.45 epilayer deposited at<br />

625 ◦ C (1539LSG) on a 4 ◦ miscut sample (Fig. 5.7a) <strong>and</strong> a flat Si(001) reference substrate<br />

(Fig. 5.7b). As already stated in Sec. 5.2.2 for the step-bunching template a seemingly hexag-<br />

onal ordering is found [85]. The moderate strain in the Si0.55Ge0.45 isl<strong>and</strong>s adjusts the isl<strong>and</strong><br />

size towards the ripple period resulting in a single-isl<strong>and</strong>-row arrangement per bunch. This<br />

leads to a mean isl<strong>and</strong> distance <strong>of</strong> 95 nm <strong>and</strong> a density <strong>of</strong> 8×10 9 cm −2 . For the flat st<strong>and</strong>ard<br />

substrate the usual simple 4-fold arrangement along the 〈100〉-directions is revealed. The<br />

importance <strong>of</strong> the periodic ripple template with respect to ordering gets clear from Fig. 5.8.<br />

In the growth procedure <strong>of</strong> the samples presented here a flat high-temperature (HT) Si-buffer<br />

was used that lacked the periodic ripples. Thus there is no ordering in miscut direction which


66 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

could be associated with the periodicity <strong>of</strong> step-bunching. Instead, the 50 ˚A Si0.55Ge0.45 epi-<br />

layer deposited at 625 ◦ C on a flat HT Si-buffer reveals a comparable hut-cluster network<br />

along the 〈100〉-directions for the 4 ◦ miscut sample as for a flat ”dummy” substrate. The<br />

guiding lines in Fig. 5.8a indicate the distorted 4-fold ordering due to the large miscut. Ob-<br />

viously several isl<strong>and</strong>s group together <strong>and</strong> share {105}-facets which is typical for hut-clusters<br />

on flat substrates. The missing peak in the FFT (Fig. 5.8a) points out the lack <strong>of</strong> ordering in<br />

miscut direction for the <strong>SiGe</strong>-dots grown on top <strong>of</strong> the HT Si-buffer. For growth on the opti-<br />

mized ripple template the two ordering trends together, namely the network formation along<br />

the 〈100〉-directions <strong>and</strong> the isl<strong>and</strong> nucleation along the step-bunches, establish the basis for<br />

a fair degree <strong>of</strong> ordering <strong>of</strong> the <strong>SiGe</strong>-isl<strong>and</strong>s already in the first <strong>SiGe</strong>-epilayer (Fig. 5.7a).<br />

The AFM images in Fig. 5.9 show a comparison <strong>of</strong> the initial coarsening process in isl<strong>and</strong><br />

growth with a 25 ˚A thin Si0.55Ge0.45 epilayer deposited at 650 ◦ C (1623LSG). The slightly<br />

higher growth temperature favors strain <strong>and</strong> a symmetric shape which can be seen with the<br />

square base <strong>of</strong> the pyramids (flat reference sample, Fig. 5.9b) <strong>and</strong> the regular rhombic shape<br />

due to the miscut on the tilted substrates (Fig. 5.9a). Both samples show poor ordering<br />

<strong>and</strong> size uniformity. Thus it is startling that growth continuation to the usual thickness<br />

<strong>of</strong> 50 ˚A can result in the well-pronounced ordering <strong>and</strong> acceptable size homogeneity for the<br />

Si0.55Ge0.45 isl<strong>and</strong>s on the optimized step-bunching buffer (compare Fig. 5.7a). Obviously a<br />

coarsening process, such as Ostwald ripening, in this case improves the homogeneity <strong>of</strong> the<br />

isl<strong>and</strong>s.<br />

The considerably increased strain in a 25 ˚A thin Si0.25Ge0.75 epilayer leads to an early onset <strong>of</strong><br />

3D-growth <strong>of</strong> dome-shaped <strong>SiGe</strong>-isl<strong>and</strong>s grown at a temperature <strong>of</strong> 600 ◦ C. Fig. 5.10 demon-<br />

strates the striking ordering <strong>of</strong> these Si0.25Ge0.75 isl<strong>and</strong>s along the step-bunches indicated<br />

Figure 5.7: AFM images <strong>and</strong> FFTs showing a fair degree <strong>of</strong> ordering in the 50 ˚A<br />

Si0.55Ge0.45-epilayer deposited at 625 ◦ C (1539LSG) on a 4 ◦ miscut sample (a) <strong>and</strong><br />

a flat dummy substrate (b).<br />

� Source: Ordering n shape 1539LSG.jpg


5.3. ORDERING AND SIZE OF SIGE-ISLANDS 67<br />

Figure 5.8: AFM images <strong>and</strong> FFTs demonstrating the importance <strong>of</strong> the ripple<br />

buffer. Growth on a flat HT Si-buffer reveals a comparable hut-cluster network<br />

along the 〈100〉-directions for the 4 ◦ miscut sample (a) as for a flat dummy substrate<br />

(b). The high miscut only results in a simple distorted 4-fold ordering with a lack<br />

<strong>of</strong> ordering in miscut direction – as indicated by guiding lines <strong>and</strong> the missing peak<br />

in the FFT (a).<br />

� Source: Ordering n shape 1616LSG.jpg<br />

by the sharp peak in miscut direction (FFT in Fig. 5.10a) in comparison to the faint 4-fold<br />

ordering for the reference sample (Fig. 5.10b).<br />

Alignment <strong>of</strong> Ge isl<strong>and</strong>s on faceted Si(001) surfaces <strong>and</strong> high-index-planes is already widely<br />

investigated <strong>and</strong> documented in literature [86, 162, 163, 164]. The AFM data in Fig. 5.11<br />

for pure Ge-dots demonstrate the effect <strong>of</strong> the increase in strain to its ultimate extent. The<br />

Ge-dots (6 ML, @ 0.05 ˚A/s) deposited on top <strong>of</strong> the optimized rippled Si-buffer at 575 ◦ C<br />

Figure 5.9: AFM images <strong>and</strong> FFTs depicting the initial coarsening process <strong>of</strong><br />

isl<strong>and</strong> growth. The slightly thinner Si0.55Ge0.45-epilayer (25 ˚A) deposited at an<br />

increased temperature <strong>of</strong> 650 ◦ C (1623LSG) shows poor ordering <strong>and</strong> a wide size<br />

distribution but a symmetric shape for the hut clusters with a rhombic base for the<br />

miscut sample (a) <strong>and</strong> a square base for the flat substrate (b).<br />

� Source: Ordering n shape 1623LSG.jpg


68 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.10: AFM images <strong>and</strong> FFTs revealing a proper ordering in miscutdirection<br />

for a 25 ˚A thin Si0.25Ge0.75-epilayer deposited at 600 ◦ C. The FFT proves<br />

a rather perfect ordering along the step-bunches for the miscut substrate (a).<br />

� Source: Ordering n shape 1624LSG.jpg<br />

show good size homogeneity with a mean isl<strong>and</strong> size <strong>of</strong> ∼ 25 nm, isl<strong>and</strong> distance <strong>of</strong> ∼ 35 nm<br />

<strong>and</strong> thus a density <strong>of</strong> ∼ 2.5×10 10 cm −2 . Preferential isl<strong>and</strong> nucleation takes place in the ripple<br />

flanks whereas the flat (001)-terraces <strong>of</strong> step-bunching do not show 3D-growth but exhibit flat<br />

denuded zones (zoom-in: Fig. 5.11c). Against first expectations, there are no small Ge-wires<br />

spanning down the ripple-flanks but there are only these small separate isl<strong>and</strong>s. Obviously<br />

the local miscut in the ripple-flanks is not high enough, so that the dots cannot merge to<br />

wires. The temperature ramp-up to 575 ◦ C is already too high <strong>and</strong> thus the amplitude <strong>of</strong> the<br />

ripple pattern <strong>and</strong> the steep flanks degrade. On the other h<strong>and</strong> for the nucleation <strong>of</strong> Ge-dots<br />

the growth temperature <strong>of</strong> 575 ◦ C is quite low <strong>and</strong> causes a high density <strong>of</strong> small isl<strong>and</strong>s.<br />

Figure 5.11: AFM data in the usual height representation (a) <strong>and</strong> in derivative<br />

mode (b). The Ge-dots (6 ML, @ 0.05 ˚A/s) deposited on top <strong>of</strong> the optimized rippled<br />

Si-buffer at 575 ◦ C show good size homogeneity. Preferential isl<strong>and</strong> nucleation takes<br />

place in the ripple flanks whereas the flat (001)-terraces <strong>of</strong> step-bunching do not<br />

show 3D-growth but exhibit flat denuded zones (c).<br />

� Source: Ordering n shape 1549LSG miscut.jpg


5.3. ORDERING AND SIZE OF SIGE-ISLANDS 69<br />

Figure 5.12: Series <strong>of</strong> AFM images for Ge-dot growth at various temperatures<br />

<strong>and</strong> a different layer thickness. The 4 ◦ miscut substrates (a-e) provide compared<br />

to untilted Si(001) substrates (f-j) a slightly wider temperature window where homogeneous<br />

Ge-dots can be found. The deposition <strong>of</strong> 6 ML Ge at a rate <strong>of</strong> 0.05 ˚A/s<br />

shows the best size homogeneity for the ripple template around 575 ◦ C (c).<br />

� Source: Ordering n shape Ge-dots miscut n dummy.jpg


70 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Fig. 5.12 shows a series <strong>of</strong> AFM images for Ge-dot growth at various temperatures <strong>and</strong> a<br />

different layer thickness. The deposition <strong>of</strong> 6 ML Ge at a rate <strong>of</strong> 0.05 ˚A/s shows the best<br />

size homogeneity for the ripple template around 575 ◦ C (Fig. 5.12c). For increasing deposi-<br />

tion temperatures the influence <strong>of</strong> the step-bunching template fades away <strong>and</strong> the ordinary<br />

coarsening <strong>and</strong> multi-modal size distribution is found (Fig. 5.12a-d). As for the flat reference<br />

substrate (Fig. 5.12f-i) pyramids, domes <strong>and</strong> super-domes showing the usual asymmetries due<br />

to the substrate miscut are already seen for a moderate Ge-layer thickness <strong>of</strong> 6 ML. The in-<br />

creased strain in a 8 ML thick Ge-layer leads to an irregular 3D-growth for the flat ”dummy”<br />

substrate (Fig. 5.12j) <strong>and</strong> for the miscut sample (Fig. 5.12e) as well, even at a growth temper-<br />

ature lowered to 550 ◦ C. The densely stepped miscut sample <strong>of</strong>fers plenty <strong>of</strong> equally favorable<br />

nucleation sites in a kinetically dominated growth regime. On flat substrate regions only a<br />

few seeds as attractive sites for material incorporation are available. Whatever substrate is<br />

used for higher temperatures <strong>and</strong> increased diffusion lengths the existing seeds capture most<br />

<strong>of</strong> the Ge-adatoms. The competition <strong>of</strong> the large attractive isl<strong>and</strong>s leads to the inordinate<br />

coarsening when thermodynamics rule. In summary, miscut substrates provide compared<br />

to untilted Si(001) a slightly wider temperature window around 575 ◦ C where homogeneous<br />

Ge-dots can be found.<br />

Miscut Si(001) may be exploited to achieve small, regular <strong>and</strong> very dense Ge-dots which<br />

are <strong>of</strong> interest for optical applications. In this approach neither use <strong>of</strong> surfactant-mediated<br />

growth [165, 166, 167, 168, 169, 170, 171] nor <strong>of</strong> extremely low growth temperatures [172, 173]<br />

is made to shrink down the dot size.<br />

5.4 Closer Look on Surface Energy Effects – Facetting<br />

This illustrative part comprises several AFM-images which show different states <strong>of</strong> facetting<br />

for various epilayers. Not only details <strong>and</strong> morphological features in well-known st<strong>and</strong>ard<br />

structures like pyramids <strong>and</strong> domes were revealed. Further examples demonstrate effects<br />

<strong>of</strong> energy minimization <strong>and</strong> facetting found by chance, when unexpected sample cleaning<br />

problems were encountered.<br />

Fig. 5.13 depicts AFM data <strong>of</strong> 50 ˚A thick Si0.55Ge0.45 epilayers grown at 625 ◦ C (1580LSG)<br />

<strong>and</strong> 700 ◦ C (1581LSG), respectively. These AFM images show the morphology <strong>of</strong> the untilted<br />

reference substrates. Details regarding the layer structure <strong>of</strong> sample 1580LSG are summarized<br />

in Tab. 5.2. The smooth Si buffer was grown with the usual ”optimized” parameters (see<br />

Tab. 5.1). Sample 1580LSG was especially grown to check the reproducibility with our MBE-<br />

system after updating our Ge-evaporation assembly (see Ch. 8).


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 71<br />

Figure 5.13: AFM data <strong>of</strong> 50 ˚A thick Si0.55Ge0.45 epilayers grown at 625 ◦ C<br />

(1580LSG) <strong>and</strong> 700 ◦ C (1581LSG), respectively. Conventional topographical data<br />

(a, e) <strong>and</strong> data recorded in derivative mode (b, f, i) are complemented with facet<br />

evaluation plots (SOM: d, h; SAP: c, g, j) for the flat reference samples.<br />

� Source: Facetting d1580 d1581.jpg


72 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Sample 1580LSG (1539LSG) 1636LSG 1635LSG<br />

Growth interrupt 425 ◦ C → 625 ◦ C 425 ◦ C → 625 ◦ C 425 ◦ C → 575 ◦ C<br />

(5 min) (5 min) (5 min)<br />

<strong>SiGe</strong>/Ge epilayer 50 ˚A Si0.55Ge0.45 150 ˚A Si0.75Ge0.25 6 ML Ge<br />

Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s<br />

Ge @ 0.1636 ˚A/s Ge @ 0.0667 ˚A/s Ge @ 0.05 ˚A/s<br />

625 ◦ C 625 ◦ C 575 ◦ C<br />

Table 5.2: Growth parameters for <strong>SiGe</strong>/Ge-dot layers on ”optimized” buffer.<br />

The conventional AFM data give the usual topographical view with height information<br />

(Fig. 5.13a), whereas the derivative recording mode during AFM operation is more sensi-<br />

tive to height fluctuations <strong>and</strong> <strong>of</strong>fers a better signal-to-noise ratio (Fig. 5.13b). Nevertheless,<br />

from these data only the changes in slope along the scan-direction are accessible. For this<br />

reason the topographic AFM data are exploited to evaluate facetting. In Fig. 5.13c the calcu-<br />

lated local slope in each AFM data point is visualized in a surface-angle-plot (SAP), whereas<br />

the already introduced histogram <strong>of</strong> specific orientations (see Sec. 5.2.2) is presented in a<br />

surface-orientation-map (SOM) Fig. 5.13d. The upper series <strong>of</strong> images (Fig. 5.13a-d) for the<br />

50 ˚A thick Si0.55Ge0.45 epilayer grown at 625 ◦ C shows bimodal isl<strong>and</strong>s with typical facetting<br />

for the flat reference Si(001) substrate. Although {105}-facets are dominant, also {113}- <strong>and</strong><br />

{15 3 23}-facets can be clearly identified [156, 174, 175, 176, 177]. Steeper facets, such as<br />

the {111}-type, are not observed here since these are usually found only for diluted Si1−xGex<br />

isl<strong>and</strong>s (x < 0.2) [178]. The circles in Fig. 5.13d serve as guides to the eye <strong>and</strong> indicate surface<br />

orientation angles for the well-known facets (from out to in: {15 3 23}, {113}, {105}). At the<br />

higher temperature <strong>of</strong> 700 ◦ C (1581LSG) the isl<strong>and</strong>s grow larger <strong>and</strong> are no longer in touch<br />

at the base (Fig. 5.13e-j), as is the case for the 625 ◦ C sample (1580LSG). The presence <strong>of</strong><br />

extended denuded zones around the widely separated isl<strong>and</strong>s results in an intensified central<br />

x y z x y z angle [�]<br />

0 0 1 1 1 1 54.74<br />

15 3 23 33.63<br />

substrate 1 1 3 25.24<br />

(reference) 1 0 5 11.31<br />

1 1 10 8.05<br />

Table 5.3: Relevant facets <strong>and</strong> angles in the <strong>SiGe</strong>-system.


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 73<br />

Figure 5.14: Topographical (a) <strong>and</strong> derivative mode AFM data (b) <strong>of</strong> a 25 ˚A thin<br />

Si0.55Ge0.45 epilayer grown at 650 ◦ C (1623LSG). The evaluation <strong>of</strong> the local slope<br />

is visualized in a surface-angle-plot (c) <strong>and</strong> a histogram <strong>of</strong> specific orientations is<br />

presented in a surface-orientation-map (d). A remarkable morphological feature is<br />

the splitting <strong>of</strong> the down-hill edge at the base <strong>of</strong> the predominantly {105}-faceted<br />

pyramids.<br />

� Source: Facetting m1623.jpg<br />

spot in the SOM image indicating the (001)-orientation <strong>of</strong> the Si(001) substrate. Generally<br />

the spots get more distinct for the higher growth temperature as the facet areas increase to-<br />

gether with the isl<strong>and</strong> size. The zoom-in images for 700 ◦ C (Fig. 5.13i-j) show several stages in<br />

the pyramid-to-dome transition. Although the upper part is typically dome-like at the base<br />

extended foothills <strong>and</strong> remainders <strong>of</strong> pyramids are present [179]. The edges <strong>of</strong> the pyramids<br />

exhibit partially a fringy shape. Seemingly, this is the way the pyramidal edges are dissolved<br />

at the base, <strong>and</strong> the boundary lines are moved in <strong>and</strong> converted from a pure 〈100〉- to a<br />

finally achieved 〈110〉-orientation (compare Fig. 5.15 <strong>and</strong> subsequent discussion). Steepening<br />

occurs mainly near the pyramid top where steps on {105}-facets bunch <strong>and</strong> {113}-facets can<br />

be formed [180]. The surface angles <strong>of</strong> the isl<strong>and</strong>s (Fig. 5.13j) agree well with the calculated<br />

values listed in Tab. 5.3.<br />

The fringy structure correlated with the shape transition was also observed for a 25 ˚A thin<br />

Si0.55Ge0.45 epilayer grown at 650 ◦ C (for details see Tab. 5.4, 1623LSG). The 4 ◦ miscut <strong>of</strong> the


74 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.15: Model <strong>of</strong> {105}-faceted pyramid with one split edge. A region<br />

<strong>of</strong>fering a steep slope is generated automatically where the pyramidal edge meets<br />

the groove between the fringes which could be an alternative growth center for a<br />

{113}-facet.<br />

� Source: Facetting edge-splitting model.jpg<br />

sample is the reason for the rhombic distortion <strong>of</strong> the pyramidal base [158, 181, 182] as already<br />

discussed for Fig. 5.9. This asymmetry causes the higher intensity for the larger down-hill<br />

{105}-facets in the SOM-representation (Fig. 5.14d) compared to the small upper facets. Es-<br />

pecially from the derivative-mode AFM data (Fig. 5.14b) <strong>and</strong> the SAP-image (Fig. 5.14c) the<br />

remarkable morphological feature, namely the splitting <strong>of</strong> the down-hill edge at the base <strong>of</strong><br />

the predominantly {105}-faceted pyramids, gets obvious. The edge is split up in two fringes<br />

which are again {105}-faceted. It is indisputable that the longer down-hill edge would show<br />

up further steepening to relief excessive strain first. Developing this splitting gives an easy<br />

access to transform the main base-orientation from 〈100〉 to 〈110〉 which is necessary to form<br />

the next steeper {113}-facet in the base region. By the splitting a region <strong>of</strong> steep slope has to<br />

be generated where the pyramidal edge meets the groove between the fringes. This could be<br />

seen as a growth center for a {113}-facet which has to be nucleated to proceed with the tran-<br />

sition from a pyramid- to dome-shaped isl<strong>and</strong>. Fig. 5.15 schematically depicts a 3D-model for<br />

this proposed mechanism. Nevertheless there is no evidence from the AFM data (Fig. 5.14d)<br />

which clearly supports this idea, <strong>and</strong>, as stated above, it is quite obvious from more detailed<br />

STM images that the main pyramid-to-dome transition takes place via bunching <strong>of</strong> steps <strong>and</strong><br />

growth from top to bottom [180]. Nevertheless the proposed model minimizes surface en-<br />

ergy <strong>and</strong> enables strain relief during the process <strong>of</strong> pyramidal edge dissolution. Additionally<br />

material is provided for the rearrangement <strong>of</strong> atoms also for the upper part <strong>of</strong> an isl<strong>and</strong> to


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 75<br />

realize finally an energetically stable dot shape.<br />

Facet generation <strong>and</strong> evolution is also found for Si-overgrowth <strong>of</strong> already formed <strong>SiGe</strong>-isl<strong>and</strong>s.<br />

Fig. 5.16 shows AFM data illustrating the morphological changes <strong>of</strong> buried <strong>SiGe</strong>-dots with<br />

dependence on Si-cap thickness <strong>and</strong> growth temperature (details in Tab. 5.4, 1609LSG). Low-<br />

temperature Si-capping results in a reversed isl<strong>and</strong> transition from dome- to pyramid-type<br />

shape [179, 183, 184, 185]. With increasing Si-layer thickness the truncated pyramids ex-<br />

tend mainly in lateral size whereas the height decreases slowly (Fig. 5.16a-b, d-e). On both<br />

substrate types the extending isl<strong>and</strong>s partially merge. For the miscut samples this leads to<br />

a formation <strong>of</strong> bunches, which exhibit still numerous constrictions (Fig. 5.16a-b). By ap-<br />

plying a temperature ramp up to 550 ◦ C during Si-capping the structure height is strongly<br />

reduced due the increased adatom diffusion giving faint step-bunches for the 4 ◦ miscut tem-<br />

plate (Fig. 5.16c) <strong>and</strong> a flat surface with step-flow growth along the 〈110〉-directions for the<br />

”dummy” Si(001) substrate (Fig. 5.16f). The change in facetting is revealed by small-scale<br />

AFM data in the early stage <strong>of</strong> Si-capping after the deposition <strong>of</strong> 50 ˚A Si at 425 ◦ C on top <strong>of</strong><br />

the optimized <strong>SiGe</strong>-isl<strong>and</strong> epilayer for a flat untilted substrate (Fig. 5.16g). The AFM-data in<br />

derivative mode (Fig. 5.16h) <strong>and</strong> the surface-orientation-map (Fig. 5.16i) indicate a seemingly<br />

octagonal base shape with facets along 〈110〉- <strong>and</strong> 〈100〉-directions. For an increased Si-cap<br />

thickness (250 ˚A) the transition from Ge-like to Si-favored high-index {11n}-type facets is<br />

completed <strong>and</strong> the base adopts a square shape (Fig. 5.16j-l). These truncated pyramids are<br />

thus arranged along a 〈110〉-direction, which is rotated by 45 ◦ with respect to the usual<br />

orientation <strong>of</strong> Ge pyramids or hut-clusters.<br />

Sample 1684LSG 1623LSG 1609LSG<br />

Growth interrupt – 425 ◦ C → 650 ◦ C 425 ◦ C → 625 ◦ C<br />

(5 min) (5 min)<br />

LT-Si Buffer / 50 ˚A Si 25 ˚A Si0.55Ge0.45 50 ˚A Si0.55Ge0.45<br />

<strong>SiGe</strong> epilayer Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s<br />

Ge @ 0.1636 ˚A/s Ge @ 0.1636 ˚A/s<br />

425 ◦ C → 350 ◦ C 650 ◦ C 625 ◦ C<br />

Growth interrupt – – 625 ◦ C → 425 ◦ C<br />

(5 min)<br />

Si-cap – – 100 ˚A + 50 ˚A + 100 ˚A Si<br />

Si @ 0.2 ˚A/s<br />

425 ◦ C → 550 ◦ C<br />

Table 5.4: LT-Si buffer <strong>and</strong> <strong>SiGe</strong>-epilayer growth with optional Si-capping.


76 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.16: AFM data illustrating the morphological changes <strong>of</strong> buried <strong>SiGe</strong>dots<br />

with dependence on Si-cap thickness <strong>and</strong> growth temperature for 4 ◦ miscut<br />

(a-c) <strong>and</strong> flat reference substrates (d-f). For increased Si-cap thickness facets change<br />

from Ge-like (g-i) to a Si-favored {11n}-type (j-l).<br />

� Source: Facetting Si-capping.jpg


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 77<br />

No clear facets can be assigned, which indicates the existence <strong>of</strong> stepped mounds [183].<br />

Steep slopes <strong>of</strong>fering angles up to those <strong>of</strong> {113}-facets are measured. Slight distortions <strong>and</strong><br />

imperfections in the 500 nm AFM-images <strong>and</strong> the calculated data based thereon are caused<br />

by sample drift during the AFM measurements (see e.g. Fig. 5.16j-l).<br />

5.4.1 Surface Energy Minimization via Excessive Facetting<br />

It was already discussed in Sec. 5.2.2 that at moderate temperatures the ripple flanks <strong>of</strong><br />

step-bunching are decorated by {105}-faceted ridges (see Fig. 5.5). There the whole surface<br />

is purely made up with (001)- <strong>and</strong> {105}-facets. Already from this, the assumption is very<br />

close, that any 3D-object, whether it is protruding from the Si-surface or recessed into it, can<br />

be fully lined with {105}-faceted features. The only condition, which has to be met, is, that<br />

the inclination angle <strong>of</strong> the surfaces, which confine the rough overall shape <strong>of</strong> these struc-<br />

tures, are close to {1 1 10}-planes. Meanwhile, such a behavior <strong>and</strong> comparable morphological<br />

features were found for overgrown 2D-pit-patterned templates by G. Chen [91, 186] <strong>and</strong> Z.<br />

Zhong [93] or for Ge-deposition on large-scale spherical dimples demonstrated by Watanabe<br />

et al. [159, 160]. Pit-patterned Si-templates are currently studied to explore the influence <strong>of</strong><br />

the faceted ridges at the slopes <strong>of</strong> the pits on the mechanism <strong>of</strong> Ge-dot nucleation [186].<br />

In this thesis, by chance, such faceted structures were found due to failed sample pre-<br />

cleaning procedures. Although it took quite a while to locate the origin <strong>of</strong> the problem in the<br />

DI-water system, this was finally not so unfortunate, as this way interesting morphologies<br />

could be investigated. Fig. 5.17 presents AFM data in conventional <strong>and</strong> derivative mode <strong>of</strong> a<br />

150 ˚A thick Si0.75Ge0.25 epilayer deposited on flat reference sample (see Tab. 5.2, 1636LSG).<br />

Surface contamination hinders perfect 2D layer-by-layer growth <strong>and</strong> several extended hills<br />

<strong>and</strong> holes are created instead. The flanks <strong>of</strong> the elevated structures exhibit angles around<br />

∼ 10 ◦ which gets clear from the SOM- <strong>and</strong> SAP-image representation <strong>of</strong> the plain AFM-data<br />

(see Fig. 5.17a-b, e, g). The same is seen for zoom-in AFM-data <strong>and</strong> its corresponding facet<br />

evaluation plots (Fig. 5.17c, f, h) revealing the morphological details <strong>of</strong> a pyramid located in<br />

the center <strong>of</strong> a pit. The major contribution <strong>of</strong> the surface is made up with {105}-facets<br />

which is typical for Ge-rich epilayers. The low Ge-content, however, promotes also Si-type<br />

facets at the edges <strong>of</strong> the pyramids <strong>and</strong> in the corners <strong>of</strong> the pits. Obviously sharp edges are<br />

energetically less favorable <strong>and</strong> flattened edges exhibiting {1 1 10}-facets are preferable for<br />

large <strong>SiGe</strong>-structures formed within a low-Ge-alloy.<br />

The sample substrate deficiencies originating from improper cleaning are also manifested in<br />

the AFM images <strong>of</strong> an epilayer consisting <strong>of</strong> 6 ML Ge grown at 575 ◦ C (Fig. 5.18 <strong>and</strong> Fig. 5.19).<br />

Fig. 5.18 gives with a 5 µm AFM scan an overview <strong>of</strong> the interesting morphology on a flat


78 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.17: AFM data in conventional (a, d) <strong>and</strong> derivative mode (b-c) <strong>of</strong> 150 ˚A<br />

Si0.75Ge0.25 epilayer on flat reference sample (1636LSG). The morphology evidences<br />

the cleaning problems at that time with faceted hill <strong>and</strong> hole structures. Due to the<br />

low Ge-content the pyramid edges are not sharp but flattened <strong>and</strong> exhibit {1 1 10}facets.<br />

This gets clarified by surface-angle-plots (SAP: e-f) <strong>and</strong> surface-orientationmaps<br />

(SOM: g-h).<br />

� Source: Facetting <strong>SiGe</strong>-features.jpg<br />

reference Si(001) substrate influenced by the failed pre-cleaning procedure. The Ge-film dec-<br />

orates the hills <strong>and</strong> holes in the underlying Si-buffer (for layer details see again Tab. 5.2,<br />

1635LSG). Small-scale AFM based data captured in conventional <strong>and</strong> derivative-mode are<br />

complemented with SAP-images representing hill- (Fig. 5.19a-f) <strong>and</strong> hole- (Fig. 5.19g-l) struc-<br />

tures that are fully lined with {105}-faceted ridges to minimize energy. The large truncated


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 79<br />

Figure 5.18: AFM image <strong>of</strong> 6 ML Ge grown at 575 ◦ C on a flat reference Si(001)<br />

substrate (1635LSG). The 5 µm scan gives an overview <strong>of</strong> the interesting morphology<br />

influenced by the failed cleaning procedure. The Ge-film decorates the hills <strong>and</strong><br />

holes in the underlying Si-buffer.<br />

� Source: Facetting Ge-features01.jpg<br />

pyramids featuring {1 1 10}-faceted flanks are well suited for the nucleation <strong>of</strong> {105}-faceted<br />

ridges (compare Fig. 5.5). The flat Si-substrate surrounding the hills <strong>and</strong> the wide plateaus<br />

on top <strong>of</strong> the Si-pyramids are decorated by the usual small Ge-pyramids. According to the bi-<br />

modal growth regime also Ge-domes are found. The small holes in the Si-buffer yield perfect<br />

nucleation sites for Ge-dots as can be seen in Fig. 5.19g-i. The side-walls <strong>of</strong> the pits are only<br />

made up <strong>of</strong> {105}-facets <strong>and</strong> therefore exhibit the same but inverted structure which is found<br />

for the hills (Fig. 5.19j-l). A similar behavior <strong>and</strong> comparable morphological features were<br />

found for overgrown 2D-pit-patterned templates by G. Chen [91, 186] <strong>and</strong> Z. Zhong [93] or for<br />

Ge-deposition on large-scale spherical dimples demonstrated by Watanabe et al. [159, 160].<br />

Fig. 5.20 shows AFM results <strong>of</strong> an Sb-mediated Ge-dot double-layer which was grown together<br />

with a guest scientist, M. M. Rzaev (layer details are listed in Tab. 5.5, 1680RSG). Although<br />

the original intent <strong>of</strong> this layer was to generate small Ge-dots for photoluminescence studies<br />

the layer reveals also in the respect <strong>of</strong> facetting interesting features which again originate<br />

from the failed wafer cleaning procedure. On the overall flat parts many small Ge-isl<strong>and</strong>s are<br />

formed because the diffusion length is reduced by the surfactant antimony (Sb) <strong>and</strong> the rel-<br />

atively low growth temperature for Ge-deposition (500 ◦ C). Although quite at the resolution<br />

limit <strong>of</strong> the AFM, the isl<strong>and</strong>s exhibit the usual pyramidal shape (Fig. 5.20b) which is already<br />

well-experienced [167]. A closer look on the deep, almost circular holes with a diameter <strong>of</strong><br />

about ∼ 150 nm reveals different remarkable details (Fig. 5.20c-e). The conventional AFM<br />

image (Fig. 5.20c) shows densely packed Ge-dots at the rim <strong>of</strong> a hole, which seem to extend<br />

down the upper part <strong>of</strong> the hole with probably {105}-faceted ridges forming a radial structure<br />

in the derivative-mode AFM image (Fig. 5.20d).


80 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.19: Small-scale AFM-based data <strong>of</strong> the 6 ML Ge film grown at 575 ◦ C<br />

(Fig. 5.18, 1635LSG). Conventional, derivative-mode AFM data <strong>and</strong> surface-angleplots<br />

<strong>of</strong> hill- (a-f) <strong>and</strong> hole- (g-l) structures lined with {105}-facets to minimize<br />

energy.<br />

� Source: Facetting Ge-features02.jpg


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 81<br />

Figure 5.20: AFM data <strong>of</strong> the Sb-mediated Ge-dot top-layer grown at 500 ◦ C<br />

together with M. M. Rzaev (1680RSG). The defective growth arising from failed<br />

substrate cleaning leads to faceted holes featuring ”inverted” dome-like shapes.<br />

� Source: Facetting GeSb.jpg<br />

The inner part <strong>of</strong> the hole is also purely faceted <strong>and</strong> <strong>of</strong>fers facets usually found on domes. All<br />

the major facets as {105}, {113} <strong>and</strong> {15 3 23} can clearly be distinguished (Fig. 5.20e). The<br />

presented hole is obviously in its structural appearance an inverted dome. The formation <strong>of</strong><br />

this structure is remarkable as strain relaxation seems to be not evident for a hole. Clearly<br />

convex surface parts are more favorable for strain relaxation compared to concave regions.<br />

But on the microscopic scale the effect <strong>of</strong> surface curvature is also important for the surface<br />

energy in terms <strong>of</strong> atomic bonding. In a concave region an atom has, on average, more<br />

neighbors, which reduces the local surface energy [149]. Obviously in the present case the<br />

(”inverted”) dome structure, which was also found by G. Chen <strong>and</strong> Z. Zhong, is dominated<br />

by the surface energy, <strong>and</strong> strain plays a subordinate role.<br />

The experimentally observed multi-faceted hill- <strong>and</strong> pit-structure is visualized in Fig. 5.21<br />

with 3D-models. For comparison Fig. 5.21a shows the 3D-representation <strong>of</strong> the AFM data<br />

already presented in Fig. 5.19a-c next to the simplified hill structure which is modeled as an<br />

extruding truncated pyramid (Fig. 5.19b, left). The underlying shape is bound by {1 1 10}-<br />

oriented side walls, has thus a square base along the 〈110〉-directions <strong>and</strong> is fully decorated


82 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

Figure 5.21: 3D-AFM data <strong>of</strong> a multi-faceted hill (see Fig. 5.19a) <strong>and</strong> models. An<br />

extruding truncated pyramid (b, left) with {1 1 10}-oriented side walls, <strong>and</strong> thus a<br />

square base along the 〈110〉-directions, is fully decorated with {105}-faceted ridges.<br />

The whole surface is only made up with (001)- <strong>and</strong> {105}-facets. Also the inverted<br />

pit-like structure (b, right) was found experimentally (see Fig. 5.19j).<br />

� Source: Facetting model.jpg<br />

with {105}-faceted ridges. The whole surface is purely made up with (001)- <strong>and</strong> {105}-<br />

facets (see especially the encircled edge in Fig. 5.19a). The same is true for the inverted<br />

pit-like structure (Fig. 5.19b, right) that was also found in experiments (compare Fig. 5.19j-l).<br />

These structures are currently investigated systematically on periodically 2D pit-patterned<br />

Sample 1680RSG<br />

Si Buffer 250 ˚A + 500 ˚A + 250 ˚A Si<br />

Si @ 1.0 ˚A/s<br />

750 ◦ C → 500 ◦ C<br />

Ge-dot layer 1 4 ML Ge + 2 ML Ge:Sb<br />

Ge @ 0.05 ˚A/s<br />

Sb @ 310 ◦ C (see Fig. A.4)<br />

500 ◦ C<br />

Si Spacer 200 ˚A + 800 ˚A Si<br />

Si @ 1.0 ˚A/s<br />

450 ◦ C → 500 ◦ C, 500 ◦ C<br />

Ge-dot layer 2 4 ML Ge + 2 ML Ge:Sb<br />

Ge @ 0.05 ˚A/s<br />

Sb @ 310 ◦ C<br />

500 ◦ C<br />

Table 5.5: Growth parameters for Sb-mediated Ge-dot layers.


5.4. CLOSER LOOK ON SURFACE ENERGY EFFECTS – FACETTING 83<br />

Figure 5.22: AFM images <strong>of</strong> 2D pit-patterned Si(001) templates with 4 ML Ge<br />

(@ 0.03 ˚A/s) deposited at 620 ◦ C. (a) 3D-AFM image showing an array <strong>of</strong> six pits<br />

with faulty lithography which led to a spread in the depth <strong>of</strong> the etch pits. (b)<br />

Laplace transformation revealing the {105}-faceted morphology for the side-walls<br />

<strong>and</strong> corners <strong>of</strong> the pits as schematically shown in Fig. 5.21b. (taken from [186])<br />

� Source: Facetting pit-patterned substrates.jpg<br />

substrates by G. Chen (see Fig. 5.22 taken from [186]). They are attributed to play an<br />

important role in the initial stage <strong>of</strong> the 2D-3D transition <strong>of</strong> a strained <strong>SiGe</strong> layer on a pit-<br />

patterned Si(001) template. Fig. 5.22 depicts an array <strong>of</strong> six pits from a part <strong>of</strong> an overgrown<br />

pit-patterned Si(001) substrate. Fig. 5.22a shows a 3D-AFM image <strong>of</strong> these pits for 4 ML Ge<br />

(@ 0.03 ˚A/s) deposited at 620 ◦ C with faulty lithography leading to a spread in the depth <strong>of</strong><br />

the etch pits. The Laplace transformation in Fig. 5.22b reveals that in the lower pit row Ge<br />

pyramids have already developed in the center, <strong>and</strong> the full symmetry <strong>of</strong> the morphological<br />

features both on the pit walls <strong>and</strong> in the corners has formed. The upper pits are shallower<br />

<strong>and</strong> show merely the staircase-like surface corrugations in the corners <strong>of</strong> the pits. This<br />

inhomogeneous region <strong>of</strong> the pit-pattern already demonstrates several stages at the 2D-3D<br />

transition where the {105}-faceted ridges seem to be important (compare schematic view<br />

in Fig. 5.21b). [186]<br />

5.4.2 Step-Bunching Templates for p-Modulation Doped <strong>SiGe</strong>-Structures<br />

A major goal <strong>of</strong> the thesis was also to investigate p-modulation doped <strong>SiGe</strong>-structures on step-<br />

bunching templates to measure an expected anisotropy in mobility (see Ch. 6). The carriers,<br />

holes confined in the <strong>SiGe</strong>-channel, should feel an additional scattering potential due to the<br />

corrugation <strong>of</strong> the step-bunches for the current-flow perpendicular to step-bunching. As pre-<br />

experiments showed the <strong>SiGe</strong>-channel has to be deposited at rather low temperatures around<br />

350 ◦ C to suppress strain-driven morphological features <strong>and</strong> to form a smooth conformal<br />

<strong>SiGe</strong>-layer (see Fig. 5.3). To avoid a growth interruption the Si-shutter was kept open during


84 CHAPTER 5. SELF-ORGANIZED GROWTH 2 – KINETICS AND STRAIN<br />

temperature ramp-down to 350 ◦ C <strong>and</strong> this way a 50 ˚A thin low-temperature Si-layer was<br />

grown on top <strong>of</strong> the optimized ripple buffer (details are listed in Tab. 5.2, 1684LSG). The<br />

AFM measurements (Fig. 5.23) <strong>and</strong> especially the evaluation with the corresponding surface-<br />

orientation-map (SOM, Fig. 5.23d) <strong>and</strong> surface-angle-plot (SAP, Fig. 5.23e) revealed that the<br />

step-bunching structure on the 4 ◦ miscut sample exhibits in the ripple flanks regions <strong>of</strong><br />

increased slope approaching {113}-facets. This steepening obviously occurred during the<br />

growth <strong>of</strong> the 50 ˚A LT-Si layer as the optimized Si-buffer grown at 425 ◦ C exhibits only step-<br />

bunches with angles reaching up to 15 ◦ (Fig. 4.9). As artifacts again holes appear in the<br />

Si-buffer which mark unintentional nucleation sites caused by bad sample pre-cleaning. The<br />

small holes are in the center <strong>of</strong> extended humps as the incoming adatoms cannot escape the<br />

vicinity <strong>of</strong> the holes due to the lowered diffusion length at 425 ◦ C (Fig. 5.23a).<br />

Several experiments in this part show that pronounced facetting can also occur for low<br />

growth temperatures (Fig. 5.23, Fig. 5.20c-e). This happens whenever surface energetics rule<br />

at least locally <strong>and</strong> thermodynamical disorder or kinetics do not smoothen the morphology.<br />

Figure 5.23: AFM data <strong>of</strong> a 50 ˚A LT-Si layer grown on top <strong>of</strong> the optimized<br />

ripple buffer (1684LSG). Conventional height (a-b) <strong>and</strong> derivative-mode (c) data<br />

are complemented with a surface-orientation-map (d) <strong>and</strong> a surface-angle-plot (e).<br />

Especially the latter reveal that the step-bunching structure on the 4 ◦ miscut sample<br />

exhibits in the ripple flanks regions <strong>of</strong> increased slope approaching {113}-facets (de).<br />

Again detrimental cleaning artifacts are visible as holes in the Si-buffer (a).<br />

� Source: Facetting LT-Si.jpg


Chapter 6<br />

p-Modulation Doping <strong>and</strong> Mobility<br />

Analysis<br />

In this chapter first results on p-modulation doped Si/<strong>SiGe</strong> heterostructures grown on top <strong>of</strong><br />

a rippled step-bunching Si-buffer are outlined. The short-scale periodic height fluctuations<br />

<strong>of</strong> the Si-buffer – which were extensively discussed in the previous chapters – are intended<br />

to form well-defined undulations in the <strong>SiGe</strong>-channel <strong>of</strong> the remotely p-doped quantum well<br />

giving rise to increased scattering. Thus an asymmetry in mobility perpendicular <strong>and</strong> par-<br />

allel to the undulations is expected which might help to uncouple the different scattering<br />

mechanisms which are conversely discussed as predominant hole-mobility limiting factors for<br />

p-modulation doped structures, namely alloy scattering <strong>and</strong> interface-roughness related scat-<br />

tering.<br />

Earlier experiments in this direction were published by Waltereit et al. [187] for n-modulation<br />

doped Si/<strong>SiGe</strong> heterostructures grown on vicinal Si(001) substrates on top <strong>of</strong> a composition-<br />

ally graded strain-relaxed Si0.72Ge0.28 buffer. Also by Neumann et al. [188, 189] anisotropic<br />

hole transport measurements on p-modulation doped <strong>SiGe</strong> channels on step-bunched vicinal<br />

Si(113) surfaces were reported. These however were performed on Si(113) which shows strong<br />

step-bunching but will hardly become <strong>of</strong> technical relevance. Our investigations are based on<br />

Si(001) substrates with a miscut <strong>of</strong> 4 ◦ which are also used commercially [8].<br />

The modulation-doped quantum well structures (MODQW) grown on a step-bunching tem-<br />

plate enables surface roughness dependent measurements on one <strong>and</strong> the same sample which<br />

makes the experiment <strong>and</strong> interpretation less sensitive to other growth-process- or sample<br />

processing-induced artifacts: especially background impurity scattering is known to be hard<br />

to control giving unpredictable results.<br />

85


86 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.1: Layer sequences typically employed for (a) n-type MODQWs with<br />

Si-channel <strong>and</strong> strain-adjusting <strong>SiGe</strong> buffer layer, (b) pseudomorphic p-MODQWs<br />

with <strong>SiGe</strong>-channel, <strong>and</strong> (c) p-MODQWs with pure Ge (or Ge-rich <strong>SiGe</strong>) channel on<br />

relaxed <strong>SiGe</strong> buffer layer. Depicted is the conventional architecture with frontside<br />

doping. [6]<br />

� Source: p-mod MODQW.jpg<br />

Although still at the beginning, the first measurements confirm a decreased low-temperature<br />

mobility across the undulations by nearly a factor <strong>of</strong> two. Such a remarkable effect was<br />

beyond expectations for the Si(001) surface. Further measurements with altered parameters<br />

such as carrier density ps, ripple period Λ <strong>and</strong> ripple amplitude H <strong>and</strong> especially Ge-content<br />

x remain to be done. The reported results, <strong>and</strong> their continuation combined with additional<br />

modeling are expected to provide a new approach toward settling the long-lasting dispute on<br />

the limiting scattering mechanisms <strong>of</strong> the hole mobility in p-MODQW structures.<br />

6.1 Introduction to p-Modulation Doped Si/<strong>SiGe</strong> Heterostruc-<br />

tures <strong>and</strong> Hole Mobility Limitations<br />

Modulation-doped Si/<strong>SiGe</strong> heterostructures were first realized in 1984 with a <strong>SiGe</strong> quantum<br />

well s<strong>and</strong>wiched between the Si substrate <strong>and</strong> an unstrained Si-cap layer [193]. Selective<br />

p-type doping in the Si cladding layers leads to an enhanced hole mobility for the estab-<br />

lished two-dimensional hole gas (2DHG) in the <strong>SiGe</strong>-channel which is formed according to<br />

the valence b<strong>and</strong> <strong>of</strong>fset. Historically later, n-doped structures featuring a two-dimensional<br />

electron gas (2DEG) were fabricated. Employing relaxed virtual Si1−xGex substrates, an<br />

in-plane tensilely strained Si-channel is formed due to the conduction b<strong>and</strong> <strong>of</strong>fset. Such re-


6.1. INTRODUCTION TO P-MODULATION DOPING 87<br />

Figure 6.2: Valence-b<strong>and</strong> (VB) <strong>and</strong> layer structure <strong>of</strong> a p-modulation doped<br />

sample with conventional architecture. The wave-function (Ψhh) clearly shows that<br />

the carriers are confined close to the upper Si/<strong>SiGe</strong> interface in the triangularly<br />

shaped potential. At cryogenic temperatures only the lowest sub-b<strong>and</strong> <strong>of</strong> the heavyhole<br />

b<strong>and</strong> (HH) is occupied.<br />

� Source: p-mod structure.jpg<br />

laxed <strong>SiGe</strong>-buffers are nowadays also used in p-type structures with Ge-rich or even pure<br />

Ge-channels. All three structures depicted in Fig. 6.1 consist <strong>of</strong> a high mobility channel,<br />

which is separated from the remote doping layer with a spacer layer to suppress remote im-<br />

purity scattering. Frontside doping is the conventional architecture for modulation-doped<br />

heterostructures having the doping layers at the side <strong>of</strong> the channel-controlling gate. This<br />

structure is usually preferred over the ”inverted” design where the doping layer is positioned<br />

beneath the channel at the substrate-side. There segregation <strong>of</strong> dopants can lead to undesir-<br />

able impurity scattering <strong>and</strong> secondly the adjusting influence <strong>of</strong> an optional gate is weekend<br />

by automatically increasing the distance between doping layer <strong>and</strong> top-gate. However, in<br />

both architectures the contributing carriers are confined in a triangularly shaped potential<br />

at the dopant-facing Si/<strong>SiGe</strong> interface (see also Fig. 6.2). For a well-chosen doping concen-


88 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

tration all dopants are ionized, all the free carriers are restricted to the channel, <strong>and</strong> only the<br />

lowest sub-b<strong>and</strong> is occupied. Therefore, in this ideal case conduction takes only place in the<br />

narrow confining potential <strong>of</strong> the channel (2DEG or 2DHG) <strong>and</strong> there is no parasitic parallel<br />

conduction path with low-mobility inside the doping region, at least not for sufficiently low<br />

temperatures (typically < 4.2 K).<br />

Whereas for n-MODQW structures extremely high mobilities were observed reaching values<br />

beyond 500 000 cm 2 V −1 s −1 , p-MODQW structures seem to be restricted to hole mobilities<br />

around 20 000 cm 2 V −1 s −1 for <strong>SiGe</strong>-channels [190, 194], <strong>and</strong> well below 100 000 cm 2 V −1 s −1<br />

even for pure Ge-channels [195, 196] with a strongly reduced effective hole mass [197] <strong>and</strong><br />

absent alloy scattering. The origin <strong>of</strong> the vast difference in low-temperature mobilities is still<br />

not clear. At least for n-doped strained-silicon heterostructures no intrinsic limitations are<br />

Figure 6.3: Calculated 2DHG hole mobilities for a 200 ˚A Si1−xGex QW <strong>and</strong><br />

a carrier density <strong>of</strong> ps = 2 × 10 11 as a function <strong>of</strong> Ge-content x. The solid lines<br />

demonstrate the estimated limitations for scattering based on deformation potential<br />

µDP , surface roughness µSR, alloy disorder µAD, piezoelectric charges µP E.<br />

The calculated total mobility µtot nicely fits the reported experimental 4 K data<br />

(filled [6, 190] <strong>and</strong> open squares [191]; data <strong>of</strong> flat reference sample 1663LSG p<br />

(see Tab. 6.1 <strong>and</strong> 6.2) is marked with a red circle). The dashed line clearly demonstrates<br />

that alloy disorder alone without screening is not able to reproduce the<br />

observed monotonic decrease in mobility even with an unjustifiable high alloy potential<br />

ual = 0.74 eV (for details see text). [192]<br />

� Source: p-mod scattering hole-mobility.jpg


6.1. INTRODUCTION TO P-MODULATION DOPING 89<br />

found, which would impede mobilities well over 1 000 000 cm 2 V −1 s −1 , as long as uncontrolled<br />

background doping <strong>and</strong> unintentional alloying in the quantum well can be suppressed during<br />

growth [198]. Alloy scattering <strong>and</strong> interface-roughness scattering together with strain fluctua-<br />

tions arising from smeared-out Si/<strong>SiGe</strong> interfaces coming along with Ge-segregation [126, 199]<br />

<strong>and</strong> interface charges are discussed as the ultimate limitations regarding mobility. How-<br />

ever many theoretical treatments are found in literature which usually adjust numerous<br />

parameters to fit the experimentally observed low-mobility results for p-modulation-doped<br />

structures. Thus <strong>of</strong>ten the predictions for the dominant limiting scattering mechanism are<br />

different. [6, 7]<br />

It seems that in the early years high relevance was attributed to alloy scattering regard-<br />

ing low-temperature mobility [200]. Especially by accounting for screening [201], even with<br />

utilizing unjustifiably high values <strong>of</strong> the alloy potential, the experimental data were not repro-<br />

duced theoretically on the basis <strong>of</strong> predominant alloy scattering [202]. Also, unrealistically<br />

high interface impurity levels were introduced into numerical calculations to explain the low-<br />

mobility observed in hole transport [203, 204]. Other investigations favor interface roughness<br />

scattering as limiting factor [191, 205, 206, 207].<br />

A more recent theory claims to solve the deficiencies <strong>of</strong> earlier explanation attempts [192].<br />

This model is not based on the ill-defined concept <strong>of</strong> interface impurity charges, but tries to<br />

adequately include all possible scattering sources, such as alloy disorder, surface (interface)<br />

roughness, deformation potential, <strong>and</strong> piezoelectric charges in a full treatment <strong>of</strong> the hole<br />

mobility. Deformation potential scattering is regarded herein as the limiting mechanism,<br />

which results from the combination <strong>of</strong> lattice-mismatch strain <strong>and</strong> surface roughness. The<br />

latter gives r<strong>and</strong>om, nonuniform valence b<strong>and</strong> shifts which are experienced by the holes <strong>and</strong><br />

subjected to non-vanishing <strong>of</strong>f-diagonal components <strong>of</strong> the strain field in the <strong>SiGe</strong> layer [192].<br />

These theoretical predictions are visualized in Fig. 6.3, where the hole mobilities limited by<br />

individual scattering mechanisms <strong>and</strong> the calculated total mobility versus Ge-content are<br />

plotted. The overall mobility µtot nicely fits the reported experimental 4 K data [6, 190, 191]<br />

for the plotted range <strong>of</strong> raising Ge-content x. Alloy disorder alone without screening is ob-<br />

viously not able to reproduce the observed monotonic decrease in mobility even with an<br />

unjustifiable high alloy potential ual = 0.74 eV (i.e. the VB <strong>of</strong>fset between Si <strong>and</strong> Ge used as<br />

a natural choice), since there appears a calculated minimum in 2DHG alloy disorder mobility<br />

µAD around x ∼ 0.4, which is inconsistent with experimental findings. However, this is not<br />

such a hard pro<strong>of</strong>, since the calculated minimum is very shallow <strong>and</strong> it is hard to grow 2D<br />

layers at higher Ge-content x. According to the model at h<strong>and</strong> a surface roughness domi-<br />

nated mobility µSR is expected only for a very low Ge-content x � 0.05, whereas just for high


90 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Ge-contents exceeding x � 0.5 piezoelectric charges get important. The commonly experi-<br />

mentally utilized Ge-content region (0.1 � x � 0.4) seems to be limited in low-temperature<br />

mobility predominantly by deformation potential scattering (see Fig. 6.3). For comparison<br />

the mobility data <strong>of</strong> the flat reference sample 1663LSG p (see Tab. 6.1 <strong>and</strong> 6.2 (page 102))<br />

was added to Fig. 6.3 <strong>and</strong> is marked with a red circle. [192].<br />

6.2 Experimental Aspects <strong>and</strong> Mobility Analysis<br />

The first p-modulation-doped structures grown within this work were fabricated by MBE on<br />

st<strong>and</strong>ard 4” Si-wafers to check the 2DHG transport parameters for flat substrates <strong>and</strong> to<br />

compare the achieved results to values reported in literature. These procedure also delivers<br />

plenty <strong>of</strong> material for testing <strong>and</strong> optimization <strong>of</strong> the Hall-bar samples. All the structures<br />

used in the following are based on the conventional p-MODQW architecture with the doping<br />

layer at the frontside within the Si-cap layers. The relevant valence b<strong>and</strong> <strong>and</strong> layer structure<br />

<strong>of</strong> the employed p-MODQW architecture is sketched in Fig. 6.2. The conductive 2DHG is<br />

confined in the triangular potential within the <strong>SiGe</strong>-channel at the upper Si/<strong>SiGe</strong> interface.<br />

The growth temperature for the Si0.75Ge0.25 channel <strong>and</strong> the subsequently deposited Si-spacer<br />

Sample 1663LSG p 1682LSG p<br />

Si Buffer 700 ˚A + 100 ˚A + 200 ˚A Si 240 ˚A + 1000 ˚A + 75 ˚A Si<br />

Si @ 0.7 ˚A/s → 0.2 ˚A/s Si @ 0.2 ˚A/s<br />

550 ◦ C → 450 ◦ C 750 ◦ C ↘ , 425 ◦ C, ↘ 350 ◦ C<br />

<strong>SiGe</strong> Channel 100 ˚A Si0.75Ge0.25 100 ˚A Si0.75Ge0.25<br />

Si @ 0.2 ˚A/s, Ge @ 0.0667 ˚A/s Si @ 0.2 ˚A/s, Ge @ 0.0667 ˚A/s<br />

450 ◦ C 350 ◦ C<br />

Si Spacer 100 ˚A Si 50 ˚A + 50 ˚A Si<br />

Si @ 0.2 ˚A/s Si @ 0.2 ˚A/s<br />

450 ◦ C 350 ◦ C → 425 ◦ C, 425 ◦ C<br />

p-doping 200 ˚A Si:B (p2.5e18 #/cm 3 ) 200 ˚A Si:B (p2.5e18 #/cm 3 )<br />

Layer Si @ 0.2 ˚A/s, B @ 1717.9 ◦ C Si @ 0.2 ˚A/s, B @ 1717.9 ◦ C<br />

450 ◦ C 425 ◦ C<br />

Si Cap 100 ˚A + 600 ˚A Si 100 ˚A + 100 ˚A + 500 ˚A Si<br />

Si @ 0.2 ˚A/s → 0.7 ˚A/s, 0.7 ˚A/s Si @ 0.2 ˚A/s → 0.7 ˚A/s<br />

450 ◦ C → 550 ◦ C, 550 ◦ C 425 ◦ C → 500 ◦ C → 550 ◦ C, 550 ◦ C<br />

Table 6.1: Layer sequence <strong>of</strong> conventional p-modulation doped structures.


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 91<br />

layer is chosen as compromise between good crystal quality <strong>and</strong> a sharp Si/<strong>SiGe</strong> interface.<br />

The growth parameters for such a structure grown on a flat Si-substrate (1663LSG p) are<br />

listed in Tab. 6.1. For the actual samples with predefined modulations in the <strong>SiGe</strong>-channel<br />

the growth temperatures had to be lowered even further (Tab. 6.1, 1682LSG p). This be-<br />

came necessary to suppress unwanted, strain-driven corrugations at the Si/<strong>SiGe</strong> interface<br />

(see Fig. 5.2). Otherwise, additional fringes perpendicular to the periodic ripples pattern<br />

could interfere with the desired surface roughness which should involve just step-bunching.<br />

According to the usual practice (Sec. 4.2), the samples on small pieces <strong>of</strong> miscut substrates<br />

were grown simultaneously with a flat reference sample (both: 17.5 mm × 17.5 mm) mounted<br />

in an all-Si substrate adapter for direct comparison. The low growth temperature should<br />

not be that detrimental for the miscut sample because the high step density still enables<br />

step-flow growth. In contrast for the flat substrates the crystal quality is expected to be<br />

strongly reduced. Nevertheless, for a first set <strong>of</strong> samples the used, but not yet optimized pa-<br />

rameters regarding layer thickness, growth temperature, doping level <strong>and</strong> <strong>SiGe</strong>-composition<br />

are envisaged to show at least the principle functionality <strong>of</strong> the approach to implement a<br />

step-bunching-induced modulation in the <strong>SiGe</strong>-channel for a mobility analysis.<br />

6.2.1 Processing <strong>of</strong> Hall-Bars<br />

The usual procedure starts with the preparation <strong>of</strong> the substrates that are intended for Hall-<br />

bar processing. The first step is cutting the samples to an adequate size. From full-size<br />

grown wafers ∼ 2 cm × 2 cm pieces are cleaved from the central region by using a diamond<br />

scribing table, whereas small substrates can be used directly. After protecting the sample<br />

surface with a thin photoresist layer the respective piece is glued with a special wax to a<br />

glass microscope slide <strong>and</strong> the chuck <strong>of</strong> the diamond wire saw [208]. This precision saw is<br />

used to cut the sample to a useful size <strong>of</strong> about 5 mm × 4 mm. A rectangular size is especially<br />

useful for miscut samples to document the orientation <strong>of</strong> the ripple pattern. After cutting<br />

<strong>and</strong> releasing the samples from the wax on a hot plate (∼ 120 ◦ C) the residues <strong>of</strong> wax <strong>and</strong><br />

photoresist are stripped in acetone (ultrasonic bath, 5 min). Further cleaning continues with<br />

the usual pre-cleaning sequence <strong>and</strong> subsequent ultrasonic cleaning steps in trichlorethylene,<br />

acetone, <strong>and</strong> methanol for 3–5 min each. A final Piranha-etch (H2SO4 : H2O2 = 5 : 1, 15–<br />

20 min) <strong>and</strong> DI-H2O rinse concludes the required cleaning procedure, <strong>and</strong> the samples are<br />

blown dry (N2).<br />

In all subsequent masking steps Shipley S1818 photoresist was used as follows. The spinning<br />

parameters were ∼ 3 sec at a rotation speed <strong>of</strong> 2000 rpm <strong>and</strong> ∼ 40 sec at 4000 rpm to realize a<br />

homogeneous photoresist layer. A s<strong>of</strong>tbake at 90 ◦ C in an oven was applied to outgas exces-


92 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.4: (a) Photograph <strong>of</strong> processed Hall-bar on structure 1682LSG p (miscut<br />

substrate) mounted <strong>and</strong> bonded onto a ”Stycast” sample carrier. (b) The zoomin<br />

shows the two perpendicular branches <strong>of</strong> the Hall-bar structure together with<br />

the pin numbers assigned for the measurements. (c) Schematics <strong>of</strong> metal contact<br />

pads (red color) <strong>and</strong> etched Hall-bars (blue color). The line pattern indicates the<br />

elongation direction <strong>of</strong> step-bunching. Thus the upper branch <strong>of</strong> the Hall-bar is<br />

used to measure the conductivity σ || parallel to the bunches <strong>and</strong> the lower branch<br />

σ⊥ perpendicular to the ripple structure.<br />

� Source: p-mod Hall-bar 1682LSGp miscut01.jpg<br />

sive solvents. A chromium-quartz mask with a Hall-bar lay-out featuring two perpendicular<br />

branches for anisotropy measurements was used. The mask was designed by T. Berer during<br />

his diploma thesis for AlGaAs-samples (see [209], p. 104, Fig. 8.13e-f). Thus a self-aligned<br />

top-gate was not implemented <strong>and</strong> is therefore not available. The photoresist was exposed<br />

in a Süss MJB3 mask-aligner for 8 sec <strong>and</strong> subsequently developed (MF319, 60 sec). Rinsing<br />

the sample for ∼ 2 min in DI-H2O stops the developing procedure.<br />

In the first masking step the contact pads are defined <strong>and</strong> the photoresist covers the inverse<br />

areas. A 20 sec short ashing-step in an O2-plasma (200 W, 1 torr) is applied to remove rem-<br />

nants <strong>of</strong> the photoresist or solvents. Immediately before the samples are transferred into the<br />

vacuum environment <strong>of</strong> the evaporation chamber an HF-dip (HF : DI-H2O � 1 : 10, 20–30 sec)<br />

is adopted to strip the natural SiO2 from the exposed contact areas. The sample are again<br />

dried with the N2-nozzle. 1000 ˚A aluminum (Al) is evaporated which is the common material<br />

for p-type contacts. In the ultrasonic bath acetone is used to lift-<strong>of</strong>f the photoresist <strong>and</strong> the<br />

metal layer above. Methanol concludes the cleaning sequence (US-bath, 3–5 min), <strong>and</strong> the<br />

samples are again blown dry. The Al-contacts are alloyed in a special annealing oven in an<br />

ambient <strong>of</strong> forming gas (Ar/H2, 0.25 bar). The annealing step at nominally 450 ◦ C for 6 min<br />

enables a Al-Si inter-diffusion providing ohmic contacts. During this heat treatment the sam-<br />

ple is separated from the heater furnace with a plain Si-wafer piece to prevent contamination


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 93<br />

from the back-side.<br />

The same masking procedure is executed a second time to adapt the metal contacts. This<br />

time after the O2-ashing step no HF-dip is applied <strong>and</strong> onto the contact pads 150 ˚A chromium<br />

(Cr) as undercoating <strong>and</strong> 1250 ˚A gold (Au) are evaporated to provide solid contact pads for<br />

wire-bonding.<br />

After lift-<strong>of</strong>f <strong>and</strong> cleaning the photoresist mask for mesa-etching is created in a working pro-<br />

cedure analogous to the steps listed above. The asher [210] is exploited a last time before<br />

the samples are placed in the reactive-ion-etching machine (RIE) [211]. The etching process<br />

defines the ”Hall-bar” structure since outside the mesa-structure the 2DHG-layers are com-<br />

pletely removed. For this purpose the employed etching parameters (2 min @ 5 mtorr; 90 W;<br />

5% O2, 25% SF6) provide an etch depth <strong>of</strong> ∼ 200 nm. Subsequent photoresist stripping yields<br />

a completely processed Hall-bar without top-gate.<br />

The samples are mounted on exchangeable ”Stycast” sample holders with a special adhesive † .<br />

Afterwards, the contact pads are inter-connected with the pins <strong>of</strong> the sample carrier using<br />

a wire-bonding machine with Au-wire. The bonded wire-ends stick nicely to the Au-covered<br />

contact pads but not so well to the pins <strong>of</strong> the self-made ”Stycast” sample holder. Therefore,<br />

the wires are additionally soldered with indium (In) at the pin-side to secure the mechanical<br />

<strong>and</strong> essential electrical connection.<br />

The samples are now prepared <strong>and</strong> ready for the electrical characterization in a 7 T super-<br />

conductivity magnet. In Fig. 6.4a a photograph <strong>of</strong> the finally processed Hall-bar on structure<br />

1682LSG p (miscut substrate) mounted <strong>and</strong> bonded onto a ”Stycast” sample carrier is repre-<br />

sented. The zoom-in (Fig. 6.4b) shows the two perpendicular branches <strong>of</strong> the Hall-bar struc-<br />

ture together with the pin numbers assigned for the cryo-measurements. Fig. 6.4c depicts<br />

schematically the metal contact pads (red color) <strong>and</strong> the Hall-bars defined by the mesa-etch<br />

(blue color).<br />

6.2.2 Cryo-Measurements <strong>and</strong> Data Evaluation<br />

The electrical characterization was performed in a 4 He-immersion cryostat with an adjustable<br />

magnetic field up to B = 7 T. The inner reservoir where the sample is located can be pumped,<br />

which extends the temperatures available for measurements from the st<strong>and</strong>ard liquid 4 He tem-<br />

perature (TLHe � 4.2 K) down to ∼ 1.6 K. A sample heating unit provides the possibility to<br />

perform temperature-dependent measurements up to room temperature (∼ 300 K). Further<br />

details <strong>of</strong> the 4 He-system can be found in Ref. [212].<br />

† also known as ”Pudalov-glue” at the institute


94 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

The basics <strong>of</strong> electronic transport <strong>and</strong> measurements regarding the Shubnikov–de Haas (SdH)<br />

<strong>and</strong> the Integer Quantum Hall (IQH) effect can be found in st<strong>and</strong>ard semiconductor physics<br />

text books [213, 214, 215]. For further reading on data evaluation from electrical <strong>and</strong> magneto-<br />

transport experiments the reader is advised to have a look at Ref. [216].<br />

All transport data presented in this section were gathered by a dc-measurement technique.<br />

A dc-current <strong>of</strong> I = 1 µA was established along the branches <strong>of</strong> the Hall-bar structure. Al-<br />

though the current was usually driven between pin 5 <strong>and</strong> pin 12 (see Fig. 6.4c) a homoge-<br />

neously current flow along each branch was realized, which is fully equivalent to currents<br />

directed separately between contacts 5 <strong>and</strong> 9 or 9 <strong>and</strong> 12, respectively. The realization <strong>of</strong><br />

the conventional 4-point measurement setup decouples the detected longitudinal resistivity<br />

RL from the quality <strong>of</strong> the contacts. Thus the normalized longitudinal resistivity ρxx can<br />

be precisely calculated from the longitudinal voltage drop between the inner contact pads,<br />

e.g. between 6 <strong>and</strong> 8 or 10 <strong>and</strong> 11, which is measured with a high impedance voltage meter.<br />

The geometry <strong>of</strong> the Hall-bar, in fact the ratio between the width W <strong>and</strong> the length L, has<br />

to be considered to extract the correct value for ρxx according to Eq. 6.1. The length L is<br />

the distance between the contacts applied for measuring the longitudinal voltage drop VL.<br />

Therefore the geometry factor is G = 1 for neighboring contacts such as 6 <strong>and</strong> 7, <strong>and</strong> G = 0.5<br />

for the preferred longitudinal voltage measurements such as between 6 <strong>and</strong> 8.<br />

G = W<br />

L , RL = VL<br />

I<br />

ρxx = RL · G<br />

(6.1)<br />

The Hall-voltage VH picked up at contacts located across the Hall-bar, e.g. between pins 3<br />

<strong>and</strong> 6 or 1 <strong>and</strong> 11, is a direct measure for the Hall-resistivity ρxy after Eq. 6.2.<br />

ρxy = VH<br />

I<br />

(6.2)<br />

At sufficiently low temperatures T , where thermodynamical broadening does not conceal the<br />

energy splitting <strong>of</strong> the L<strong>and</strong>au levels, <strong>and</strong> at a reasonably high magnetic fields B the two-<br />

dimensionality <strong>of</strong> the hole gas (2DHG) gets obvious from the Shubnikov–de Haas oscillations<br />

in ρxx <strong>and</strong> the Integer Quantum Hall effect in ρxy. For all investigations the samples were<br />

orientated in the cryostat in such a way that the magnetic field was aligned in perpendicular<br />

direction to the 2DHG. Thus, the relevant perpendicular magnetic field component B⊥ is<br />

equal to the absolute value <strong>of</strong> the applied magnetic field B (B⊥ ≡ B).<br />

Fig. 6.5 shows the measured data gathered from a magnetic field sweep for the extracted<br />

resistivity along (ρxx) <strong>and</strong> across (ρxy) the upper branch <strong>of</strong> the Hall-bar <strong>of</strong> sample 1663LSG p,<br />

which was grown on a flat st<strong>and</strong>ard Si-substrate. Although not shown here, the transport<br />

investigation for the lower branch produced comparable curves since there is no intrinsic


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 95<br />

anisotropy in the structure. The Hall-bars are oriented along the two perpendicular 〈110〉-<br />

directions. The B-sweep at ∼ 1.7 K yields pronounced SdH-oscillations in ρxx <strong>and</strong> extended<br />

plateaus featuring the Integer Quantum Hall effect in ρxy. The increase <strong>of</strong> the resistivity<br />

ρxx around B = 0 T can be attributed to a phenomenon called weak localization effect. This<br />

is a quantum-mechanical interference effect arising from coherent backscattering <strong>of</strong> single<br />

electrons, which is associated with an additional contribution to resistivity (see also [212,<br />

216]).<br />

Fig. 6.6 represents the same data depicted in the previous figure (Fig. 6.5) but plotted as a<br />

function <strong>of</strong> the inverse magnetic field 1/B at the abscissa, <strong>and</strong> with the Hall-resistivity ρxy<br />

normalized to the universal resistance 25 812.8 Ω (h/e 2 ) [217]. The quantized Hall resistance<br />

gives integer filling factors ν <strong>of</strong> the L<strong>and</strong>au levels according to Eq. 6.3.<br />

ρxy = 1 h<br />

ν e2 (6.3)<br />

Fig. 6.6b demonstrates the filling factor ν <strong>of</strong> the L<strong>and</strong>au levels with extended plateaus for<br />

even values due to spin degeneracy gs = 2 <strong>and</strong> valley degeneracy gv = 1 in the valence b<strong>and</strong>.<br />

For magnetic fields exceeding ∼ 5 T the spin degeneracy is already slightly lifted, which gets<br />

obvious from the faintly appearing plateau at ν = 5.<br />

The transport parameters <strong>of</strong> the holes confined in the <strong>SiGe</strong>-channel were predominantly<br />

calculated from longitudinal resistivity ρxx <strong>and</strong> especially via the SdH-oscillations. These<br />

deliver information which is based purely on the electronic properties <strong>of</strong> the 2DHG <strong>and</strong><br />

do not account for the carriers in a potential parasitic channel in the doping region. The<br />

hole sheet carrier density ps in the 2DHG, therefore <strong>of</strong>ten also referred to as p2D, is either<br />

calculated from the periodicity <strong>of</strong> the SdH-oscillations as a function <strong>of</strong> inverse magnetic field<br />

∆(1/B) (Eq. 6.4) or directly from the position <strong>of</strong> an SdH-minimum with regard <strong>of</strong> magnetic<br />

field B, whenever the filling factor ν is known (Eq. 6.5).<br />

ps = e<br />

h gsgv<br />

�<br />

∆ 1<br />

�−1 B<br />

(6.4)<br />

ps = ν e<br />

B (6.5)<br />

h<br />

The mobility µ can be recalculated from the longitudinal conductivity σxx or the resistivity<br />

ρxx for zero magnetic field (B = 0 T). Since the mobility µ depends on the carrier density<br />

(Eq. 6.6) an accurate determination <strong>of</strong> the hole density ps is necessary for a reliable assessment<br />

<strong>of</strong> the mobility µ.<br />

σxx = (ρxx) −1 = pseµ → µ = 1 −1<br />

(ρxx)<br />

pse<br />

(6.6)


96 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.5: Measured data <strong>of</strong> (a) Shubnikov–de Haas (SdH) <strong>and</strong> (b) Integer<br />

Quantum Hall (IQH) effect at ∼ 1.7 K in a perpendicular magnetic field for a flat<br />

p-modulation doped <strong>SiGe</strong>-structure (1663LSG p).<br />

� Source: p-mod Hall-bar 1663LSGp SdH QH B.jpg<br />

In this work the measured data point ρxx(B=0T) was not directly extracted from the B-<br />

sweep. Instead, the additional resistivity contribution <strong>of</strong> the weak localization effect was<br />

neglected by extrapolating the resistivity curve from finite magnetic fields towards B = 0 T.<br />

This leads generally to slightly reduced resistivity values <strong>and</strong> finally results in enhanced mo-<br />

bility readings. The extracted transport properties <strong>of</strong> sample 1663LSG p characterized at<br />

4.2 K <strong>and</strong> 1.66 K are summarized in Tab. 6.2 (see page 102). The recalculated values for ps


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 97<br />

Figure 6.6: Representation <strong>of</strong> the same experimental data from Fig. 6.5 plotted<br />

as a function <strong>of</strong> the inverse magnetic field. The equidistant SdH-oscillations in this<br />

plot (a) are used to derive the carrier concentration ps. The lower graph (b) shows<br />

the filling factor ν <strong>of</strong> the L<strong>and</strong>au levels with extended plateaus for even values (spin<br />

degeneracy gs = 2, valley degeneracy gv = 1).<br />

� Source: p-mod Hall-bar 1663LSGp SdH nu invB.jpg<br />

<strong>and</strong> µ have to be considered with an uncertainty or error bar <strong>of</strong> up to 10–15%. The achieved<br />

mobility is comparable to literature values for a Si0.75Ge0.25 channel although clearly below<br />

the reported record mobilities [6, 190, 191] (see again Fig. 6.3).<br />

After the pre-experiments with flat ”high”-mobility p-modulation doped material the ac-


98 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.7: Schematic drawings <strong>of</strong> p-<strong>SiGe</strong> conventional modulation-doped structure<br />

with frontside doping where the conduction takes place at the upper interface<br />

<strong>of</strong> the <strong>SiGe</strong>-channel. The modulation <strong>of</strong> the channel yields expected differences in<br />

conductivity for measurements parallel (σ ||) <strong>and</strong> perpendicular (σ⊥) to the ripple<br />

structure.<br />

� Source: p-mod conductivity model.jpg<br />

tual measurements were addressed. Here the p-MODQW structures grown on step-bunching<br />

templates were envisaged for measuring a mobility anisotropy. Again, the branches <strong>of</strong> the<br />

Hall-bar structure were processed with orientation along the 〈110〉-direction. This way the<br />

buried ripple structure – giving a modulation at the upper Si/<strong>SiGe</strong>-interface, <strong>and</strong> thus <strong>of</strong> the<br />

2DHG – is aligned parallel to the upper branch <strong>and</strong> perpendicular to the lower branch <strong>of</strong> the<br />

Hall-bar (see Fig. 6.4c). The modulation <strong>of</strong> the channel yields expected differences in con-<br />

ductivity for the measurements parallel (σ ||) <strong>and</strong> perpendicular (σ⊥) to the ripple structure.<br />

Fig. 6.7 depicts schematic drawings <strong>of</strong> p-<strong>SiGe</strong> modulation-doped structure with frontside dop-<br />

ing where the conduction takes place at the upper interface <strong>of</strong> the <strong>SiGe</strong>-channel. The electrical<br />

properties were investigated in detail for two sets <strong>of</strong> samples (1882LSG p, 1883LSG p), each<br />

a sample on miscut substrate <strong>and</strong> on a flat reference substrate. The growth sequence for<br />

p-MODQW structure 1883LSG p is identical to 1882LSG p except for the lowered doping<br />

concentration (p1.0e18 #/cm 3 ) in the Si:B doping layer.<br />

Fig. 6.8 shows selected B-sweep measurement curves <strong>of</strong> the longitudinal resistivity ρxx<br />

(Fig. 6.8a) <strong>and</strong> the transversal resistivity ρxy (Fig. 6.8b) for the Hall-bars parallel (solid<br />

curves) <strong>and</strong> perpendicular (dashed curves) to the periodic modulations in the <strong>SiGe</strong>-channel<br />

due to step-bunching. A LED mounted near the sample is utilized to illuminate the sample.<br />

The light generates electron-hole pairs which modifies the carrier concentration depending on<br />

the illumination level. This is a poor, but successful, substitute to tuning the carrier density


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 99<br />

Figure 6.8: Plots <strong>of</strong> the longitudinal resistivity ρxx (a) <strong>and</strong> transversal resistivity<br />

ρxy (b) for the Hall-bars parallel (solid curves) <strong>and</strong> perpendicular (dashed curves)<br />

to the periodic modulations in the <strong>SiGe</strong>-channel due to step-bunching. The data<br />

for miscut sample 1682LSG p recorded at 4.2 K <strong>and</strong> ∼ 1.6 K are plotted for different<br />

stages <strong>of</strong> sample illumination.<br />

� Source: p-mod Hall-bar 1682LSGp miscut01 SdH QH B.jpg<br />

with a top-gate. Data recorded at 4.2 K <strong>and</strong> ∼ 1.6 K at different stages <strong>of</strong> sample illumination<br />

are plotted for miscut sample 1682LSG p (see Fig. 6.8). Tab. 6.2 gives a complete overview on<br />

the measurement parameters <strong>and</strong> the extracted values for carrier density <strong>and</strong> mobility <strong>of</strong> the<br />

investigated samples. Again it has to be remarked that the confidence <strong>of</strong> the data evaluation


100 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.9: Evaluated data from Fig. 6.8 summarized in a plot showing mobility<br />

µ versus carrier concentration ps for miscut sample 1682LSG p. The data clearly<br />

prove a lower carrier mobility perpendicular to the ripples due to increased scattering.<br />

� Source: p-mod Hall-bar 1682LSGp miscut01 mu p.jpg<br />

is estimated to be within a 10–15% uncertainty. From Fig. 6.8a it can clearly be seen that the<br />

resistivity ρxx for the investigation along (||) <strong>and</strong> across (⊥) the ripple modulation forms two<br />

groups <strong>of</strong> curves in the resistivity plot. Obviously, the conductivity µ || is by nearly a factor <strong>of</strong><br />

two higher in the upper branch <strong>of</strong> the Hall-bar structure although the carrier densities ps are<br />

comparable (see Fig. 6.8b). This indicates a strong influence <strong>of</strong> the step-bunching structure<br />

on the 2DHG mobility. The evaluated data <strong>of</strong> the miscut sample 1682LSG p are summarized<br />

in a plot showing the mobility µ versus carrier concentration ps (Fig. 6.9). The data clearly<br />

prove a lower carrier mobility perpendicular to the ripples due to increased scattering.<br />

The change in temperature from 4.2 K to ∼ 1.6 K does not affect the carrier concentration or<br />

mobility remarkably. Nevertheless, the lowered temperature clearly enhances the oscillation<br />

amplitude <strong>of</strong> the SdH-effect <strong>and</strong> produces well-pronounced IQH-plateaus (Fig. 6.8). As pre-<br />

dicted for a considerable illumination level, the carrier concentration in the <strong>SiGe</strong>-channel is<br />

increased which directly lowers the resistivity. Additionally, the raised level <strong>of</strong> free carriers<br />

favors higher mobilities due to the effect <strong>of</strong> screening where perturbing scattering potentials<br />

are partially shielded (Fig. 6.9). This second effect further lowers the classical longitudinal<br />

resistivity ρxx on which the quantum-mechanical Shubnikov–de Haas oscillations are over-


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 101<br />

layed. In ρxy the overall slope is decreased along with enhanced carrier density since this<br />

slope is a direct measure in conventional Hall characterization to evaluate the carrier density<br />

ps. For all the involved illumination levels the detected ps is nearly identical for both Hall-bar<br />

branches, which is proved by the positions <strong>of</strong> the SdH-minima in the ρxx-plot (Fig. 6.8a) <strong>and</strong><br />

the almost congruent ρxy-curves (Fig. 6.8b).<br />

6.2.3 Data Interpretation<br />

In all investigated samples the doping level was rather high so that already without illu-<br />

mination the <strong>SiGe</strong>-channel was almost fully populated with carriers (ps ∼ 5 ×10 11 cm −2 , see<br />

Tab. 6.2). Therefore the lower doped samples 1883LSG p showed carrier densities compara-<br />

ble to 1882LSG p. Long-lasting permanent effects arising from previous sample illumination<br />

were not observed. Usually, some minutes decay time provided, the recorded data before <strong>and</strong><br />

after illumination were comparable.<br />

Some inconsistencies with the results are found for the flat reference sample 1882LSG p<br />

(referred as 1882 (d) as abbreviation for 1882LSG p dummy) itself <strong>and</strong> furthermore by com-<br />

parison with the lower doped sample 1883LSG p. Although no intrinsic anisotropy should be<br />

involved for the two branches <strong>of</strong> the flat 2DHG-sample 1882LSG p (dummy), the character-<br />

ization showed an extremely low mobility value for the upper branch (see Tab. 6.2: 1682 (d)<br />

before illumination <strong>and</strong> at 4.2 K). Since a contact problem was suspected, the sample was<br />

unmounted, the contacts were checked <strong>and</strong> the sample was reinstalled into the measurement<br />

set-up, rotated by 90 ◦ . After the involved warm-up <strong>and</strong> cool-down procedure the sample<br />

behaved differently (1682 (d)*) <strong>and</strong> more reasonably. The unexpected anisotropy observed<br />

before was gone (or even slightly inverted). This indicates fluctuations in the sample which<br />

might be attributed to structure induced defects. It was not further investigated whether<br />

this anomalousness arises from growth defects or from post-processing.<br />

More interestingly, the observed mobilities for the different samples do not fit into a simple<br />

systematics. Whereas the lower doped miscut sample 1883LSG p (m) shows unpredictable<br />

results except for high illumination, the corresponding dummy sample (1883LSG p (d)) per-<br />

forms well already without illumination. This flat reference sample even outperforms the<br />

nominally higher doped dummy 1882LSG p (d), except for high illumination where the latter<br />

sample seems to recover performance. A decrease in carrier density with increasing illumi-<br />

nation is found for sample 1882LSG p (d) <strong>and</strong> 1883LSG p (m) (see Tab. 6.2) which is an un-<br />

expected result. Generally the samples 1882LSG p (d), 1883LSG p (m), <strong>and</strong> 1883LSG p (d)<br />

would benefit definitely from top gates, which should give more reproducible <strong>and</strong> reliable<br />

measurement results since surface potentials are then well defined.


102 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Sample<br />

upper branch (Pin 5-9) lower branch (Pin 9-12)<br />

mobility µ || hole conc. ps mobility µ⊥ hole conc. ps<br />

[cm 2 V −1 s −1 ] [×10 11 cm −2 ] [cm 2 V −1 s −1 ] [×10 11 cm −2 ]<br />

illum. temp.<br />

1663 2690 5.65 2670 5.65 – 4.2 K<br />

1663 2880 5.60 2830 5.60 – 1.66 K<br />

1682 (m) 1350 5.00 745 5.00 before 4.2 K<br />

1682 (m) 1330 5.00 685 5.00 before 1.64 K<br />

1682 (m) 1190 4.80 630 4.80 after 1.64 K<br />

1682 (m) 1360 5.20 695 5.30 5 nA 1.64 K<br />

1682 (m) 1450 5.50 845 5.70 50 nA 1.61 K<br />

1682 (m) 1500 5.80 875 5.85 250 nA 1.61 K<br />

1682 (m) 1280 4.85 690 4.85 after 1.59 K<br />

1683 (m) 340 5.00 290 6.10 before 1.64 K<br />

1683 (m) 390 5.30 540 4.20 after 1.67 K<br />

1683 (m) 1230 4.40 695 4.40 50 nA 1.68 K<br />

1683 (m) 1300 4.70 760 4.60 250 nA 1.69 K<br />

1682 (d) 590 4.95 900 4.95 before 4.2 K<br />

1682 (d) 2400 4.55 2400 4.55 250 nA 4.2 K<br />

1682 (d)* 950 4.60 860 4.70 – 4.2 K<br />

1682 (d)* 940 4.60 850 4.70 – 1.65 K<br />

1683 (d) 1300 5.40 1300 5.30 – 4.2 K<br />

1683 (d) 1300 5.40 1300 5.30 – 1.65 K<br />

Table 6.2: 2DHG transport parameters evaluated from SdH-measurements. The<br />

mobilities parallel (µ ||) <strong>and</strong> perpendicular (µ⊥) to step-bunching are listed for the<br />

upper <strong>and</strong> lower branch <strong>of</strong> the different Hall-bar samples, respectively. The different<br />

samples are specified in abbreviated notation: (m) denotes miscut <strong>and</strong> (d) flat<br />

reference (”dummy”) samples. Sample 1663 was grown on a flat st<strong>and</strong>ard 4”-Si-<br />

wafer. The ”*” indicates a second measurement after sample reinstallation (rotated<br />

by 90 ◦ ), <strong>and</strong> associated warm-up/cool-down procedure for sample 1682 (d).


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 103<br />

The highest mobility was found for the flat sample 1663LSG p. This is not surprising<br />

as in the growth procedure <strong>of</strong> this sample a more appropriate growth temperature could be<br />

applied (see Tab. 6.1). Especially for the flat reference samples 1682LSG p <strong>and</strong> 1683LSG p<br />

the crystal structure suffers from the reduced substrate temperature, which became necessary<br />

to prevent strain-driven features at the Si/<strong>SiGe</strong>-interface perpendicular to step-bunching (see<br />

also preamble <strong>of</strong> Sec. 6.2). For another reason it is not possible to directly compare the mis-<br />

cut <strong>and</strong> reference samples. Although these were grown simultaneously they were grown at a<br />

slightly different location in the MBE-system with regard to the sources. The samples were<br />

oriented in the growth chamber with the miscut substrate at the Ge-rich side <strong>and</strong> the reference<br />

dummy substrate at the Si-rich side. Therefore, some differences in the <strong>SiGe</strong>-composition –<br />

layer thickness <strong>and</strong> Ge-content – <strong>and</strong> doping concentration (see Ch. A) are always present.<br />

These are complications which impede a direct comparison <strong>of</strong> several measurement results.<br />

According to the evaluation <strong>of</strong> calibration data (see Ch. A, especially Tab. A.3) deviations<br />

from the nominal parameters <strong>of</strong> the <strong>SiGe</strong>-channel were estimated <strong>and</strong> are listed in Tab. 6.3.<br />

The differences in doping level should play only a subordinate role since in all cases the chan-<br />

nel seems to be highly filled with hole-type carriers. The <strong>SiGe</strong>-channel thickness can also<br />

be neglected because the carriers are confined at the upper Si/<strong>SiGe</strong> interface in a triangular<br />

potential anyway. Presumably the most significant restriction is in this sense probably the<br />

Ge-content, since its value deviates from the nominal value, which is calibrated for the center<br />

position in the growth chamber, by x = 0.25 ± 0.05 for the Ge-rich <strong>and</strong> Si-rich side, respec-<br />

tively (Tab. 6.3). Although the dependence <strong>of</strong> Ge-content on the mobility is already weak<br />

around x = 0.25 (see Fig. 6.3), the Ge-content is next to the growth temperature the major<br />

complication for a direct data comparison.<br />

Therefore, the following considerations are mainly related to miscut sample 1682LSG p.<br />

Although there was the justified expectation to measure an anisotropy in mobility for a<br />

Hall-bar oriented parallel to the ripple modulation (upper branch) <strong>and</strong> perpendicular (lower<br />

branch), such a remarkable effect with a factor <strong>of</strong> two difference in µ seemed to be well out<br />

<strong>of</strong> scope. To rule out severe constrictions in the <strong>SiGe</strong>-layer, cross-sectional transmission elec-<br />

tron microscopy (XTEM) investigations were conducted on miscut sample 1882LSG p (m).<br />

growth position center (calibr.) Si-rich side Ge-rich side<br />

<strong>SiGe</strong>-channel thickness Lz [˚A] 100.0 102.4 93.7<br />

Ge-content x 0.250 0.225 0.275<br />

B doping conc. p [#/cm 3 ] 2.5e18 2.9e18 1.8e18<br />

Table 6.3: Deviations <strong>of</strong> the <strong>SiGe</strong>-channel parameters for different positions.


104 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.10: XTEM images stiched together depicting the dark Si0.75Ge0.25<br />

channel <strong>of</strong> miscut sample 1682LSG p, embedded between Si-buffer <strong>and</strong> Si-spacer<br />

layer. The view along the ripple direction [110] reveals a seemingly homogeneous<br />

Si0.75Ge0.25 layer thickness. The Si/<strong>SiGe</strong> interface is only partially smeared out due<br />

to slight modulations along the bunches lying within the finite sample thickness in<br />

view direction.<br />

� Source: p-mod XTEM 1682LSGp Imag10-14.jpg<br />

Fig. 6.10 presents a set <strong>of</strong> XTEM images stiched together depicting the dark Si0.75Ge0.25<br />

channel embedded between Si-buffer <strong>and</strong> Si-spacer layer. The view along the ripple direction<br />

[110] reveals a seemingly homogeneous Si0.75Ge0.25 layer thickness. The <strong>SiGe</strong> channel repli-<br />

cates the underlying step-bunched Si-buffer in a conformal manner. The Si/<strong>SiGe</strong> interface<br />

is partially smeared out due to slight modulations along the bunches lying within the finite<br />

sample thickness in view direction. The contrast in the presented XTEM data was strongly<br />

enhanced <strong>and</strong> optimized to reveal the morphological details <strong>of</strong> the s<strong>and</strong>wiched <strong>SiGe</strong>-layer.<br />

Therefore the epoxy glue, which would appear on top is not visible within the shown gray<br />

scale. The top part <strong>of</strong> the Si-cap layer is clearly amorphous which is an artifact generated by<br />

ion-sputtering <strong>and</strong> sample thinning during the TEM-preparation procedure. Fig. 6.11a shows<br />

another more detailed XTEM image <strong>of</strong> the Si0.75Ge0.25 channel <strong>of</strong> miscut sample 1682LSG p.<br />

The red box serves as guide to the eye to help resolving the periodic modulations (Λ ∼ 100 nm)<br />

<strong>of</strong> the Si0.75Ge0.25 channel. The modulation amplitude <strong>of</strong> the <strong>SiGe</strong> quantum well is clearly<br />

on the nanometer scale. For comparison the <strong>SiGe</strong>-channel thickness reads ∼ 10 nm. As far


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 105<br />

Figure 6.11: (a) More detailed XTEM image <strong>of</strong> the Si0.75Ge0.25 channel <strong>of</strong> miscut<br />

sample 1682LSG p. The red box serves as guide to the eye to help resolving the<br />

periodic modulations (Λ ∼ 100 nm) <strong>of</strong> the Si0.75Ge0.25 channel grown on top <strong>of</strong> the<br />

step-bunching template. (b) Representation <strong>of</strong> the same image squeezed together<br />

in lateral direction to emphasize the modulation <strong>of</strong> step-bunching for the <strong>SiGe</strong>channel.<br />

The red arrows indicate the minima <strong>of</strong> the undulations with ∼ 100 nm<br />

periodicity.<br />

� Source: p-mod XTEM 1682LSGp Imag09.jpg<br />

as can be seen in these XTEM images both Si/<strong>SiGe</strong> interfaces are rather sharp (except for<br />

aforementioned sporadic undulations <strong>of</strong> the <strong>SiGe</strong>-channel in view direction) <strong>and</strong> do not in-<br />

dicate significant intermixing or Ge-segregation. As expected, the doping layer cannot be<br />

identified from these views. The interface between substrate <strong>and</strong> epitaxial layers is located<br />

∼ 130 nm below the <strong>SiGe</strong>-channel <strong>and</strong> is not decorated with defects. This proves a proper<br />

crystal growth based on a successfully performed precleaning procedure. In Fig. 6.11b the<br />

XTEM image <strong>of</strong> Fig. 6.11a is represented squeezed together in lateral direction to emphasize<br />

the modulation <strong>of</strong> step-bunching for the <strong>SiGe</strong>-channel. The red arrows indicate the minima<br />

<strong>of</strong> the undulations with ∼ 100 nm periodicity.<br />

It seems to be straightforward that the strongly reduced mobility is due to roughness scatter-<br />

ing events as the holes experience the modulation <strong>of</strong> the Si/<strong>SiGe</strong> interface when drifting across<br />

the step-bunches. To confirm this it is necessary to compare the length scales <strong>of</strong> typical trans-<br />

spin degeneracy gs<br />

valley degeneracy gv<br />

effective mass m ∗ 0.28 m0 [kg]<br />

carrier density ps 5.0×10 11 [cm −2 ]<br />

mobility (perp.) µ 700 [cm 2 V −1 s −1 ]<br />

Table 6.4: Parameters employed for the estimation <strong>of</strong> scattering.<br />

2<br />

1


106 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

port related parameters <strong>and</strong> the period <strong>of</strong> the undulations <strong>of</strong> step-bunching. The thickness<br />

fluctuations <strong>of</strong> the <strong>SiGe</strong> quantum well are characterized in terms <strong>of</strong> correlation length Λ <strong>and</strong><br />

amplitude ∆ <strong>of</strong> the surface roughness. Usually only short-range surface roughness is treated<br />

(kF · Λ ≪ 1) <strong>and</strong> used as fitting parameter to experimental mobility data [192, 218, 219]. In<br />

the present case the roughness parameters for the <strong>SiGe</strong>-channel are Λ ∼ 100 nm <strong>and</strong> ∆ ∼ 4 nm.<br />

According to simple transport related formulas <strong>and</strong> typical parameters for the characterized<br />

p-modulation doped structures on miscut samples (Tab. 6.4) the Fermi wave number (Eq. 6.7)<br />

kF =<br />

� 4πps<br />

gsgv<br />

(6.7)<br />

<strong>and</strong> the mean free path lmfp (Eq. 6.8) are calculated. The latter is a product <strong>of</strong> the Fermi<br />

velocity <strong>and</strong> the momentum relaxation time which is also known as transport lifetime.<br />

vF = �kF<br />

m∗ µ = eτm<br />

m∗ → τm = m∗ µ<br />

e<br />

lmfp = vF · τm = �kF<br />

µ<br />

e<br />

(6.8)<br />

The estimated values are listed in Tab. 6.5. The comparison <strong>of</strong> the Fermi wave number kF <strong>and</strong><br />

the interface roughness correlation length Λ shows that the condition for short-range surface<br />

roughness (kF · Λ ≪ 1) is not fulfilled in the present case. The reciprocal Fermi wave number<br />

k −1<br />

F<br />

is by an order in magnitude smaller than the period <strong>of</strong> step-bunching Λ. Interestingly, the<br />

mean free path is by more than a factor <strong>of</strong> 10 smaller than the ripple distance as well. This<br />

means that on average several scattering events occur between two adjacent ripples. Along the<br />

bunches the mean free path is just by the discussed factor <strong>of</strong> two larger <strong>and</strong> reaches estimated<br />

values around lmfp ∼ 15 nm. Hence, generally the travel paths between successive scattering<br />

events are too small to underst<strong>and</strong> the huge anisotropy in mobility for conduction along (σ ||)<br />

<strong>and</strong> perpendicular (σ⊥) to the ripple pattern. A sophisticated model has to be developed<br />

<strong>and</strong> theoretical calculations have to be performed for an interpretation <strong>and</strong> to eventually<br />

underst<strong>and</strong> the underlying physics. For now it can only be speculated whether long-range<br />

Fermi wave number kF 1.8×10 8 [m −1 ]<br />

reciprocal Fermi wave number k −1<br />

F 5.6×10 −9 [m] → 5.6 [nm]<br />

momentum relaxation time τm 1.1×10 −13 [s] → 0.11 [ps]<br />

Fermi velocity vF 7.3×10 4 [ms −1 ]<br />

mean free path lmfp 8.2×10 −9 [m] → 8.2 [nm]<br />

Table 6.5: Estimated values for scattering related parameters.


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 107<br />

Figure 6.12: Schematic visualization <strong>of</strong> ideas to explain the differences in mobility<br />

measured parallel <strong>and</strong> across the ripple structure. The transport anisotropy is<br />

assumed to be related to an inhomogeneous distribution <strong>of</strong> scattering centers. These<br />

increased scattering potentials could dominantly be active at the bending edges <strong>of</strong><br />

the 2DHG (a). Overall conduction has to be seen as parallel <strong>and</strong> series connection<br />

<strong>of</strong> different ohmic resistances (b). According to the ripple structure in the 2DHG<br />

(c) different resistivity regions can be assumed: d) terraces <strong>and</strong> ripple flanks, e)<br />

low- <strong>and</strong> high-curvature regions, f) continuous model for (e).<br />

� Source: p-mod mobility model1.jpg<br />

surface roughness scattering directly or strain-induced scattering potentials correlated with<br />

the periodically modulated <strong>SiGe</strong>-channel limit the mobility so strongly across the bunches.<br />

Even strain-induced fluctuations in the <strong>SiGe</strong>-composition aiming at alloy scattering cannot<br />

be ruled out completely [220, 221, 222]. The origin <strong>and</strong> nature <strong>of</strong> the relevant scattering<br />

potentials involved in the presented structures remain unrevealed by now.<br />

With the present data at h<strong>and</strong> it can only be speculated about the microscopically dom-<br />

inant scattering mechanisms. Just one thing seems to be clear by now: there are different<br />

”channels” contributing unequally to the overall measured total conductivity σtot <strong>and</strong> the<br />

derived mobility µtot. The scattering centers are supposed to be aligned along the ripple<br />

pattern forming stripes <strong>of</strong> different mobility parallel to these step-bunches. The lateral ex-<br />

tension <strong>and</strong> even the position with respect to the ripples <strong>of</strong> these less-conducting paths in


108 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

Figure 6.13: Two simple macroscopic models to explain the measured mobility<br />

anisotropy. Different conductivities for the flat terraces <strong>and</strong> the steep flanks <strong>of</strong> the<br />

step-bunching structure are assumed. a) anisotropic mobility for the parallel <strong>and</strong><br />

perpendicular direction; b) isotropic mobility within the terrace regions <strong>and</strong> the<br />

flanks, respectively.<br />

� Source: p-mod mobility model2.jpg<br />

the 2DHG are unclear. Several geometries <strong>of</strong> the high-resistivity stripes can be assumed.<br />

Fig. 6.12 schematically visualizes some ideas to explain the different mobilities measured par-<br />

allel <strong>and</strong> across the ripple structure. The electrically evidenced anisotropy is assumed to be<br />

related to an inhomogeneous distribution <strong>of</strong> scattering centers. These numerous scattering<br />

potentials could be dominantly be active at the bending edges <strong>of</strong> the 2DHG (Fig. 6.12a).<br />

Overall conduction σtot has to be seen as parallel <strong>and</strong> series connection <strong>of</strong> different ohmic<br />

resistances (Fig. 6.12b). The grey-scale in Fig. 6.12c-f is used to illustrate the different resis-<br />

tivity regions in the 2DHG. Bright colors are used for higher conductivity regions <strong>and</strong> dark<br />

colors symbolize lower conductivity. A rather unrealistic model assumes different mobilities<br />

for the flat terraces <strong>and</strong> the steep ripple flanks (Fig. 6.12d), whereas low <strong>and</strong> high curvature<br />

regions at intermediate zones seem to be more realistic regions for increased resistivity. This<br />

is shown in Fig. 6.12e with a homogeneously lower conductivity at the edges. A more compli-<br />

cated assumption would invoke a continuous model (Fig. 6.12f) with a continuous variation<br />

<strong>of</strong> (isotropic) conductivities σi.<br />

Simple mathematical considerations are applied to discuss the plain macroscopic approach<br />

with a two-stripe-phase conductivity. For simplicity the width <strong>of</strong> the two channels <strong>of</strong> different<br />

conduction are supposed to be the same. For Fig. 6.13 the different channels are related to<br />

the flat terraces <strong>and</strong> the ripple flanks (compare Fig. 6.12d). The more realistic splitting <strong>and</strong><br />

re-allocating <strong>of</strong> the low-conduction areas to the high curvature regions in the 2DHG (edges<br />

in the <strong>SiGe</strong>-channel, Fig. 6.12e) is for the mathematical treatment absolutely equivalent. Ac-


6.2. EXPERIMENTAL ASPECTS AND MOBILITY ANALYSIS 109<br />

cording to Fig. 6.13 two simple macroscopic models are discussed to explain the measured<br />

anisotropy in mobility. The overall mobilities for the parallel µ ||<br />

tot <strong>and</strong> perpendicular µ⊥ tot<br />

direction are calculated using the basic formulas to analyze ohmic resistivity circuits with<br />

parallel <strong>and</strong> series connections <strong>of</strong> different ohmic resistances (Eq. 6.9).<br />

||: µ || 1<br />

tot = 2 ·<br />

�<br />

µ ||<br />

⊥:<br />

�<br />

1 + µ|| 2<br />

µ ⊥ �<br />

µ ⊥<br />

1 ·µ<br />

tot = 2 ·<br />

⊥ �<br />

2<br />

(6.9)<br />

µ ⊥ 2 +µ⊥ 1<br />

In case 1 an anisotropic mobility for the parallel <strong>and</strong> perpendicular direction with respect<br />

to step-bunching is assumed. This is motivated by interface-roughness- (parallel to step-<br />

bunching) induced scattering, which reduces the mobility only for the perpendicular direction.<br />

In a different scenario (case 2) the mobility could be isotropic within the terrace regions <strong>and</strong><br />

the flanks <strong>of</strong> the ripple pattern, respectively. Slightly different growth conditions in the<br />

steep ripple flanks could locally increase or decrease the crystal quality or at least modify<br />

the Si/<strong>SiGe</strong> interface in a way which gives differing mobilities for the different areas. The<br />

resulting formulas to estimate the mobilities for the different zones under both assumptions<br />

are comprised under Eq. 6.10.<br />

case 1: case 2:<br />

µ ||<br />

tot = µ|| 1 = µ|| 2<br />

µ ⊥ tot = µ ⊥ 1 = µ⊥ 2<br />

For case 1 the calculated mobilities µ ||<br />

1<br />

µ1,2 = µ± = µ ||<br />

tot ±<br />

� �<br />

µ ||<br />

tot<br />

� 2<br />

− µ ||<br />

tot · µ⊥ tot<br />

for equal width <strong>of</strong> conduction paths (σ1- <strong>and</strong> σ2-regions)<br />

<strong>and</strong> µ||<br />

2<br />

(6.10)<br />

coincide with the measured anisotropic mo-<br />

bilities (µ ||<br />

tot <strong>and</strong> µ⊥ tot) in both regions. Only in the second case the averaged mobilities<br />

overestimate or underestimate the mobilities for the distinct conduction paths. Hence, as-<br />

suming an equal width for the conduction channels (σ1- <strong>and</strong> σ2-regions) typical measurement<br />

results (1882LSG p: µ⊥ = 700 cm 2 V −1 s −1 , µ || = 1340 cm 2 V −1 s −1 ) lead to calculated values <strong>of</strong><br />

µ+ ∼ 2266 cm 2 V −1 s −1 <strong>and</strong> µ− ∼ 416 cm 2 V −1 s −1 (see Eq. 6.10). The higher value µ+ could be<br />

assumed for the flat (001)-oriented terrace part, whereas the lower value µ− may be used to<br />

estimate poor conduction in the steep ripple flanks. The higher value would nicely coincide<br />

with the measured mobility value for the optimized 2DHG-sample grown on a flat untilted<br />

Si(001)-substrate (1663LSG p: µ ∼ 2750 cm 2 V −1 s −1 ). However, also case 1 should not be<br />

discarded hastily, since with the assumption <strong>of</strong> low-mobility ”edges” (see Fig. 6.12e-f) the<br />

period <strong>of</strong> the correlation length Λ is reduced by a factor <strong>of</strong> two. Thus the discrepancy to the<br />

values for the mean free path perpendicular to the bunches l⊥ mfp ∼ 8 nm <strong>and</strong> the new reduced<br />

Λr ∼ 55 nm is relativized <strong>and</strong> appears more meaningful again. Furthermore, the mean free<br />

path parallel to the undulations in the <strong>SiGe</strong>-channel (l ||<br />

mfp ∼ 16 nm) is merely less than a


110 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

factor <strong>of</strong> four smaller than the interface roughness correlated length Λr. Since lmfp is just an<br />

averaged value more than a negligible portion <strong>of</strong> holes could travel across the wide terraces<br />

<strong>and</strong> reach the edges <strong>of</strong> the neighboring flanks where severe scattering occurs.<br />

6.3 Outlook <strong>and</strong> Perspectives<br />

For future investigations several new experimental setups are proposed. First, a p-modula-<br />

tion-doped structure could be grown onto a highly miscut substrate on top <strong>of</strong> a flat high-<br />

temperature Si-buffer. Thus a highly miscut substrate could be utilized to match extensively,<br />

over its whole surface area, the inclination angle with the slope <strong>of</strong> the ripple flanks <strong>of</strong> the<br />

4 ◦ miscut samples showing step-bunching. With this additional reference experiment the<br />

found decrease in mobility perpendicular to the ripples could be unambiguously linked to<br />

the mesoscopic periodic undulations <strong>of</strong> step-bunching. Although the higher density <strong>of</strong> closely<br />

spaced atomic steps in miscut direction is not expected to cause a significant detectable<br />

mobility anisotropy, this way it should be possible to rule out a transport anisotropy for a<br />

planar 2DHG on tilted substrates definitely.<br />

Up to now, only electrical measurements for the two prominent directions – parallel <strong>and</strong><br />

perpendicular to the ripple structure – were conducted. The Hall-bars could be oriented in<br />

any arbitrary direction to analyze the angle dependence <strong>of</strong> the mobility. This way according<br />

to the chosen angle α any corrugation period along the measurement path can be adjusted<br />

after Eq. 6.11.<br />

Λα (α, Λ) =<br />

Λ<br />

cos (α) , Λα > Λ (6.11)<br />

The effective distance between the undulations Λα depends on the ripple period Λ <strong>and</strong> the<br />

rotation angle α, where an angle α = 0 represents the measurement direction perpendicular<br />

to the ripples. Generally p-MODQW grown on ripple templates featuring different periods<br />

<strong>and</strong> amplitudes would be interesting to compare. Especially short-periodic ripple templates<br />

(Λ < 50 nm) would be favorable for further experiments. However, it can be doubted whether<br />

the realization <strong>of</strong> such structures based on fully self-organized means is possible.<br />

New optical masks have to be designed which employ the opportunity to put self-aligned gates<br />

on top <strong>of</strong> the Hall-bars. This would be essential to precisely adjust the carrier density ps in<br />

the <strong>SiGe</strong>-channel <strong>and</strong> to get reliable results. Although the position <strong>of</strong> the wave-function is<br />

shifted with respect to the Si/<strong>SiGe</strong>-interface by making use <strong>of</strong> the top-gate, still, ps-dependent<br />

mobility data could help to identify the dominant scattering mechanisms.<br />

Maybe the most important goal should be to optimize the p-modulation doped structure to<br />

achieve higher mobilities. Accordingly, the relevant parameter, namely the mean free path


6.3. OUTLOOK AND PERSPECTIVES 111<br />

lmfp, would be enhanced. In an ultimately ideal case the mobility would thereafter only be<br />

limited for the hole transport in perpendicular direction <strong>and</strong> all the scattering events would<br />

occur due to the interface-roughness-induced scattering potential extending along the ripple<br />

structure. Several ways are assumed to achieve this mobility improvement. The layer thick-<br />

ness <strong>of</strong> Si-spacer <strong>and</strong> <strong>SiGe</strong>-channel may require optimization. So an increase in thickness<br />

towards 150–200 ˚A might be beneficial. Of course, also the growth temperature in relation<br />

with the growth rates are crucial parameters for the final quality <strong>of</strong> the structures <strong>and</strong> have<br />

to be reconsidered.<br />

Switching to the inverted p-MODQW architecture with the doping region beneath the chan-<br />

nel could also be advantageous. In this approach an, at a short scale, smoother Si/<strong>SiGe</strong><br />

interface is expected, since Ge-segregation should be less pronounced for the lower interface<br />

when compared with the upper Si/<strong>SiGe</strong>-interface [6].<br />

However, lowering the Ge-content x in the <strong>SiGe</strong>-channel should give the largest contribution<br />

for the desired increase in mobility (see Fig. 6.3). Estimated Ge-contents as low as x ∼ 5–10%<br />

should give a boost in mobility by a factor <strong>of</strong> 10. Too low Ge-contents are not useful as the<br />

carriers are no longer well-confined in the shallow <strong>SiGe</strong> quantum well, <strong>and</strong> the wave-function<br />

is less defined at the interface, which is a drawback for the characterization <strong>of</strong> the surface<br />

roughness related scattering.<br />

Also with the currently available samples further measurements could be performed at lower<br />

temperatures in a different cryostat. Thus cooling down the samples to 300 mK in the 3 He-<br />

refrigerator or even to 30 mK in the dilution refrigerator should significantly suppress thermal<br />

effects <strong>and</strong> enhance significantly the otherwise blurred quantum effects. Especially the onset<br />

<strong>of</strong> SdH-oscillations is expected to be shifted to lower magnetic fields. There, spin splitting<br />

does not distort the periodic pattern, <strong>and</strong> hence, the position <strong>and</strong> amplitude <strong>of</strong> the SdH-<br />

oscillation can be evaluated according to the Dingle analysis to extract the quantum lifetime<br />

τq [216]. The ratio <strong>of</strong> the transport lifetime <strong>and</strong> the quantum lifetime τm/τq is known to re-<br />

veal information on the nature <strong>of</strong> the scattering potential <strong>and</strong> the scattering mechanism [6].<br />

Low-temperature magnetoresistance experiments with tilted magnetic fields are also inter-<br />

esting to study the edge-channel transport in the Si/<strong>SiGe</strong> QH-system [223]. The existence<br />

<strong>of</strong> stripe phases causing a strong transport anisotropy at half filled L<strong>and</strong>au levels (LL) was<br />

found first for high-mobility GaAs/AlGaAs heterostructures [224, 225, 226]. There a signifi-<br />

cantly high in-plane magnetic field increases the Zeeman splitting <strong>and</strong> causes a cross-over <strong>of</strong><br />

adjacent L<strong>and</strong>au levels, i.e. the spin-down configuration <strong>of</strong> the higher L<strong>and</strong>au level (N+1, ↓)<br />

crosses the energetic position <strong>of</strong> the spin-up state (N, ↑). For the edge-channel transport<br />

<strong>of</strong> electrons in Si/<strong>SiGe</strong> heterostructures it was revealed that inter-edge-channel scattering is


112 CHAPTER 6. P-MODULATION DOPING AND MOBILITY ANALYSIS<br />

strongly suppressed over macroscopic distance between (0 ↓, 0 ↑) edge channels, whereas it is<br />

promoted between (0 ↓, 1 ↓) edge channels [227]. Thus, it could be interesting to study the<br />

influence <strong>of</strong> the anisotropic ripple structure, which is present in our 2DHG-samples, on such<br />

a stripe-phase formation <strong>and</strong> the inter-edge-channel scattering.<br />

In conclusion, more measurement data have to be collected from different electrical ex-<br />

periments to develop a more sophisticated model. In a second step, based on the extracted<br />

parameters, theoretical simulations have to be applied to fit the experimental results <strong>and</strong><br />

to learn more about the microscopic physics <strong>of</strong> the limiting scattering mechanisms for p-<br />

modulation doped structures.


Part III<br />

Additional Work<br />

113


Chapter 7<br />

Transient-Enhanced Si Diffusion on<br />

Natural-Oxide-Covered Si(001)<br />

This chapter comprises the work on ”Transient-enhanced Si diffusion on natural-oxide-covered<br />

Si(001) nanostructures during vacuum annealing”, which was started in the course <strong>of</strong> the<br />

diploma thesis [1]. Further experiments proposed as outlook there were conducted here in<br />

continuation. Thus all the related results are discussed here in compact form on the basis <strong>of</strong><br />

the two resulting publications [228, 229] in order to keep the context.<br />

The morphology <strong>of</strong> patterned Si(001) wire-templates after annealing was studied by several<br />

techniques. An enormous Si-mass-transport on the Si-surface at usual oxide desorption tem-<br />

peratures around 900 ◦ C under UHV-conditions was found. Heat treatment <strong>of</strong> 5 min trans-<br />

forms the initially rectangular wire pr<strong>of</strong>iles with a height <strong>of</strong> 300 nm to flat (< 100 nm) <strong>and</strong><br />

faceted triangular ridges exhibiting thermodynamically preferred {111}- <strong>and</strong> {113}-facets.<br />

It was found that the natural SiO2 on the predefined wire pattern must be responsible for<br />

the degradation <strong>of</strong> the wire structure. Removing the SiO2-layer from the Si wires ex situ<br />

with an HF-dip preserves the rectangular structures during high temperature annealing. The<br />

Si/SiO2-interface was investigated with high-resolution transmission electron microscopy to<br />

image the Si wire surface <strong>and</strong> the natural-oxide-layer in detail.<br />

The new experiment employed low-temperature Ge-deposition to generate in situ a Ge-<br />

contrast layer which conserved the Si/SiO2 interface for later HRXTEM studies (high reso-<br />

lution cross-sectional transmission electron microscopy). This way the role <strong>of</strong> the SiO2-cover<br />

in the vast transient Si surface diffusion process could be enlightened.<br />

115


116 CHAPTER 7. TRANSIENT-ENHANCED SI DIFFUSION<br />

Figure 7.1: (a) HRXTEM image <strong>of</strong> the outer edge including zoom-in <strong>of</strong> a Si-wire<br />

after RIE. (b) Cross-sectional SEM image <strong>of</strong> 800 nm Si-wire template tilted slightly<br />

out <strong>of</strong> the wire-direction. (c) Schematic drawing indicating the respective location<br />

<strong>of</strong> image (a) with regard to the complete wire cross sections.<br />

� Source: TEM n SEM rect.jpg<br />

7.1 Introduction<br />

State-<strong>of</strong>-the-art semiconductor devices have reached critical dimensions well below 100 nm.<br />

For shrinking the device dimensions it becomes increasingly important to control the shape<br />

<strong>and</strong> size <strong>of</strong> small structures during subsequent processing. Our investigations concentrate<br />

on the shape stability <strong>of</strong> Si nanostructures during vacuum annealing at around 900 ◦ C for a<br />

few minutes. Such thermal steps are typically employed for natural oxide desorption prior<br />

to epitaxial growth, but similar thermal budgets are frequently required during device pro-<br />

cessing, e.g. after ion implantation. While the shape evolution <strong>of</strong> structured Si surfaces into<br />

thermodynamically stable facets ({111} <strong>and</strong> {113}) is well described [230, 231], most <strong>of</strong> the<br />

experimental studies employed long-term, high-temperature anneals. Also, quite <strong>of</strong>ten exotic<br />

annealing procedures (e.g. a flash to 1200 ◦ C [232]) <strong>and</strong> direct current heating (prone to<br />

electro-migration artifacts [233]) were used. The discussed experiments here only employed<br />

cleaning <strong>and</strong> annealing procedures adapted from st<strong>and</strong>ard Si device processes.<br />

7.2 Template Preparation<br />

For our experiments Si wire-arrays were processed on 17 mm × 17 mm Si-wafer pieces that<br />

were cut from (001)-orientated st<strong>and</strong>ard Si-wafers supplied by Wacker Siltronic (4”, p-doped,


7.3. EXPERIMENTAL RESULTS 117<br />

> 1000 Ωcm resistivity, 525 µm thick). Stripe-patterns with a period <strong>of</strong> typically 800 nm along<br />

the [110]-direction were produced by holographic lithography (λ = 457.9 nm) <strong>and</strong> subsequent<br />

reactive ion etching in an SF6 process. Typical etch depths were 250–300 nm. The residual<br />

photoresist-mask was removed by plasma-ashing. The homogeneity <strong>of</strong> the fabricated wire-<br />

structures was checked with a Park Scientific atomic force microscope (AFM) operated in a<br />

non-contact mode with Olympus TESP tips. The shape <strong>of</strong> the etched pr<strong>of</strong>iles was controlled<br />

with cross-sectional views in a JEOL JSM 6400 scanning electron microscope (SEM) or at<br />

200 keV with a Jeol 2011 FasTEM transmission electron microscope. The wire-templates<br />

show rectangular pr<strong>of</strong>iles with small corner rounding <strong>of</strong> < 5 nm (see Fig. 7.1).<br />

7.3 Experimental Results<br />

The wire-templates that had been pre-characterized with AFM were successively cleaned in<br />

trichlorethylene, acetone, methanol using an ultrasonic bath, rinsed with deionized water (DI-<br />

H2O) <strong>and</strong> dried using the flow <strong>of</strong> a nitrogen nozzle. Additionally, the Si-pieces were cleaned<br />

with a mixture <strong>of</strong> sulphuric acid (H2SO4, 96%) <strong>and</strong> hydrogen peroxide (H2O2, 30%) for 10 min<br />

(H2SO4 : H2O2 = 5 : 1) to remove organic impurities, <strong>and</strong> again rinsed with DI-H2O. The con-<br />

ventional cleaning procedure was concluded with the st<strong>and</strong>ard RCA clean [71, 72] without any<br />

HF-treatments. The RCA-clean leaves a passivating SiO2-layer <strong>of</strong> ≈ 2 nm on the Si-surface.<br />

Additionally, for some experiments the samples were HF-dipped (HF 40% : DI-H2O ≈ 1 : 5)<br />

immediately before substrate-transfer into the UHV <strong>of</strong> our Riber SIVA 45 MBE-system.<br />

This optional HF dip removes the oxide <strong>and</strong> provides a hydrogen terminated surface that is<br />

stable against oxidation for a few hours. After introduction <strong>of</strong> the samples into the UHV-<br />

environment <strong>of</strong> the MBE-chamber the prepatterned substrates were annealed to desorb the<br />

SiO2 from the Si-surface <strong>and</strong> to provide a clean surface for epitaxial overgrowth. After sample<br />

degassing at 550 ◦ C the temperature was ramped up to 900–950 ◦ C at 230 ◦ C/min, kept at<br />

this maximum temperature for 1–5 min, <strong>and</strong> was then rapidly quenched to room temperature.<br />

The substrate temperature was controlled by a calibrated thermocouple (≈ 15 ◦ C), located in<br />

the radiation area between wafer <strong>and</strong> heater, <strong>and</strong> additionally monitored with a pyrometer.<br />

The samples were characterized before <strong>and</strong> after annealing on air by AFM to reveal structural<br />

changes. Selected samples were imaged by high-resolution cross-sectional transmission elec-<br />

tron microscopy (HRXTEM). Fig. 7.2 <strong>and</strong> Fig. 7.3 show 3D-AFM data <strong>and</strong> AFM line scans<br />

<strong>of</strong> a sample covered with natural oxide before <strong>and</strong> after annealing at 950 ◦ C for 4 min. It is<br />

obvious that a dramatic morphological change has taken place that converted the originally<br />

300 nm high, almost rectangular wire structures into multiple faceted wires that have lost up


118 CHAPTER 7. TRANSIENT-ENHANCED SI DIFFUSION<br />

Figure 7.2: 3D-AFM images <strong>of</strong> a natural-oxide covered sample before (a) <strong>and</strong><br />

after (b) annealing at 950 ◦ C for 4 min.<br />

� Source: Wire-Samples 3D-AFM.jpg<br />

to 75% <strong>of</strong> their original height (see Fig. 7.3a). The central, triangular part <strong>of</strong> the wires is<br />

made up <strong>of</strong> {113}-facets, whereas the lower part <strong>of</strong> the cross section is inclined against the<br />

[100]-direction by about 4.5 ◦ (see Fig. 7.3b). To uncouple the oxide-desorption process from<br />

the Si-mass-transport phenomenon the experiment was repeated with an identical sample<br />

that had the natural oxide removed in 4% HF. In that case no noticeable change <strong>of</strong> the wire<br />

cross section was observed after annealing. To further investigate the relation between oxide<br />

desorption <strong>and</strong> surface diffusion, the shape evolution <strong>of</strong> the wire arrays was studied. Fig. 7.4<br />

shows the normalized wire height after annealing at 900 ◦ C <strong>and</strong> 950 ◦ C versus annealing time<br />

for samples with <strong>and</strong> without a final HF dip. The upper curve in Fig. 7.4 indicates that there<br />

is no visible change <strong>of</strong> the wire-height without an initial SiO2-layer. Especially the data point<br />

for 4 min annealing at 950 ◦ C proves that after an HF-dip also an extended heat-treatment<br />

Figure 7.3: AFM line scans <strong>of</strong> a natural-oxide covered sample before (broken line)<br />

<strong>and</strong> after annealing at 950 ◦ C for 4 min (solid line). (a) Direct comparison <strong>of</strong> AFM<br />

line scans <strong>of</strong> wire-templates before <strong>and</strong> after annealing. (b) The central, triangular<br />

part <strong>of</strong> the annealed wire is made up <strong>of</strong> {113}-facets<br />

� Source: Wire-Samples Linescans.jpg


7.3. EXPERIMENTAL RESULTS 119<br />

Figure 7.4: Normalized wire height vs. annealing time for samples with (RCA),<br />

<strong>and</strong> without (HF) natural-oxide. The arrow links the data points from a sample<br />

that was annealed first without, <strong>and</strong> then again with natural oxide coverage.<br />

� Source: Annealing plot.jpg<br />

does not affect the Si-nanostructures. The lower curve for thermal oxide-desorption reveals<br />

that the wire height <strong>of</strong> the oxide-covered samples initially decreases with increasing annealing<br />

time, <strong>and</strong> then saturates after about 3 min. The onset <strong>of</strong> saturation agrees well with the time<br />

required for complete oxide desorption at 900 ◦ C [234]. The saturation behavior is consistent<br />

with the results <strong>of</strong> the oxide-free samples, which do not show any structural changes. To rule<br />

out that possible hydro-carbon contaminations from the HF treatment had interfered with<br />

Si surface diffusion, a control experiment was performed: A sample that had been annealed<br />

after an HF-dip underwent another RCA clean to generate a natural oxide layer. Repeating<br />

the annealing step on this recycled sample resulted in the same loss <strong>of</strong> height as expected<br />

for an RCA-cleaned sample (arrow in Fig. 7.4). It can be ruled out that possible hydrocar-<br />

bon contaminations after HF treatment have impeded the surface migration <strong>of</strong> Si after an<br />

HF dip. These would have reacted with Si during the first annealing step to SiC, which is<br />

stable against the RCA-clean for the second experiment. Thus the greatly enhanced surface<br />

mass transport on Si(001) can unambiguously be associated to the presence <strong>of</strong> a thin layer<br />

<strong>of</strong> natural oxide. It is also interesting to note that annealing at 950 ◦ C leads to a saturated<br />

wire height that is significantly lower than the saturation height after a 900 ◦ C anneal. This<br />

is a clear indication that the activation energies for the oxide desorption mechanism <strong>and</strong> for<br />

the Si mass transport are different. HRXTEM images with the viewing direction along the


120 CHAPTER 7. TRANSIENT-ENHANCED SI DIFFUSION<br />

Figure 7.5: HRXTEM images <strong>of</strong> a natural-oxide covered sample after annealing<br />

for 60 s at 900 ◦ C. Faceted wire-top (a) <strong>and</strong> wire-edge (b) showing thermodynamically<br />

stable facets. (c) Multiple faceted wire flank. (d) Schematic drawings indicating<br />

the respective location <strong>of</strong> the HRXTEM images with regard to the complete<br />

wire cross sections.<br />

� Source: TEMs facet-evolution.jpg


7.4. DETAILED CHARACTERIZATION AND DISCUSSION 121<br />

[110]-oriented wires were recorded to study details <strong>of</strong> facet formation. The samples were<br />

covered ex-situ with a polycrystalline Ti-film to enhance contrast. Fig. 7.5a shows the cross<br />

section <strong>of</strong> the uppermost part <strong>of</strong> a wire after 60 s at 900 ◦ C. A top (001)-facet still exists, but<br />

the former {110}-sidewall facet (compare Fig. 7.1) has been transformed into a lower {111}-,<br />

<strong>and</strong> an upper {113}-facet (see Fig. 7.5b). We also found some regions with multiple facet<br />

orientations (see Fig. 7.5c), <strong>and</strong> transitions regions that cannot be assigned to known low<br />

energy facets. The respective location <strong>of</strong> the HRXTEM images with regard to the wire cross<br />

sections is sketched in Fig. 7.5d.<br />

7.4 Detailed Characterization <strong>and</strong> Discussion<br />

The vast Si-self-diffusibility on the Si-surface was also found with other experiments. Fig. 7.6a<br />

shows an XTEM-image <strong>of</strong> a SiO2 wire structure on Si(001) annealed for 6 min at 950 ◦ C. It can<br />

be clearly seen that a large amount <strong>of</strong> Si has diffused towards the SiO2 structures, resulting<br />

in a bowed Si-surface between two neighboring SiO2-ridges [12]. The SiO2 wires seemingly<br />

Figure 7.6: (a) XTEM image <strong>of</strong> a SiO2 wire-structure on Si(001) after annealing<br />

for 6 min @ 950 ◦ C. As indicated with the schematic graph (b) the Si surface atoms<br />

diffuse towards the ridges <strong>and</strong> react with the SiO2 to volatile SiO molecules. This<br />

process provides the lateral undercut <strong>of</strong> the SiO2 wires <strong>and</strong> leads to the bowed Sisurface<br />

between neighboring oxide wires. The dashed line indicates the Si-surface<br />

before annealing [12].<br />

� Source: TEM oxide-wires.jpg<br />

attract the Si-atoms. Evaluating the contact-angles between the SiO2-structures <strong>and</strong> the Si-<br />

surface yields values slightly exceeding 90 ◦ . Apparently, wetting cannot be attributed as the<br />

driving force for the Si pile-up next to the SiO2-wires. The transport <strong>of</strong> the large amount <strong>of</strong> Si<br />

within a short period <strong>of</strong> time at relatively low temperatures is a pro<strong>of</strong> for the high Si-surface<br />

mobility. The lateral undercut <strong>of</strong> the SiO2-structures can be explained using the well-known<br />

oxide-desorption reaction at elevated temperatures around 900 ◦ C. This mechanism for oxide-<br />

desorption for the patterned SiO2-template is illustrated in Fig. 7.6b <strong>and</strong> follows the reaction


122 CHAPTER 7. TRANSIENT-ENHANCED SI DIFFUSION<br />

path [235]<br />

SiO2 + Si → 2 SiO ↑ . (7.1)<br />

In contrast to SiO2, SiO is volatile at typical annealing temperatures around 900 ◦ C. As long<br />

as a continuous oxide film exists, SiO forms predominantly at the interface to the oxide [236].<br />

Thermal desorption would then require SiO diffusion through the SiO2 film. It was, how-<br />

ever, found that oxide desorption occurs mainly via the expansion <strong>of</strong> voids that form during<br />

an early stage in the oxide [237, 238]. Surface diffusion on the Si surface, which becomes<br />

exposed within these voids, is high even at 900 ◦ C [239]. This allows Si transport towards<br />

the periphery <strong>of</strong> the voids, were reaction Eq. 7.1 <strong>and</strong> the desorption <strong>of</strong> SiO can readily oc-<br />

cur. On flat substrates no correlation between void formation <strong>and</strong> interface structures has<br />

been found [237]. Instead, void nucleation has been associated with contaminations on or<br />

in the oxide [238, 240]. It is, however, not clear whether this applies to our wire-structured<br />

substrates, which expose almost atomically sharp intersections between facets. Such a large<br />

perturbation might be expected to affect the nucleation <strong>and</strong> anisotropy <strong>of</strong> void formation.<br />

Therefore XTEM decoration experiments were conducted to visualize the location <strong>of</strong> the<br />

voids with respect to the wire structures. For that purpose natural-oxide covered samples<br />

with a wire period <strong>of</strong> 400 nm that had undergone an annealing cycle for 1 min at 900 ◦ C were<br />

covered in-situ by a 20 nm thick Ge layer at a deposition temperature < 100 ◦ C to passivate<br />

the underlying Si/SiO2 interface. Under these conditions the Ge film becomes amorphous on<br />

areas above the Si surface with intact SiO2-film, <strong>and</strong> forms polycrystalline Ge-grains in the<br />

voids where the deposited Ge is in direct contact with the crystal-order <strong>of</strong> the Si-surface (see<br />

Fig. 7.7a). Because <strong>of</strong> the large mass contrast, the residual natural oxide is clearly visible as<br />

a lighter stripe between the crystalline Si-substrate <strong>and</strong> the amorphous Ge-cap. The darker<br />

appearance <strong>of</strong> the poly grains decorates the void regions even in the lower resolution images<br />

in Fig. 7.7b, which show cross sections <strong>of</strong> several {113}-faceted wires. It is obvious that the<br />

voids are not correlated with the period <strong>of</strong> the wire template. It is especially striking that<br />

most <strong>of</strong> the ridges are still completely covered by oxide despite their shape transformation<br />

from an originally rectangular cross section. This clearly indicates that the strongly enhanced<br />

diffusion occurs predominantly underneath the oxide, <strong>and</strong> even more, that the oxide follows<br />

the shape transformation. The voids do have some influence, as can be seen at the arrow-<br />

marked ridge in Fig. 7.7b that coincides with a void, <strong>and</strong> has become even flatter than the<br />

neighboring oxide-covered ridges. A similar effect might have caused the height fluctuations<br />

in Fig. 7.2b. Nevertheless, most <strong>of</strong> the mass transport takes place beneath the oxide, <strong>and</strong> is<br />

most likely associated with the SiO phase that forms with substantial partial pressure [236]<br />

at the Si/SiO2 interface upon annealing.


7.4. DETAILED CHARACTERIZATION AND DISCUSSION 123<br />

Figure 7.7: HRXTEM images <strong>of</strong> an annealed sample that was in-situ covered<br />

with Ge. (a) Voids in the oxide are decorated by polycrystalline Ge (p-Ge), which<br />

appears darker than the amorphous Ge (a-Ge) that forms on SiO2. (b) Lower<br />

resolution images <strong>of</strong> several Si wires.<br />

� Source: TEMs Ge-cover.jpg<br />

The presented study shows that the initially rectangular pr<strong>of</strong>iles <strong>of</strong> periodic Si-wire-arrays<br />

are degraded during radiative annealing in UHV for 1–5 min at 900–950 ◦ C to {113}-faceted<br />

trapezoids concomitant with a loss <strong>of</strong> up to 75% <strong>of</strong> the structure height. This shape trans-<br />

formation requires drastically enhanced mass transport, which occurs only in the presence <strong>of</strong><br />

SiO2 on the surface, <strong>and</strong> consequently ceases after complete oxide desorption.


124 CHAPTER 7. TRANSIENT-ENHANCED SI DIFFUSION


Chapter 8<br />

Germanium Source Reconstruction<br />

The initial MBE-system as supplied by RIBER consisted <strong>of</strong> 3 single-crucible e-guns. A large<br />

Temescal SFIH-270-3 canon with a crucible volume <strong>of</strong> 156 cm 3 was <strong>and</strong> is still used as electron<br />

beam evaporator for silicon, which is the main matrix material in our <strong>SiGe</strong>C apparatus.<br />

Two small Temescal SFIH-270-2 canons with a capacity <strong>of</strong> 40 cm 3 were originally used for the<br />

other two matrix materials, germanium <strong>and</strong> carbon, respectively. Therefore the evaporable<br />

amount <strong>of</strong> germanium was fairly limited <strong>and</strong> even further diminished as the source material is<br />

not evaporated directly from the copper hearth <strong>of</strong> the e-gun, but out <strong>of</strong> a Si-liner. This action<br />

is taken to prevent Cu in-diffusion into the Ge material, which is known to be detrimental es-<br />

pecially for photoluminescence results. The carbon e-beam evaporator was shut down a long<br />

time ago <strong>and</strong> a carbon sublimation cell, which was better to control, was used instead. Thus<br />

there was enough space for a larger Ge evaporator which could provide several improvements:<br />

The larger amount <strong>of</strong> germanium exp<strong>and</strong>s the refilling cycle <strong>and</strong> therefore there should be no<br />

need to open the growth chamber frequently, which improves the vacuum conditions <strong>and</strong> also<br />

the quality <strong>of</strong> the grown layers. Especially for n-modulation-doped structures, where thick<br />

graded buffers with up to 30% germanium are involved, a lot <strong>of</strong> material is needed. Addi-<br />

tionally, with the larger crucible higher growth rates are feasible. Instead <strong>of</strong> about 0.2 ˚A/s in<br />

the small evaporator, with the new large e-gun 1.2 ˚A/s can be readily realized. Due to higher<br />

available growth rates it is now possible to speed up growth <strong>and</strong> cut down growth times<br />

from > 12 hours to about 4 hours for n-modulation doped structures involving thick graded<br />

buffers. As our group started up in a new field with growth on Ge-substrates, excessive use<br />

<strong>of</strong> Ge-material is made especially for pure Ge-buffers. The maximum Ge-rate is chosen to<br />

be 1.25 ˚A/s to keep a meaningful relation to the maximum available Si-rate (2.5 ˚A/s). The<br />

limited sensor range yields an optimized sensitivity <strong>and</strong> resolution in the low growth-rate<br />

125


126 CHAPTER 8. GERMANIUM SOURCE RECONSTRUCTION<br />

region, which is extremely important for Ge-dot growth with sub-monolayer accuracy.<br />

As already mentioned, there were several good reasons for updating our machine with a pow-<br />

erful Ge-source. We decided to buy another Temescal SFIH-270-3 e-gun for Ge-evaporation<br />

to stay compatible with our electronics, <strong>and</strong> to keep our stock <strong>of</strong> spare-parts small. On the<br />

basis <strong>of</strong> original 2D-drawings <strong>of</strong> our whole MBE-system supplied by Episerve (i.e. the lo-<br />

cal distributor for RIBER products) 3D-drawings had to be generated to fit all parts into<br />

our apparatus. Although the former water-cooled ro<strong>of</strong> <strong>of</strong> the Ge/C-assembly <strong>and</strong> the Si-<br />

evaporation assembly were used as model for the new Ge-assembly, several aspects had to<br />

be taken into account. The water-cooled ro<strong>of</strong> was designed to reach a maximum <strong>of</strong> thermal<br />

shielding with regard to the chamber walls. The ro<strong>of</strong> was extended to block excessive heat<br />

radiation from the melt <strong>and</strong> to keep the surrounding temperature as low as possible, which<br />

improves the growth pressure. Baffle plates inserted into the hollow space <strong>of</strong> the ro<strong>of</strong> ensure<br />

Figure 8.1: Top-view <strong>of</strong> rendered 3D-drawings <strong>of</strong> most important parts in our<br />

<strong>SiGe</strong>C MBE-system with new Ge-evaporation assembly.<br />

� Source: GC Drawing51 3d 2 new07Feb2006 3a label.jpg


Figure 8.2: Elevated front-view <strong>of</strong> rendered 3D-drawings <strong>of</strong> most important parts<br />

in our <strong>SiGe</strong>C MBE-system with new Ge-evaporation assembly.<br />

� Source: GC Drawing51 3d 2 new07Feb2006 3c.jpg<br />

proper water-flow <strong>and</strong> cooling. Most important were the openings <strong>and</strong> bores in the ro<strong>of</strong> to<br />

ensure an unobstructed intervisibility between crucible <strong>and</strong> substrate, viewport window <strong>and</strong><br />

flux sensing units (Sentinel, QMS), respectively. Additionally, the space for the installation<br />

<strong>of</strong> a long-term planned phosphorous effusion cell had to be provided. Therefore, to check all<br />

the details, <strong>and</strong> to perfectly tailor the new evaporation assembly into the existing system the<br />

elaborate designing in 3D <strong>of</strong> all involved parts was inevitable.<br />

The 3D-drawings <strong>of</strong> the growth-chamber were generated with AutoCAD R14 with as many<br />

details as required. Only the parts <strong>of</strong> the Ge-ro<strong>of</strong> assembly had to be constructed with every<br />

detail. Rendered 3D-views <strong>of</strong> the most important parts in our <strong>SiGe</strong>C MBE-system already<br />

with the new Ge-evaporation assembly are shown in Fig. 8.1 <strong>and</strong> Fig. 8.2.<br />

In Fig. 8.3 various rendered 3D-views <strong>of</strong> the final Ge-ro<strong>of</strong> assembly are depicted. The images<br />

include all parts <strong>of</strong> the custom-made construction including main-flange, water-cooled ro<strong>of</strong>,<br />

underframe for e-beam source mounting <strong>and</strong> all waterlines. The main flange was provided<br />

with several flanges for various feedthroughs, such as the high-voltage connection, electron<br />

127


128 CHAPTER 8. GERMANIUM SOURCE RECONSTRUCTION<br />

Figure 8.3: Various rendered 3D-views <strong>of</strong> the Ge-ro<strong>of</strong> assembly. The images<br />

include all parts <strong>of</strong> the custom-made construction including main-flange, watercooled<br />

ro<strong>of</strong>, underframe for e-beam source mounting <strong>and</strong> all waterlines.<br />

� Source: Ge-ro<strong>of</strong>02.jpg<br />

beam sweep control, water-cooling <strong>and</strong> linear shutter-motion. The hole in the underframe is<br />

necessary to connect the electron beam source with the high-voltage cables.<br />

Fig. 8.4 presents a photograph <strong>of</strong> the new Ge-evaporation unit <strong>and</strong> two group members ad-<br />

justing the Ge-shutter. Fig. 8.5 shows the fully assembled Ge-assembly mounted on the<br />

”service-trolley” with installed electron gun, shielding plates, shutter mechanics, electric <strong>and</strong>


water lines.<br />

Figure 8.4: Photograph taken during assembling <strong>of</strong> Ge-evaporation unit showing<br />

two group members adjusting the Ge-shutter.<br />

� Source: Photo Ge-ro<strong>of</strong>01.jpg<br />

Due to s<strong>of</strong>tware incompatibilities, the company, which was assigned with manufacturing, had<br />

to redraw the construction drawings. Fig. 8.6 <strong>and</strong> 8.7 show the final drawings <strong>of</strong> VTS-Schwarz<br />

that match our 3D-drawings <strong>and</strong> thus were accepted for fabrication.<br />

Figure 8.5: Photograph <strong>of</strong> Ge-evaporation assembly ready to slide into the growth<br />

chamber.<br />

� Source: Photo Ge-ro<strong>of</strong>03.jpg<br />

129


130 CHAPTER 8. GERMANIUM SOURCE RECONSTRUCTION<br />

Figure 8.6: Part 1 <strong>of</strong> final construction drawings for Ge-evaporation assembly redrawn by VTS-Schwarz.<br />

� Source: Ge-evaporation-assembly Schwarz04Mar04 2 ok 1.jpg


Figure 8.7: Part 2 <strong>of</strong> final construction drawings for Ge-evaporation assembly redrawn by VTS-Schwarz.<br />

� Source: Ge-evaporation-assembly Schwarz04Mar04 2 ok 2.jpg<br />

131


132 CHAPTER 8. GERMANIUM SOURCE RECONSTRUCTION


Appendix A<br />

Calibration <strong>and</strong> Characterization <strong>of</strong><br />

Sources<br />

In this part the influence <strong>of</strong> the Ge source reconstruction (see Ch. 8) on the <strong>SiGe</strong> calibration<br />

is discussed. Photoluminescence measurements on single <strong>SiGe</strong> quantum wells were performed<br />

to verify the cleanliness <strong>of</strong> the new Ge source material.<br />

A.1 Calibrations<br />

The substitution <strong>of</strong> the two small electron guns (for Ge <strong>and</strong> C) with the larger Ge gun re-<br />

sulted in a new position <strong>of</strong> the Ge crucible. The whole geometry <strong>of</strong> Ge evaporation is therefore<br />

changed, which significantly alters the distribution <strong>of</strong> Ge in grown epilayers across a wafer.<br />

For <strong>SiGe</strong> calibration it is now st<strong>and</strong>ard to grow a single <strong>SiGe</strong> epilayer with about 750–1000 ˚A<br />

thickness <strong>and</strong> a Ge-content between 10% <strong>and</strong> 20% (see Tab. A.1 for details). The parameters<br />

<strong>of</strong> the pseudomorphic epilayer are calculated from x-ray data. A single ω-2θ scan around the<br />

symmetric (004)-reflex <strong>of</strong> Si reveals the sharp Si substrate peak <strong>and</strong> the <strong>SiGe</strong> peak shifted<br />

to lower angles. The peak positions indicate the perpendicular lattice constants a⊥ <strong>of</strong> Si<br />

<strong>and</strong> the pseudomorphic <strong>SiGe</strong> epilayer. The position <strong>of</strong> the <strong>SiGe</strong> peak is used to extract the<br />

Ge-content x. This is accessible due to the monotonic relation <strong>of</strong> the Ge-content x <strong>and</strong> the<br />

fully strained pseudomorphic lattice constant <strong>of</strong> the epilayer a<strong>SiGe</strong>,⊥. The thickness <strong>of</strong> the<br />

epilayer can either be derived from the width <strong>of</strong> the <strong>SiGe</strong> peak, or, alternatively, from the<br />

spacing <strong>of</strong> the thickness oscillations [241].<br />

The measured x-ray data were compared <strong>and</strong> fitted with simulated spectra calculated by the<br />

Matlab-program ”simx” developed <strong>and</strong> maintained by our X-Ray group. This is a convenient<br />

133


134 APPENDIX A. CALIBRATION AND CHARACTERIZATION OF SOURCES<br />

way to extract the Ge-content <strong>and</strong> the total <strong>SiGe</strong> epilayer thickness. From these two values<br />

the Si <strong>and</strong> Ge rate can easily be computed.<br />

Usually the calibration is only actualized for the wafer center. Nevertheless, all sources show<br />

a distinct flux distribution across the substrate. Therefore, also growth on larger substrates,<br />

e.g. on a whole 4”-wafer, does not yield identical sample material. Out-<strong>of</strong>-center regions do<br />

not supply that high quality material due to glide lines arising from thermal stress during<br />

the high temperature oxide desorption step at the wafer rims.<br />

Often, not much material is needed for later experiments or special substrate miscuts are<br />

used. So, we save substrate material by growing on small pieces, which are mounted in all-Si<br />

adapter wafers (see Sec. 2.3). Several adapters are available which usually can hold several<br />

pieces. The position <strong>of</strong> the samples is sometimes <strong>of</strong>f-center in the growth chamber. Therefore<br />

the calibration data <strong>and</strong> the distribution across the substrate should be available for all the<br />

different sources.<br />

To get the distribution <strong>of</strong> the <strong>SiGe</strong> composition the MRD x-ray system is used, where a whole<br />

4”-wafer can be mounted <strong>and</strong> the measurement spots can be easily adjusted due to the wide<br />

accessible range <strong>of</strong> the motorized x-y-stage. At each position an ω-2θ scan can be recorded<br />

fully automated for later data evaluation.<br />

Fig. A.2 shows the evaluation <strong>of</strong> the x-ray data <strong>of</strong> Sample 1536LSG which was grown with<br />

the old Ge-evaporation-assembly, that means, before the small Ge source was replaced with<br />

the new larger electron gun for Ge. The comparison <strong>of</strong> the distribution <strong>of</strong> the normalized Ge<br />

Sample 1536LSG 1823LSG 1653LS n 1661LS p<br />

Buffer 700 ˚A Si 700 ˚A Si 840 ˚A Si 500 ˚A Si<br />

Si @ 1.5 ˚A/s Si @ 1.5 ˚A/s Si @ 0.7 ˚A/s Si @ 1.0 ˚A/s<br />

550 ◦ C → 500 ◦ C 550 ◦ C → 500 ◦ C 550 ◦ C 550 ◦ C<br />

+<br />

840 ˚A Si<br />

Si @ 0.7 ˚A/s<br />

550 ◦ C → 350 ◦ C<br />

Epilayer 1000 ˚A Si0.90Ge0.10 1000 ˚A Si0.90Ge0.10 300 ˚A Si:Sb 1000 ˚A Si:B<br />

Si @ 1.5 ˚A/s Si @ 1.5 ˚A/s Sb pre-dep.: 120 s B @ 1850 ◦ C<br />

Ge @ 0.1667 ˚A/s Ge @ 0.1667 ˚A/s Sb @ 310 ◦ C Si @ 1.0 ˚A/s<br />

500 ◦ C 500 ◦ C Si @ 0.7 ˚A/s 550 ◦ C<br />

350 ◦ C<br />

Table A.1: Typical structure <strong>and</strong> growth parameters <strong>of</strong> several calibration layers.


A.1. CALIBRATIONS 135<br />

Figure A.1: X-Ray data (ω-2θ scan) <strong>of</strong> a <strong>SiGe</strong>-epilayer grown for calibration <strong>of</strong> Si<br />

<strong>and</strong> Ge sources. Nominally, the epilayer consists <strong>of</strong> 1000 ˚A Si0.90Ge0.10. Using our<br />

MRD x-ray system, the whole 4”-wafer can be mounted <strong>and</strong> the measurement spots<br />

can be easily adjusted due to the wide accessible range <strong>of</strong> the motorized x-y-stage.<br />

The plotted graph shows x-ray data recorded at the wafer center (green solid line)<br />

together with the fitted, simulated curve (red dashed line).<br />

� Source: 1823LSG xpos0 ypos0 2.jpg<br />

flux between the old <strong>and</strong> the new Ge-evaporator shows, as suspected, a significant change.<br />

This can be seen from Fig. A.3, where the evaluated data from sample 1823LSG are visual-<br />

ized.<br />

To get reproducible results, or to compare different samples within a growth series, it is es-<br />

sential to be aware <strong>of</strong> the layer distribution. This is especially the case for growth on small<br />

substrate pieces using all-Si adapters. Whenever samples <strong>of</strong> consecutive growth processes<br />

should be directly compared, one has to take care to rotate them to the same growth posi-<br />

tion.<br />

The calibration data were completed with measuring the flux distribution <strong>of</strong> our Sb <strong>and</strong><br />

B effusion cells used for doping. These sources are calibrated with doped epilayers which are<br />

grown with direct doping for p-doped layers (B) or pre-deposition <strong>of</strong> Sb <strong>and</strong> subsequent Si


136 APPENDIX A. CALIBRATION AND CHARACTERIZATION OF SOURCES<br />

Figure A.2: Normalized Si <strong>and</strong> Ge distribution across a grown wafer (1536LSG)<br />

evaluated via x-ray measurements. The different measurement positions are<br />

marked with black spots. The black circle indicates the circumference <strong>of</strong> a 100 mm<br />

wafer. The layers were grown with the old Ge-evaporation-assembly.<br />

� Source: wafercalibration 1536Si fig01.jpg, wafercalibration 1536Ge fig02.jpg<br />

overgrowth with Sb incorporation for n-doped layers. The n-doped layers have to be grown at


A.1. CALIBRATIONS 137<br />

Figure A.3: Normalized Si-distribution on calibration layer 1823LSG grown<br />

with the new Ge-evaporation-assembly. Markers point up the side in the growth<br />

chamber from which the different sources emit.<br />

� Source: wafercalibration 1823Si fig05.jpg, wafercalibration 1823Ge fig06.jpg<br />

lowest possible substrate temperatures to suppress the Sb-segregation but still to guarantee<br />

good crystalline quality <strong>of</strong> the epilayers. More details <strong>of</strong> the layer structure can be found


138 APPENDIX A. CALIBRATION AND CHARACTERIZATION OF SOURCES<br />

Figure A.4: Sb <strong>and</strong> B flux distribution <strong>of</strong> calibration layer 1653LS n <strong>and</strong><br />

1661LS p, respectively. The plotted data are based on 4-point measurements which<br />

take into account only the active doping concentration which contributes to the<br />

electrical conductivity.<br />

� Source: wafercalibration 1653Sb fig10.jpg, wafercalibration 1661B fig14.jpg<br />

again in Tab. A.1.<br />

The epilayers are characterized electrically by measuring the sheet-resistance with a 4-point


A.2. PHOTOLUMINESCENCE 139<br />

Sample 1596LSG<br />

Buffer 1000 ˚A Si<br />

Si @ 1.0 ˚A/s; 750 ◦ C → 550 ◦ C<br />

+<br />

1000 ˚A Si<br />

Si @ 1.0 ˚A/s → 0.45 ˚A/s; 550 ◦ C → 450 ◦ C<br />

<strong>SiGe</strong> quantum-well 25 ˚A Si0.75Ge0.25<br />

Si @ 0.45 ˚A/s, Ge @ 0.15 ˚A/s; 450 ◦ C<br />

Cap 500 ˚A Si<br />

Si @ 0.45 ˚A/s → 1.0 ˚A/s; 450 ◦ C → 650 ◦ C<br />

+<br />

1500 ˚A Si<br />

Si @ 1.0 ˚A/s; 650 ◦ C<br />

Table A.2: Layer structure <strong>and</strong> growth parameters <strong>of</strong> a typical PL structure.<br />

measurement setup across the central region <strong>of</strong> a 4”-wafer. Therefore the calculated fluxes<br />

presented in Fig. A.4 for boron (1661LS p) <strong>and</strong> antimony (1653LS n) take into account only<br />

the active doping concentration which contributes to the electrical conductivity. SIMS mea-<br />

surements are known to give significantly different fluxes for the Sb source in particular.<br />

Nevertheless for electrical applications only the activated dopants are relevant <strong>and</strong> although<br />

the absolute values for the fluxes may vary, the distribution <strong>of</strong> the doping concentration<br />

remains unaffected.<br />

A.2 Photoluminescence<br />

As already mentioned in this chapter’s preamble, photoluminescence measurements (PL) on<br />

single <strong>SiGe</strong> quantum wells were performed to verify the cleanliness <strong>of</strong> our source material.<br />

The layer structure <strong>and</strong> growth parameters <strong>of</strong> a typical photoluminescence structure are listed<br />

in Tab. A.2. The PL data presented in this chapter were acquired together with our optics<br />

group. The samples were irradiated with an Ar + laser (λ = 514.5 nm) with a power <strong>of</strong> 58 mW<br />

<strong>and</strong> measured at a temperature <strong>of</strong> 4.2 K with an InGaAs CCD-line detector. Fig. A.5 shows<br />

a typical spectrum <strong>of</strong> a single Si0.75Ge0.25 quantum well with a well width <strong>of</strong> 25 ˚A grown with<br />

the new Ge-evaporation assembly. Sample 1596LSG shows well-behaved photoluminescence<br />

signal from which can be stated that the first crucible filling <strong>of</strong> our new Ge source is not<br />

contaminated. The comparison <strong>of</strong> the PL-signal <strong>of</strong> structure 1596LSG with the reference


140 APPENDIX A. CALIBRATION AND CHARACTERIZATION OF SOURCES<br />

Figure A.5: Photoluminescence spectrum <strong>of</strong> single Si0.75Ge0.25 quantum well<br />

with a well width <strong>of</strong> 25 ˚A grown with the new Ge-evaporation assembly (sample<br />

1596LSG). The labels indicate the different PL-signals.<br />

� Source: photolum fig09.jpg<br />

data <strong>of</strong> sample 1080MSG, which was grown long ago with the old small Ge evaporator,<br />

Figure A.6: Schematic drawing indicating various measurement positions used<br />

for recording photoluminescence spectra.<br />

� Source: PL measure points2.jpg


A.2. PHOTOLUMINESCENCE 141<br />

Figure A.7: Photoluminescence spectra taken at different wafer positions indicating<br />

a significant variation <strong>of</strong> the Si <strong>and</strong> Ge thickness. Origin <strong>of</strong> PL signals with<br />

regard to sample position can be seen in Fig. A.6.<br />

� Source: photolum fig08.jpg<br />

shows, that the samples match regarding PL intensity (see Fig. A.7). In principle the peaks<br />

<strong>of</strong> samples 1080MSG <strong>and</strong> 1596LSG at measurement spot ”D” (central wafer position) should<br />

coincide. The significant <strong>of</strong>fset between the positions <strong>of</strong> the <strong>SiGe</strong> no-phonon (NP) peak can<br />

be deduced from slight <strong>SiGe</strong> calibration deficiencies. Various PL measurements (Fig. A.6<br />

<strong>and</strong> Fig. A.7) were performed on sample 1596LSG across the wafer from the ”Si-rich” to<br />

the ”Ge-rich” region. The significant shift <strong>of</strong> the <strong>SiGe</strong> NP-peak can be explained by the<br />

distribution <strong>of</strong> Si <strong>and</strong> Ge <strong>and</strong> the resulting changes in Ge-content <strong>and</strong> layer thickness. For<br />

the relatively thin <strong>SiGe</strong> quantum well examined here the shift <strong>of</strong> the no-phonon line depends<br />

on quantum confinement <strong>and</strong> on the reduced <strong>SiGe</strong> energy gap (which is coupled to the Ge-<br />

content) in about equal parts. Obviously, either the quantum well thickness <strong>of</strong> the grown<br />

layer is gradually increasing from the ”Si-side” towards the ”Ge-side” or the the Ge-content<br />

is heavily increased. Both trends would red-shift the <strong>SiGe</strong> NP-line as a wide well yields low<br />

quantum confinement energy having the energy levels near the bottom, <strong>and</strong> a high Ge-content<br />

results in a lower excitonic b<strong>and</strong>gap. The Si <strong>and</strong> Ge flux distribution known from x-ray<br />

analyses <strong>of</strong> <strong>SiGe</strong> calibration layers reveal that the maximum thickness has to be expected<br />

rather at the ”Si-side”. Therefore, the observed red-shift has to be attributed to a noticeable


142 APPENDIX A. CALIBRATION AND CHARACTERIZATION OF SOURCES<br />

Ge gradient.<br />

Figure A.8: Calculated <strong>SiGe</strong> NP-line position as a function <strong>of</strong> Ge-content <strong>and</strong><br />

quantum well width. The computation includes the approximation <strong>of</strong> the <strong>SiGe</strong><br />

excitonic b<strong>and</strong>gap <strong>and</strong> the quantum confinement energy in an ideal rectangular<br />

quantum well only.<br />

� Source: NP <strong>SiGe</strong> position fig03c.jpg<br />

The position <strong>of</strong> the <strong>SiGe</strong> NP-line can be estimated by considering only the <strong>SiGe</strong> excitonic<br />

b<strong>and</strong>gap E <strong>SiGe</strong><br />

GX <strong>and</strong> the quantum confinement energy ∆Equant in the valence b<strong>and</strong> (VB)<br />

Eq. A.1.<br />

E <strong>SiGe</strong><br />

NP = E <strong>SiGe</strong><br />

GX + ∆Equant (A.1)<br />

Calculating the energy levels for a textbook-like rectangular box gives a good approximation<br />

for the confinement energy <strong>of</strong> the single <strong>SiGe</strong> quantum well. Only the out-<strong>of</strong>-plane heavy hole<br />

masses m⊥ hh have to be taken into account. The whole b<strong>and</strong>gap difference is assumed to occur<br />

in the VB <strong>and</strong> is used as barrier height <strong>of</strong> the quantum well Vb (see Wachter pp. 27 [242]).<br />

Values for the excitonic <strong>SiGe</strong> b<strong>and</strong>gap are taken from a fit to measured <strong>SiGe</strong> NP-line positions<br />

for thick <strong>SiGe</strong> quantum wells <strong>of</strong> known Ge-content x, which show no significant quantization<br />

effect Eq. A.2<br />

EGX(x) = 1.155 − (0.848 ± 0.029)x + (0.173 ± 0.108)x 2<br />

[eV] (A.2)


A.2. PHOTOLUMINESCENCE 143<br />

Figure A.9: Zoom-in into Fig. A.8 for better comparability <strong>of</strong> the measured <strong>SiGe</strong><br />

NP-lines presented in Fig A.5. The calculated data (well width Lz = 2.5 nm, Gecontent<br />

x = 0.25) match the measured PL-data <strong>of</strong> 1596LSG B (central wafer position)<br />

nicely. This indicates an accurate <strong>SiGe</strong> calibration.<br />

� Source: NP <strong>SiGe</strong> position fig04c.jpg<br />

(after Wachter pp. 53 [242]).<br />

In Fig. A.8 the thus calculated <strong>SiGe</strong> NP-line position as a function <strong>of</strong> Ge-content <strong>and</strong> quan-<br />

tum well width is depicted. It is clearly visible that the <strong>SiGe</strong> NP-line position is dominated<br />

by the Ge-content for broad wells which give no significant confinement energy contribution,<br />

whereas for thin quantum wells the Ge-content plays a subordinate role. Fig. A.9 shows a<br />

zoom-in into Fig. A.8 for better comparability with the measured <strong>SiGe</strong> NP-lines presented<br />

in Fig. A.5. The calculated data (well width Lz = 2.5 nm, Ge-content x = 0.25) match nicely<br />

PL measurement position 1596LSG B 1596LSG D 1596LSG A<br />

recalculated well thickness Lz [˚A] 25.59 25.00 23.43<br />

recalculated Ge-content x 0.225 0.250 0.275<br />

measured <strong>SiGe</strong> NP-energy E<strong>SiGe</strong> NP, meas [meV] 1039.0 1025.5 1015.0<br />

calculated <strong>SiGe</strong> NP-energy E<strong>SiGe</strong> NP, calc [meV] 1041.0 1025.0 1016.0<br />

Table A.3: Comparison <strong>of</strong> the measured <strong>and</strong> calculated <strong>SiGe</strong> NP-lines E <strong>SiGe</strong><br />

NP .


144 APPENDIX A. CALIBRATION AND CHARACTERIZATION OF SOURCES<br />

x m ⊥ hh Vb [meV]<br />

0.00 0.2778 0.0<br />

0.05 0.2729 38.6<br />

0.10 0.2681 77.4<br />

0.15 0.2635 116.1<br />

0.20 0.2591 155.0<br />

0.25 0.2548 193.9<br />

0.30 0.2506 232.9<br />

0.35 0.2466 272.0<br />

Table A.4: Values used for quantum confinement energy calculation ∆Equant.<br />

with the measured PL-data <strong>of</strong> 1596LSG D (central wafer position), as the calculated <strong>SiGe</strong><br />

NP-transition energy E<strong>SiGe</strong> NP, calc ≈ 1025 meV is quite close to the measured value, which reads<br />

E <strong>SiGe</strong><br />

NP, meas<br />

≈ 1025.5 meV. Additionally the experimentally found <strong>SiGe</strong> NP-line positions from<br />

the <strong>of</strong>f-center measurements (1596LSG A <strong>and</strong> 1596LSG B) can be reproduced with the cal-<br />

culations. By applying corrections for the quantum well thickness <strong>and</strong> Ge-content according<br />

to the distribution <strong>of</strong> the Si <strong>and</strong> Ge source (known from x-ray calibration, Sec. A.1) the dis-<br />

crepancy between measured <strong>and</strong> calculated data is negligible (see Tab. A.3). This actually<br />

proves the accurate <strong>SiGe</strong> calibration <strong>of</strong> our MBE-system.<br />

The calculation is partly based on a Matlab-program developed by T. Fromherz [243]. For<br />

more details on that program <strong>and</strong> parameters involved see also the diploma thesis written by<br />

P. Rauter (pp. 17) [244] <strong>and</strong> references therein [245, 246, 247]. Several values computed with<br />

the aforementioned program <strong>and</strong> used for the calculation <strong>of</strong> the quantum confinement energy<br />

∆Equant, namely the perpendicular heavy hole mass m ⊥ hh <strong>and</strong> the barrier height Vb for the<br />

<strong>SiGe</strong> quantum well as function <strong>of</strong> Ge-content x, can be found in Tab. A.4.


Appendix B<br />

General Physical Data<br />

B.1 Stereographic Projection<br />

Figure B.1: Stereographic projection <strong>of</strong> important Si-crystal planes relative to the<br />

(001)-surface [12].<br />

� Source: Stereographic Proj.jpg<br />

145


146 APPENDIX B. GENERAL PHYSICAL DATA<br />

B.2 Physical Constants<br />

International System <strong>of</strong> Units (SI) From: physics.nist.gov/constants<br />

Fundamental Physical Constants – Frequently used constants<br />

Quantity Symbol Value Unit<br />

speed <strong>of</strong> light in vacuum c, c0 299792458 m s−1 magnetic constant µ0 4π × 10−7 N A−2 = 12.566370614... × 10−7 N A−2 electric constant 1/µ0c2 ε0 8.854187817... × 10−12 F m−1 Newtonian constant<br />

<strong>of</strong> gravitation G 6.673(10) × 10−11 m3 kg−1 s−2 Planck constant h 6.62606876(52) × 10 −34 J s<br />

h/2π � 1.054571596(82) × 10 −34 J s<br />

elementary charge e 1.602176462(63) × 10 −19 C<br />

magnetic flux quantum h/2e Φ0 2.067833636(81) × 10 −15 Wb<br />

conductance quantum 2e 2 /h G0 7.748091696(28) × 10 −5 S<br />

electron mass me 9.10938188(72) × 10 −31 kg<br />

proton mass mp 1.67262158(13) × 10 −27 kg<br />

proton-electron mass ratio mp/me 1836.1526675(39)<br />

fine-structure constant e 2 /4πε0�c α 7.297352533(27) × 10 −3<br />

inverse fine-structure constant α −1 137.03599976(50)<br />

Rydberg constant α 2 mec/2h R∞ 10973731.568549(83) m −1<br />

Avogadro constant NA, L 6.02214199(47) × 10 23 mol −1<br />

Faraday constant NAe F 96485.3415(39) C mol −1<br />

molar gas constant R 8.314472(15) J mol −1 K −1<br />

Boltzmann constant R/NA k 1.3806503(24) × 10 −23 J K−1<br />

Stefan-Boltzmann constant<br />

(π 2 /60)k 4 /� 3 c 2 σ 5.670400(40) × 10 −8 W m −2 K −4<br />

Non-SI units accepted for use with the SI<br />

electron volt: (e/C)J eV 1.602176462(63) × 10 −19 J<br />

(unified) atomic mass unit<br />

1 u = mu = 1<br />

12 m(12 C) u 1.66053873(13) × 10 −27 kg<br />

= 10 −3 kg mol −1 /NA<br />

Source: Peter J. Mohr <strong>and</strong> Barry N. Taylor, CODATA Recommended Values <strong>of</strong> the Fundamental<br />

Physical Constants: 1998, Journal <strong>of</strong> Physical <strong>and</strong> Chemical Reference Data, Vol. 28, No. 6, 1999<br />

<strong>and</strong> Reviews <strong>of</strong> Modern Physics, Vol. 72, No. 2, 2000.


Bibliography<br />

[1] Herbert Lichtenberger, Characterization <strong>and</strong> Overgrowth <strong>of</strong> Prestructured Silicon-Substrates,<br />

Diploma-thesis, Institut für Halbleiterphysik, Johannes Kepler Universität Linz, Juni 2002.<br />

[2] S. M. Sze, Semiconductor Devices, John Wiley & Sons, 1985.<br />

[3] O. Madelung (ed.), Semiconductors – Group IV <strong>and</strong> III-V Compounds, Data in Science <strong>and</strong><br />

Technology, Springer, 1991.<br />

[4] Robert Hull (ed.), Properties <strong>of</strong> Crystalline Silicon, no. 20, INSPEC, 1999.<br />

[5] Erich Kasper (ed.), Properties <strong>of</strong> <strong>Strain</strong>ed <strong>and</strong> Relaxed Silicon-Germanium, no. 12, INSPEC,<br />

1995.<br />

[6] F. Schäffler, Semicond. Sci. Technol. 12 (1997), 1515–1549.<br />

[7] Y. Shiraki <strong>and</strong> A. Sakai, Surface Science Reports 59 (2005), 153–207.<br />

[8] V. Zielasek, F. Liu, <strong>and</strong> M. G. Lagally, Structure <strong>of</strong> clean silicon surfaces: vicinal Si(001) <strong>and</strong><br />

Si(111) surfaces, Properties <strong>of</strong> Crystalline Silicon (Robert Hull, ed.), no. 20, INSPEC, 1999,<br />

pp. 183–188.<br />

[9] D. J. Chadi, Phys. Rev. Lett. 59 (1987), 1691–1694.<br />

[10] P. E. Wierenga, J. A. Kubby, <strong>and</strong> J. E. Griffith, Phys. Rev. Lett. 59 (1987), 2169–2172.<br />

[11] B. S. Swartzentruber, Y.-W. Mo, R. Kariotis, M. G. Lagally, <strong>and</strong> M. B. Webb, Phys. Rev. Lett.<br />

65 (1990), 1913–1916.<br />

[12] Christoph Schelling, Growth <strong>and</strong> Characterization <strong>of</strong> <strong>Self</strong>-Organized <strong>and</strong> ”Organized” Si <strong>and</strong><br />

Si1−xGex Nanostructures, Ph.D. thesis, Institut für Halbleiterphysik, Johannes Kepler Univer-<br />

sität Linz, November 2000.<br />

[13] O. L. Alerh<strong>and</strong>, D. V<strong>and</strong>erbilt, R. D. Meade, <strong>and</strong> J. D. Joannopoulos, Phys. Rev. Lett. 61<br />

(1988), 1973–1976.<br />

[14] F. K. Men, W. E. Packard, <strong>and</strong> M. B. Webb, Phys. Rev. Lett. 61 (1988), 2469–2471.<br />

[15] J. Ters<strong>of</strong>f <strong>and</strong> E. Pehlke, Phys. Rev. Lett. 68 (1992), 816–819.<br />

[16] R. M. Tromp <strong>and</strong> M. C. Reuter, Phys. Rev. Lett. 68 (1992), 820–822.<br />

147


148 BIBLIOGRAPHY<br />

[17] R. M. Tromp <strong>and</strong> M. C. Reuter, Phys. Rev. B 47 (1993), 7598–7601.<br />

[18] O. L. Alerh<strong>and</strong>, A. N. Berker, J. D. Joannopoulos, D. V<strong>and</strong>erbilt, R. J. Hamers, <strong>and</strong> J. E.<br />

Demuth, Phys. Rev. Lett. 64 (1990), 2406–2409.<br />

[19] B. S. Swartzentruber, N. Kitamura, M. G. Lagally, <strong>and</strong> M. B. Webb, Phys. Rev. B 47 (1993),<br />

13432–13441.<br />

[20] J. R. Chelikowsky <strong>and</strong> M. L. Cohen, Phys. Rev. B 14 (1976), 556–582.<br />

[21] Mark Winter (The University <strong>of</strong> Sheffield <strong>and</strong> WebElements Ltd, UK), WebElements - The<br />

Periodic Table on the WWW, http://www.webelements.com.<br />

[22] J. E. Field (ed.), The Properties <strong>of</strong> Diamond, Academic Press, 1979.<br />

[23] J. Hennis, J. Res. Bur. Nat. St<strong>and</strong>. 68A (1964), 529.<br />

[24] M. E. Straumanis <strong>and</strong> A. Z. Aka, J. Appl. Phys. 23 (1952), 330–334.<br />

[25] P. Becker, P. Seyfried, <strong>and</strong> H. Siegert, Z. Physik B 48 (1982), 17–21.<br />

[26] J. F. C. Baker <strong>and</strong> M. Hart, Acta Crystallogr. 31a (1975), 364–367.<br />

[27] W. Kaiser <strong>and</strong> W. L. Bond, Phys. Rev. 115 (1959), 857–863.<br />

[28] W. Bludau, A. Onton, <strong>and</strong> W. Heinke, J. Appl. Phys. 45 (1974), 1846–1848.<br />

[29] G. G. MacFarlane, T. P. McLean, J. E. Quarrington, <strong>and</strong> V. Roberts, Phys. Rev. 108 (1957),<br />

1377–1383.<br />

[30] F. J. Himpsel, J. A. Knapp, J. A. van Vechten, <strong>and</strong> D. E. Eastman, Phys. Rev. B 20 (1979),<br />

624–627.<br />

[31] H. Armon <strong>and</strong> J. P. F. Sellschop, Phys. Rev. B 26 (1982), 3289–3296.<br />

[32] D. R. Ma˘sović, F. R. Vukajlović, <strong>and</strong> S. Zeković, J. Phys. C 16 (1983), 6731–6738.<br />

[33] S. Zwerdling, B. Lax, L. M. Roth, <strong>and</strong> K. J. Button, Phys. Rev. 114 (1959), 80–89.<br />

[34] F. Nava, C. Canali, C. Jacoboni, L. Reggiani, <strong>and</strong> S. F. Kozlov, Solid State Commun. 33 (1980),<br />

475–477.<br />

[35] J. C. Hensel, H. Hasegawa, <strong>and</strong> M. Nakayama, Phys. Rev. 138 (1965), A225–A238.<br />

[36] D. Fink <strong>and</strong> R. Braunstein, Phys. Status Solidi (b) 73 (1976), 361–370.<br />

[37] L. Reggiani, D. Waechter, <strong>and</strong> S. Zukotynski, Phys. Rev. B 28 (1983), 3550–3555.<br />

[38] H. D. Barber, Solid-State Electronics 10 (1967), 1039–1051.<br />

[39] R. N. Dexter, H. J. Zeiger, <strong>and</strong> B. Lax, Phys. Rev. 104 (1956), 637–644.<br />

[40] R. L. Aggarwal, Phys. Rev. B 2 (1970), 446–458.<br />

[41] F. J. Morin <strong>and</strong> J. P. Maita, Phys. Rev. 96 (1954), 28–35.


BIBLIOGRAPHY 149<br />

[42] M. B. Prince, Phys. Rev. 92 (1953), 681–687.<br />

[43] F. Szmulowicz <strong>and</strong> F. L. Madarasz, Phys. Rev. B 27 (1983), 2605–2608.<br />

[44] F. van der Maesen <strong>and</strong> J. A. Brenkman, Philips. Res. Rep. 9 (1954), 225.<br />

[45] L. Reggiani, S. Bosi, C. Canali, F. Nava, <strong>and</strong> S. F. Kozlov, Phys. Rev. B 23 (1981), 3050–3057.<br />

[46] R. Hultgren, P. D. Desai, D. T. Hawkins, M. Gleiser, K. K. Kelly, <strong>and</strong> D. D. Wagman, The<br />

Thermodynamic Properties <strong>of</strong> the Elements, American Society for Metals, 1973.<br />

[47] F. P. Bundy, J. Chem. Phys. 38 (1963), 631–643.<br />

[48] W. Fulkerson, J. P. Moore, R. K. Williams, R. S. Graves, <strong>and</strong> D. L. McElroy, Phys. Rev. 167<br />

(1968), 765–782.<br />

[49] S. R. Bakchieva, N. P. Kekelidse, <strong>and</strong> M. G. Kekua, Phys. Status Solidi (a) 83 (1984), 139–145.<br />

[50] R. Berman <strong>and</strong> M. Martinez, Diamond Research 1976 (Suppl. Ind. Diamond Rev.) 7.<br />

[51] J. P. Dismukes, L. Ekstrom, <strong>and</strong> R. J. Paff, J. Phys. Chem. 68 (1964), 3021–3027.<br />

[52] D. V. Lang, R. People, J. C. Bean, <strong>and</strong> A. M. Sergent, Appl. Phys. Lett. 47 (1985), 1333–1335.<br />

[53] Michael Mühlberger, Modulationsdotierte Si/Si1−xGex– und Si/Si1−yCy–Heterostrukturen,<br />

Diploma-thesis, Institut für Halbleiterphysik, Johannes Kepler Universität Linz, August 1998.<br />

[54] Michael Mühlberger, Silicon-based Heterostructues: Growth, Modulation-Doping <strong>and</strong> Spin Prop-<br />

erties, Ph.D. thesis, Institut für Halbleiterphysik, Johannes Kepler Universität Linz, 2003.<br />

[55] M. A. Herman <strong>and</strong> H. Sitter, Molecular Beam Epitaxy – Fundamentals <strong>and</strong> Current Status, 2nd<br />

ed., Springer, 1996.<br />

[56] Robin F. C. Farrow, Molecular Beam Epitaxy – Applications to Key Materials, Noyes Publica-<br />

tions, 1995.<br />

[57] M. A. Herman, Vacuum 32 (1982), 555–565.<br />

[58] G. J. Davies <strong>and</strong> D. Williams, III-V MBE growth systems, The Technology <strong>and</strong> Physics in<br />

Molecular Beam Epitaxy (E. H. C. Parker, ed.), Plenum, 1985, p. 15.<br />

[59] K. Ploog, Molecular beam epitaxy <strong>of</strong> III-V compounds, Crystal Growth, Properties <strong>and</strong> Appli-<br />

cations (H. C. Freyhardt, ed.), vol. 3, Springer, 1980, p. 73.<br />

[60] L. D. Schmidt, Condensation <strong>Kinetic</strong>s <strong>and</strong> Mechanisms, The Physical Basis for Hetergeneous<br />

Catalysis (E. Drauglis <strong>and</strong> R. I. Jaffee, eds.), Plenum, 1975, p. 451.<br />

[61] I. G. Pitt, R. G. Gilbert, <strong>and</strong> K. R. Ryan, Surf. Sci. 324 (1995), 69–89.<br />

[62] P. Kisliuk, J. Phys. Chem. Solids 3 (1957), 95–101.<br />

[63] P. Kisliuk, J. Phys. Chem. Solids 5 (1958), 78–84.<br />

[64] M. A. Herman, Semiconductor Superlattices, Akademie-Verlag, 1986.


150 BIBLIOGRAPHY<br />

[65] J. A. Venables, G. D. T. Spiller, <strong>and</strong> M. Hanbuecken, Rep. Prog. Phys. 47 (1984), 399–459.<br />

[66] E. Bauer, Z. Kristallogr. 110 (1958), 372.<br />

[67] J. H. Van der Merve, Surf. Sci. 31 (1972), 198–228.<br />

[68] J. W. Matthews <strong>and</strong> A. E. Blakeslee, J. Cryst. Growth 27 (1974), 118–125.<br />

[69] J. C. Bean, L. C. Feldman, A. T. Fiory, S. Nakahara, <strong>and</strong> I. K. Robinson, J. Vac. Sci. Technol.<br />

A 2 (1984), 436–440.<br />

[70] H.-J. Herzog, H. Jorke, E. Kasper, <strong>and</strong> S. Mantl, J. Electrochem. Soc. 136 (1989), 3026.<br />

[71] W. Kern <strong>and</strong> D. A. Puotinen, RCA Rev. 31 (1970), 1887–1892.<br />

[72] W. Kern, J. Electrochem. Soc. 137 (1990), 1887–1892.<br />

[73] D. B. Williams <strong>and</strong> C. B. Carter, Transmission Electron Microscopy – A Textbook for Material<br />

Science, Plenum, 1996.<br />

[74] ThermoMicroscopes Corp., 1171 Borregas Avenue, Sunnyvale, CA 94089, A Practical Guide to<br />

Scanning Probe Microscopes.<br />

[75] L. Reimer, Scanning Electron Microscopy – Physics <strong>of</strong> Image Formation <strong>and</strong> Microanalysis, 2nd<br />

ed., Springer, 1998.<br />

[76] L. Reimer, Transmission Electron Microscopy – Physics <strong>of</strong> Image Formation <strong>and</strong> Microanalysis,<br />

4th ed., Springer, 1997.<br />

[77] JEOL LTD., 1-2 Musashino 3-chome Akishima Tokyo 196-8558 Japan, JEM-2010, 1999.<br />

[78] Thomas Berer, Electronic <strong>and</strong> spin properties <strong>of</strong> Si/<strong>SiGe</strong> heterostructures, Ph.D. thesis, Institut<br />

für Halbleiterphysik, Johannes Kepler Universität Linz, 2006.<br />

[79] Karin Wiesauer, Nanolithographie von Halbleitern durch mechanische Modifikation von Pho-<br />

toresist mit dem Atomkraftmikroskop, Diploma-thesis, Institut für Halbleiterphysik, Johannes<br />

Kepler Universität Linz, Oktober 1998.<br />

[80] R. Wiesendanger, Scanning Probe Microscopy <strong>and</strong> Spectroscopy, Cambridge University Press,<br />

1994.<br />

[81] Park Scientific Instruments, 1171 Borregas Avenue, Sunnyvale, California 94089, Operating<br />

Instructions: Non-Contact AFM, Intermittent-Contact AFM <strong>and</strong> MFM, May 5 1995.<br />

[82] Heinz Seyringer, Nanostrukturierung und Charakterisierung von Si/<strong>SiGe</strong>-Heterostrukturen,<br />

Ph.D. thesis, Institut für Halbleiterphysik, Johannes Kepler Universität Linz, März 2000.<br />

[83] Digital Instruments Veeco Metrology Group, 112 Robin Hill Road, Santa Barbara, CA 93117,<br />

Dimension� 3100 Manual, ver 4.43b ed., June 1 2000.<br />

[84] Olympus MicroCantilever – OMCL Series,<br />

http://www.olympus.co.jp/en/insg/probe/img/catalog05E.pdf.


BIBLIOGRAPHY 151<br />

[85] J. Zhu, K. Brunner, <strong>and</strong> G. Abstreiter, Appl. Phys. Lett. 73 (1998), 620–622.<br />

[86] K. Sakamoto, H. Matsuhata, M. O. Tanner, D. Wang, <strong>and</strong> K. L. Wang, Thin Solid Films 321<br />

(1998), 55–59.<br />

[87] M. Abdallah, I. Berbezier, P. Dawson, M. Serpentini, G. Bremond, <strong>and</strong> B. Joyce, Thin Solid<br />

Films 336 (1998), 256–261.<br />

[88] C. Teichert, Physics Reports 365 (2002), 335–432.<br />

[89] H. Lichtenberger, M. Mühlberger, C. Schelling, <strong>and</strong> F. Schäffler, J. Cryst. Growth 278 (2005),<br />

78–82.<br />

[90] H. Lichtenberger, M. Mühlberger, <strong>and</strong> F. Schäffler, Appl. Phys. Lett. 86 (2005), 131919.<br />

[91] G. Chen, H. Lichtenberger, F. Schäffler, G. Bauer, <strong>and</strong> W. Jantsch, Mat. Sci. Eng. C 26 (2006),<br />

795–799.<br />

[92] Z. Zhong, H. Lichtenberger, G. Chen, M. Mühlberger, C. Schelling, J. Mysliveĉek, A. Halilovic,<br />

J. Stangl, G. Bauer, W. Jantsch, <strong>and</strong> F. Schäffler, Microelectr. Eng. 83 (2006), 1730–1735.<br />

[93] Z. Zhong, O. G. Schmidt, <strong>and</strong> G. Bauer, Appl. Phys. Lett. 87 (2005), 133111.<br />

[94] Z. Zhong, A. Halilovic, M. Mühlberger, F. Schäffler, <strong>and</strong> G. Bauer, Appl. Phys. Lett. 82 (2003),<br />

445–447.<br />

[95] Z. Zhong, A. Halilovic, H. Lichtenberger, F. Schäffler, <strong>and</strong> G. Bauer, Physica E 23 (2004),<br />

243–247.<br />

[96] M. G. Lagally, Jpn. J. Appl. Phys. 32 (1993), 1493–1501.<br />

[97] G. Ehrlich <strong>and</strong> F. G. Hudda, J. Chem. Phys. 44 (1966), 1039–1049.<br />

[98] R. L. Schwoebel <strong>and</strong> E. J. Shipsey, J. Appl. Phys. 37 (1966), 3682–3686.<br />

[99] R. L. Schwoebel, J. Appl. Phys. 40 (1969), 614–618.<br />

[100] G. S. Bales <strong>and</strong> A. Zangwill, Phys. Rev. B 41 (1990), 5500–5508.<br />

[101] J. Krug, M. Plischke, <strong>and</strong> M. Siegert, Phys. Rev. Lett. 70 (1993), 3271–3274.<br />

[102] J. Krug, Adv. Phys. 46 (1997), 139.<br />

[103] J. Zhu, K. Brunner, <strong>and</strong> G. Abstreiter, Appl. Phys. Lett. 73 (1998), 2438–2440.<br />

[104] I. Berbezier, A. Ronda, A. Portavoce, <strong>and</strong> N. Motta, Appl. Phys. Lett. 83 (2003), 4833–4835.<br />

[105] A. Ronda <strong>and</strong> I. Berbezier, Physica E 23 (2004), 370–376.<br />

[106] C. Schelling, G. Springholz, <strong>and</strong> F. Schäffler, Phys. Rev. Lett. 83 (1999), 995–998.<br />

[107] C. Schelling, G. Springholz, <strong>and</strong> F. Schäffler, Thin Solid Films 369 (2000), 1–4.<br />

[108] C. Schelling, M. Mühlberger, G. Springholz, <strong>and</strong> F. Schäffler, Phys. Rev. B 64 (2001), 041301(R).


152 BIBLIOGRAPHY<br />

[109] M. Mühlberger, C. Schelling, G. Springholz, <strong>and</strong> F. Schäffler, Mat. Sci. Eng. B 89 (2002),<br />

257–262.<br />

[110] M. Mühlberger, C. Schelling, G. Springholz, <strong>and</strong> F. Schäffler, Physica E 13 (2002), 990–994.<br />

[111] M. Mühlberger, C. Schelling, G. Springholz, <strong>and</strong> F. Schäffler, Surf. Sci. 532–535 (2003), 721–<br />

726.<br />

[112] J. Ters<strong>of</strong>f, Y. H. Phang, Z. Zhang, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 75 (1995), 2730–2733.<br />

[113] F. Liu, J. Ters<strong>of</strong>f, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 80 (1998), 1268–1271.<br />

[114] Y. H. Phang, C. Teichert, M. G. Lagally, L. J. Peticolos, J. C. Bean, <strong>and</strong> E. Kasper, Phys. Rev.<br />

B 50 (1994), 14435–14445.<br />

[115] C. Teichert, J. C. Bean, <strong>and</strong> M. G. Lagally, Appl. Phys. A 67 (1998), 675–685.<br />

[116] C. Teichert, Y. H. Phang, L. J. Peticolas, J. C. Bean, <strong>and</strong> J. Ters<strong>of</strong>f, Surface Diffusion: Atomistic<br />

<strong>and</strong> Collective Processes, Stress-driven morphological changes <strong>of</strong> <strong>SiGe</strong> films grown on vicinal<br />

Si(001) substrates (M. C. Tringides, ed.), Plenum Press, New York, 1997, pp. 297–307.<br />

[117] P. Politi <strong>and</strong> J. Krug, Surf. Sci. 446 (2000), 89–97.<br />

[118] J. Mysliveček, C. Schelling, F. Schäffler, G. Springholz, P. ˇ Smilauer, J. Krug, <strong>and</strong> B. Voigtländer,<br />

cond–mat/0212331.<br />

[119] J. Mysliveček, C. Schelling, G. Springholz, F. Schäffler, B. Voigtländer, <strong>and</strong> P. ˇ Smilauer, Mat.<br />

Sci. Eng. B 89 (2002), 410–414.<br />

[120] J. Mysliveček, C. Schelling, F. Schäffler, G. Springholz, P. ˇ Smilauer, J. Krug, <strong>and</strong> B. Voigtländer,<br />

Surf. Sci. 520 (2002), 193–206.<br />

[121] T. Frisch <strong>and</strong> A. Verga, Phys. Rev. Lett. 94 (2005), 226102.<br />

[122] F. Slanina, J. Krug, <strong>and</strong> M. Kotrla, Phys. Rev. E 71 (2005), 041605.<br />

[123] A. Pimpinelli, V. Tonchev, A. Videcoq, <strong>and</strong> M. Vladimirova, Phys. Rev. Lett. 88 (2002), 206103.<br />

[124] S. Fukatsu, K. Fujita, H. Yaguchi, Y. Shiraki, <strong>and</strong> R. Ito, Appl. Phys. Lett. 59 (1991), 2103–<br />

2105.<br />

[125] K. Fujita, S. Fukatsu, H. Yaguchi, Y. Shiraki, <strong>and</strong> R. Ito, Appl. Phys. Lett. 59 (1991), 2240–<br />

2241.<br />

[126] D. J. Godbey, J. V. Lill, J. Deppe, <strong>and</strong> K. D. Hobart, Appl. Phys. Lett. 65 (1994), 711–713.<br />

[127] F. Wu, X. Chen, Z. Zhang, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 74 (1995), 574–577.<br />

[128] B. Voigtländer <strong>and</strong> M. Kästner, Phys. Rev. B 60 (1999), R5121–R5124.<br />

[129] A. I. Yakimov, A. V. Dvurechenskii, V. V. Kirienko, Y. I. Yakovlev, A. I. Nikiforov, <strong>and</strong> C. J.<br />

Adkins, Phys. Rev. B 61 (2000), 10868–10876.


BIBLIOGRAPHY 153<br />

[130] C. S. Peng, Q. Huang, W. Q. Cheng, J. M. Zhou, Y. H. Zhang, T. T. Sheng, <strong>and</strong> C. H. Tung,<br />

Phys. Rev. B 57 (1998), 8805–8808.<br />

[131] Y. Shiraki, H. Sunamura, N. Usami, <strong>and</strong> S. Fukatsu, Appl. Surf. Sci. 102 (1996), 263–271.<br />

[132] J.-M. Baribeau, X. Wu, N. L. Rowell, <strong>and</strong> D. J. Lockwood, J. Phys.: Condens. Matter 18<br />

(2006), R139–R174.<br />

[133] O. G. Schmidt, K. Eberl, <strong>and</strong> Y. Rau, Phys. Rev. B 62 (2000), 16715–16720.<br />

[134] A. Karmous, A. Cuenat, A. Ronda, I. Berbezier, S. Atha, <strong>and</strong> R. Hull, Appl. Phys. Lett. 85<br />

(2004), 6401–6403.<br />

[135] G. Jin, J. L. Liu, <strong>and</strong> K. L. Wang, Appl. Phys. Lett. 83 (2003), 2847–2849.<br />

[136] G. Springholz, V. Holy, M. Pinczolits, <strong>and</strong> G. Bauer, Science 282 (1998), 734–737.<br />

[137] J. Ters<strong>of</strong>f, C. Teichert, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 76 (1996), 1675–1678.<br />

[138] J. Brunner, P. Schittenhelm, J. Gondermann, B. Spangenberg, B. Hadam, T. Köster, H. G.<br />

Roskos, H. Kurz, H. Gossner, I. Eisele, <strong>and</strong> G. Abstreiter, J. Cryst. Growth 150 (1995), 1060–<br />

1064.<br />

[139] E. S. Kim, N. Usami, <strong>and</strong> Y. Shiraki, Appl. Phys. Lett. 72 (1998), 1617–1619.<br />

[140] Y. Nitta, M. Shibata, K. Fujita, <strong>and</strong> M. Ichikawa, Surf. Sci. 462 (2000), L587–L593.<br />

[141] H. Omi, D. J. Bottomley, Y. Homma, <strong>and</strong> T. Ogino, Phys. Rev. B 67 (2003), 115302.<br />

[142] M. Borgström, V. Zela, <strong>and</strong> W. Seifert, Nanotechnology 14 (2003), 264–267.<br />

[143] F. Leroy, J. Eymery, P. Gentile, <strong>and</strong> F. Fournel, Appl. Phys. Lett. 80 (2002), 3078–3080.<br />

[144] G. Jin, J. L. Liu, S. G. Thomas, Y. H. Luo, K. L. Wang, <strong>and</strong> B.-Y. Nguyen, Appl. Phys. Lett.<br />

75 (1999), 2752–2754.<br />

[145] L. Vescan, K. Grimm, M. Goryll, <strong>and</strong> B. Holländer, Mat. Sci. Eng. B 69–70 (2000), 324–328.<br />

[146] L. Vescan <strong>and</strong> T. Stoica, J. Appl. Phys. 91 (2002), 10119–10126.<br />

[147] T. Kitajima, B. Liu, <strong>and</strong> S. R. Leone, Appl. Phys. Lett. 80 (2002), 497–499.<br />

[148] Z. Zhong <strong>and</strong> G. Bauer, Appl. Phys. Lett. 84 (2004), 1922–1924.<br />

[149] B. Yang, F. Liu, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 92 (2004), 025502.<br />

[150] T. I. Kamins, D. A.A. Ohlberg, R. S. Williams, W. Zhang, <strong>and</strong> S. Y. Chou, Appl. Phys. Lett.<br />

74 (1999), 1773–1775.<br />

[151] Paul Scherrer Institute (PSI), Laboratory for Micro- <strong>and</strong> Nanotechnology: XIL, X-ray interfer-<br />

ence lithography facility at the Swiss Light Source (SLS), http://lmn.web.psi.ch/xil/intro.htm.<br />

[152] M.-I. Richard, T.-U. Schülli, E. Wintersberger, G. Renaud, <strong>and</strong> G. Bauer, NIM B 246 (2006),<br />

35–38.


154 BIBLIOGRAPHY<br />

[153] M. Kästner <strong>and</strong> B. Voigtländer, Phys. Rev. Lett. 82 (1999), 2745–2748.<br />

[154] A. Vailionis, B. Cho, G. Glass, P. Desjardins, D. G. Cahill, <strong>and</strong> J. E. Greene, Phys. Rev. Lett.<br />

85 (2000), 3672–3675.<br />

[155] Y.-W. Mo, D. E. Savage, B. S. Schwartzentruber, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 65 (1990),<br />

1020–1023.<br />

[156] F. M. Ross, R. M. Tromp, <strong>and</strong> M. C. Reuter, Science 286 (1999), 1931–1934.<br />

[157] A. Ronda, I. Berbezier, A. Pascale, A. Portavoce, <strong>and</strong> F. Volpi, Mat. Sci. Eng. B 101 (2003),<br />

95–101.<br />

[158] J. Zhu, K. Brunner, <strong>and</strong> G. Abstreiter, Appl. Phys. Lett. 72 (1998), 424–426.<br />

[159] F. Watanabe, D. G. Cahill, S. Hong, <strong>and</strong> J. E. Greene, Appl. Phys. Lett. 85 (2004), 1238–1240.<br />

[160] F. Watanabe, D. G. Cahill, <strong>and</strong> J. E. Greene, Phys. Rev. Lett. 94 (2005), 066101.<br />

[161] F. Liu, S. E. Davenport, H. M. Evans, <strong>and</strong> M. G. Lagally, Phys. Rev. Lett. 82 (1999), 2528–2531.<br />

[162] H. Omi <strong>and</strong> T. Ogino, Phys. Rev. B 59 (1999), 7521–7527.<br />

[163] K. Brunner, J. Zhu, C. Miesner, G. Abstreiter, O. Kienzle, <strong>and</strong> F. Ernst, Physica E 7 (2000),<br />

881–886.<br />

[164] K. Brunner, J. Zhu, G. Abstreiter, O. Kienzle, <strong>and</strong> F. Ernst, Thin Solid Films 369 (2000),<br />

39–42.<br />

[165] J. Qin, F. Xue, Y. Wang, L. H. Bai, J. Cui, X. J. Yang, Y. L. Fan, <strong>and</strong> Z. M. Jiang, J. Cryst.<br />

Growth 278 (2005), 136–141.<br />

[166] D. J. Eaglesham, F. C. Unterwald, <strong>and</strong> D. C. Jacobson, Phys. Rev. Lett. 70 (1993), 966–969.<br />

[167] A. Portavoce, A. Ronda, <strong>and</strong> I. Berbezier, Mat. Sci. Eng. B 89 (2002), 205–210.<br />

[168] B. P. Tinkham, D. M. Goodner, D. A. Walko, <strong>and</strong> M. J. Bedzyk, Phys. Rev. B 67 (2003),<br />

035404.<br />

[169] O. G. Schmidt, C. Lange, K. Eberl, O. Kienzle, <strong>and</strong> F. Ernst, Appl. Phys. Lett. 71 (1997),<br />

2340–2342.<br />

[170] O. Leifeld, E. Müller, D. Grützmacher, B. Müller, <strong>and</strong> K. Kern, Appl. Phys. Lett. 74 (1999),<br />

994–996.<br />

[171] Y. Wakayama, L. V. Sokolov, N. Zakharov, P. Werner, <strong>and</strong> U. Gösele, Appl. Surf. Sci. 216<br />

(2003), 419–423.<br />

[172] G. Abstreiter, P. Schittenhelm, C. Engel, E. Silveira, A. Zrenner, <strong>and</strong> D. Meertens W. Jäger,<br />

Semicond. Sci. Technol. 11 (1996), 1521–1528.<br />

[173] M. W. Dashiell, U. Denker, C. Müller, G. Costantini, C. Manzano, K. Kern, <strong>and</strong> O. Schmidt,<br />

Appl. Phys. Lett. 80 (2002), 1279–1281.


BIBLIOGRAPHY 155<br />

[174] G. Medeiros-Ribeiro, A. M. Bratkovski, T. I. Kamins, D. A. A. Ohlberg, <strong>and</strong> R. S. Williams,<br />

Science 279 (1998), 353–355.<br />

[175] A. Rastelli <strong>and</strong> H. von Känel, Surf. Sci. Lett. 515 (2002), L493–L498.<br />

[176] Z. Gai, X. Li, R. G. Zhao, <strong>and</strong> W. S. Yang, Phys. Rev. B 57 (1998), R15060–R15063.<br />

[177] G. Costantini, A. Rastelli, C. Manzano, R. Songmuang, O. G. Schmidt, K. Kern, <strong>and</strong> H. von<br />

Känel, Appl. Phys. Lett. 85 (2004), 5673–5675.<br />

[178] E. Sutter, P. Sutter, <strong>and</strong> J. E. Bernard, Appl. Phys. Lett. 84 (2004), 2262–2264.<br />

[179] O. Kirfel, E. Müller, D. Grützmacher, K. Kern, A. Hesse, J. Stangl, V. Holy, <strong>and</strong> G. Bauer,<br />

Appl. Surf. Sci. 224 (2004), 139–142.<br />

[180] F. Montalenti, P. Raiteri D. B. Migas, H. von Känel, A. Rastelli, C. Manzano, G. Costantini,<br />

U. Denker, O. G. Schmidt, K. Kern, <strong>and</strong> L. Miglio, Phys. Rev. Lett. 93 (2004), 216102.<br />

[181] I. Berbezier, M. Descoins, B. Ismail, H. Maaref, <strong>and</strong> A. Ronda, J. Appl. Phys. 98 (2005), 063517.<br />

[182] P. Sutter, E. Sutter, <strong>and</strong> L. Vescan, Appl. Phys. Lett. 87 (2005), 161916.<br />

[183] A. Rastelli, M. Kummer, <strong>and</strong> H. von Känel, Phys. Rev. Lett. 87 (2001), 256101.<br />

[184] M. St<strong>of</strong>fel, G. S. Kar, U. Denker, A. Rastelli, H. Sigg, <strong>and</strong> O. G. Schmidt, Physica E 23 (2004),<br />

421–427.<br />

[185] A. Rastelli, E. Müller, <strong>and</strong> H. von Känel, Appl. Phys. Lett. 80 (2002), 1438–1440.<br />

[186] G. Chen, H. Lichtenberger, G. Bauer, W. Jantsch, <strong>and</strong> F. Schäffler, Phys. Rev. B 74 (2006),<br />

035302.<br />

[187] P. Waltereit, J. M. Fernández, S. Kaya, <strong>and</strong> T. J. Thornton, Appl. Phys. Lett. 72 (1998),<br />

2262–2264.<br />

[188] R. Neumann, J. Zhu, K. Brunner, <strong>and</strong> G. Abstreiter, Thin Solid Films 380 (2000), 124–126.<br />

[189] R. Neumann, K. Brunner, <strong>and</strong> G. Abstreiter, Physica E 13 (2002), 986–989.<br />

[190] T. E. Whall, J. Cryst. Growth 157 (1995), 353–361.<br />

[191] D. R. Leadley, M. J. Kearney, A. I. Horrell, H. Fischer, L. Risch, E. C. H. Parker, <strong>and</strong> T. E.<br />

Whall, Semicond. Sci. Technol. 17 (2002), 708–715.<br />

[192] D. N. Quang, V. N. Tuoc, T. D. Huan, <strong>and</strong> P. N. Phong, Phys. Rev. B 70 (2004), 195336.<br />

[193] R. People, J. C. Bean, D. V. Lang, A. M. Sergent, H. L. Störmer, K. W. Wecht, R. T. Lynch,<br />

<strong>and</strong> K. Baldwin, Appl. Phys. Lett. 45 (1984), 1231–1233.<br />

[194] E. Basaran, A. Kubiak, T. E. Whall, <strong>and</strong> E. H. C. Parker, Appl. Phys. Lett. 64 (1994), 3470–<br />

3472.<br />

[195] H. von Känel, M. Kummer, G. Isella, E. Müller, <strong>and</strong> T. Hackbarth, Appl. Phys. Lett. 80 (2002),<br />

2922–2924.


156 BIBLIOGRAPHY<br />

[196] M. Myronov, T. Irisawa, O. A. Mironov, S. Koha, Y. Shiraki, T. E. Whall, <strong>and</strong> E. H. C. Parker,<br />

Appl. Phys. Lett. 80 (2002), 3117–3119.<br />

[197] T. Manku <strong>and</strong> A. Nathan, J. Appl. Phys. 69 (1991), 8414–8416.<br />

[198] D. Monroe, Y. H. Xie, E. A. Fitzgerald, P. J. Silverman, <strong>and</strong> G. P. Watson, J. Vac. Sci. Technol.<br />

B 11 (1993), 1731–1737.<br />

[199] N. Usami, S. Fukatsu, <strong>and</strong> Y. Shiraki, Appl. Phys. Lett. 63 (1993), 388–390.<br />

[200] B. Laikhtman <strong>and</strong> R. A. Kiehl, Phys. Rev. B 47 (1993), 10515–10527.<br />

[201] A. D. Plews, N. L. Mattey, P. J. Phillips, E. H. C. Parker, <strong>and</strong> T. E. Whall, Semicond. Sci.<br />

Technol. 12 (1997), 1231–1234.<br />

[202] M. J. Kearney <strong>and</strong> A. I. Horrell, Semicond. Sci. Technol. 13 (1998), 174–180.<br />

[203] C. J. Emeleus, T. E. Whall, D. W. Smith, R. A. Kubiak, E. H. C. Parker, <strong>and</strong> M. J. Kearney,<br />

J. Appl. Phys. 73 (1993), 3852–3856.<br />

[204] M. A. Sadeghzadeh, A. I. Horrell, O. A. Mironov, T. E. Whall, E. H. C. Parker, <strong>and</strong> M. J.<br />

Kearney, Appl. Phys. Lett. 76 (2000), 2568–2570.<br />

[205] R. J. P. L<strong>and</strong>er, M. J. Kearney, A. I. Horrell, E. H. C. Parker, P. J. Phillips, <strong>and</strong> T. E. Whall,<br />

Semicond. Sci. Technol. 12 (1997), 1064–1071.<br />

[206] M. Myronov, P. J. Phillips, T. E. Whall, <strong>and</strong> E. H. C. Parker, Appl. Phys. Lett. 80 (2002),<br />

3557–3559.<br />

[207] S. Tsujino, C. V. Falub, E. Müller, M. Scheinert, L. Diehl, U. Gennser, T. Fromherz, A. Borak,<br />

H. Sigg, D. Grützmacher, Y. Campidelli, O. Kermarrec, <strong>and</strong> D. Bensahel, Appl. Phys. Lett. 84<br />

(2004), 2829–2831.<br />

[208] Well – W. Ebner, Crêt-Vaillant 17, CH-2400-LeLocle, Schweiz, Well-Drahtsäge, Modell 3032-4.<br />

[209] Thomas Berer, Fabrication <strong>of</strong> Quantum Point Contacts in the AlGaAs-System, Diploma-thesis,<br />

Institut für Halbleiterphysik, Johannes Kepler Universität Linz, August 2001.<br />

[210] TePla Technics Plasma GmbH, Dieselstr. 22a, D-85551 Kirchheim bei München, Plasma Prozes-<br />

sor 100-E.<br />

[211] Oxford Instruments Plasma Technology, Oxford Plasmalab 80 Plus Reactor, North End, Yatton,<br />

Bristol, BS194AP, UK.<br />

[212] Adrian Prinz, Schwache Lokalisierung und Elektron-Electron-Wechselwirkung in Si/<strong>SiGe</strong>-<br />

Heterostrukturen, Diploma-thesis, Institut für Halbleiterphysik, Johannes Kepler Universität<br />

Linz, Juli 1996.<br />

[213] C. Weisbuch <strong>and</strong> B. Vinter, Quantum Semiconductor Structures - Fundamentals <strong>and</strong> Applica-<br />

tions, Academic Press, 1991.


BIBLIOGRAPHY 157<br />

[214] G. Bastard, Wave Mechanics applied to Semiconductor Heterostructures, Éditions de Physique,<br />

1992.<br />

[215] K. Seeger, Semiconductor Physics - An Introduction, Springer, 6th edition, 1991.<br />

[216] Adrian Prinz, Magnetotransport Investigations <strong>of</strong> the Two-Dimensional Metallic State in Silicon<br />

Metal-Oxide-Semiconductor Structures, Ph.D. thesis, Institut für Halbleiterphysik, Johannes<br />

Kepler Universität Linz, 2002.<br />

[217] K. von Klitzing, Rev. Mod. Phys. 58 (1986), 519–531.<br />

[218] A. Gold <strong>and</strong> V. T. Dolgopolov, Phys. Rev. B 33 (1986), 1076–1084.<br />

[219] T. Ando, A. B. Fowler, <strong>and</strong> F. Stern, Rev. Mod. Phys. 54 (1982), 437–672.<br />

[220] V. Hol´y, A. A. Darhuber, J. Stangl, G. Bauer, J. Nützel, <strong>and</strong> G. Abstreiter, Phys. Rev. B 57<br />

(1998), 12435–12442.<br />

[221] B. J. Spencer, P. W. Voorhees, <strong>and</strong> J. Ters<strong>of</strong>f, Appl. Phys. Lett. 76 (2000), 3022–3024.<br />

[222] J. Ters<strong>of</strong>f, Appl. Phys. Lett. 83 (2003), 353–355.<br />

[223] U. Zeitler, H. W. Schumacher, A. G. M. Jansen, <strong>and</strong> R. J. Haug, Phys. Rev. Lett. 86 (2001),<br />

866–869.<br />

[224] M. P. Lilly, K. B. Cooper, J. P. Eisenstein, L. N. Pfeiffer, <strong>and</strong> K. W. West, Phys. Rev. Lett. 83<br />

(1999), 824–827.<br />

[225] W. Pan, R. R. Du, H. L. Stormer, D. C. Tsui, L. N. Pfeiffer, K. W. Baldwin, <strong>and</strong> K. W. West,<br />

Phys. Rev. Lett. 83 (1999), 820–823.<br />

[226] K. von Klitzing, Phil. Trans. R. Soc. A 363 (2005), 2203–2219.<br />

[227] K. Hamaya, S. Masabuchi, K. Hirakawa, S. Ishida, Y. Arakawa, K. Sawano, Y. Shiraki, <strong>and</strong><br />

T. Machida, Phys. Rev. B 73 (2006), 121304(R).<br />

[228] H. Lichtenberger, M. Mühlberger, <strong>and</strong> F. Schäffler, Appl. Phys. Lett. 82 (2003), 3650–3652.<br />

[229] H. Lichtenberger, M. Mühlberger, C. Schelling, W. Schwinger, S. Senz, <strong>and</strong> F. Schäffler, Physica<br />

E 23 (2004), 442–448.<br />

[230] J. M. Bermond, J. J. Métois, X. Egéa, <strong>and</strong> F. Floret, Surf. Sci. 330 (1995), 48–60.<br />

[231] T. Suzuki, J. J. Métois, <strong>and</strong> K. Yagi, Surf. Sci. 339 (1995), 105–113.<br />

[232] S. Tanaka, C. C. Umbach, <strong>and</strong> J. M. Blakely, J. Vac. Sci. Technol. A 15 (1997), 1345–1350.<br />

[233] L. V. Litvin, A. B. Krasilnikov, <strong>and</strong> A. V. Latyshev, Surf. Sci. Lett. 244 (1991), L121–L124.<br />

[234] E. Kasper, M. Bauer, <strong>and</strong> M. Oehme, Thin Solid Films 321 (1998), 148–152.<br />

[235] R. E. Walkup <strong>and</strong> S. I. Raider, Appl. Phys. Lett. 53 (1998), 888–890.<br />

[236] E. A. Gulbransen <strong>and</strong> S. A. Jansson, Oxid. Met. 4 (1972), 181–201.


158 BIBLIOGRAPHY<br />

[237] H. Watanabe, K. Fujita, <strong>and</strong> M. Ichikawa, Appl. Phys. Lett. 70 (1997), 1095–1097.<br />

[238] D. Jones <strong>and</strong> V. Palermo, Appl. Phys. Lett. 80 (2002), 673–675.<br />

[239] R. Tromp, G. W. Rubl<strong>of</strong>f, P. Balk, <strong>and</strong> F. K. LeGoues, Phys. Rev. Lett. 55 (1985), 2332–2335.<br />

[240] G. W. Rubl<strong>of</strong>f, K. H<strong>of</strong>mann, M. Liehr, <strong>and</strong> D. R. Young, Phys. Rev. Lett. 58 (1987), 2379–2382.<br />

[241] Julian Stangl, Fortgeschrittenenpraktikum: X-Ray Diffraction from a Semiconductor Het-<br />

erostructure.<br />

[242] Martin Wachter, Optische Untersuchungen an Potentialtopfstrukturen aus Si/<strong>SiGe</strong>, Diploma-<br />

thesis, Abteilung Halbleiterphysik, Universität Ulm, September 1992.<br />

[243] Thomas Fromherz, Infrared Spectroscopy <strong>of</strong> Electronic <strong>and</strong> Vibrational Excitons in Semicon-<br />

ductor Quantum Wells <strong>and</strong> Superlattices, Ph.D. thesis, Institut für Halbleiterphysik, Johannes<br />

Kepler Universität Linz, 1994.<br />

[244] Patrick Rauter, Intrab<strong>and</strong> Absorption <strong>and</strong> Photospectroscopy <strong>of</strong> <strong>SiGe</strong> Quantum Cascades,<br />

Diploma-thesis, Institut für Halbleiterphysik, Johannes Kepler Universität Linz, August 2003.<br />

[245] M. M. Rieger <strong>and</strong> P. Vogl, Phys. Rev. B 48 (1993), 14276–14287.<br />

[246] L. Colombo, R. Resta, <strong>and</strong> S. Baroni, Phys. Rev. B 44 (1991), 5572–5579.<br />

[247] C. G. Van de Walle, Phys. Rev. B 39 (1989), 1871–1883.


Postface<br />

Curriculum Vitae<br />

⋄ ∗ 13. 09. 1976 born in Linz<br />

⋄ Sept. 1983 – July 1987 VS Ottensheim (primary school)<br />

⋄ Sept. 1987 – July 1995 Akademisches Gymnasium Linz Spittelwiese (secondary<br />

school)<br />

⋄ Oct. 1995 – Sept. 1996 alternative (military) service<br />

⋄ Oct. 1996 – Sept. 2002 Technical Physics studies at the Johannes Kepler Uni-<br />

versity Linz<br />

⋄ May 2001 – July 2002 Diploma thesis at the Institute <strong>of</strong> Semiconductor <strong>and</strong><br />

Solid State Physics on ”Characterization <strong>and</strong> Over-<br />

growth <strong>of</strong> Prestructured Silicon-Substrates”<br />

⋄ from Oct. 2002 PhD thesis at the Institute <strong>of</strong> Semiconductor <strong>and</strong> Solid<br />

State Physics <strong>and</strong> composing the thesis at h<strong>and</strong><br />

159


160 POSTFACE<br />

List <strong>of</strong> Publications<br />

� H. Lichtenberger, M. Mühlberger <strong>and</strong> F. Schäffler, ”Transient-enhanced Si diffusion on<br />

native-oxide-covered Si(001) nanostructures during vacuum annealing”, Appl. Phys.<br />

Lett. 82, 3650–3652 (2003)<br />

� Z. Zhong, A. Halilovic, H. Lichtenberger, F. Schäffler, G. Bauer, ”Growth <strong>of</strong> Ge isl<strong>and</strong>s<br />

on prepatterned Si(001) substrates”, Physica E 23, 243–247 (2004)<br />

� H. Lichtenberger, M. Mühlberger, C. Schelling, W. Schwinger, S. Senz <strong>and</strong> F. Schäffler,<br />

”Transient-enhanced Si diffusion on natural-oxide-covered Si(001) nano-structures dur-<br />

ing vacuum annealing”, Physica E 23, 442–448 (2004)<br />

� C. Schelling, J. Mysliveček, M. Mühlberger, H. Lichtenberger, Z. Zhong, B. Voigtländer,<br />

G. Bauer <strong>and</strong> F. Schäffler ”<strong>Kinetic</strong> <strong>and</strong> <strong>Strain</strong>-Driven Growth Phenomena on Si(001)”,<br />

Phys. stat. sol. (a) 201, 321–328 (2004)<br />

� J. P. Leitão, A. Fonseca, N. A. Sobolev, M. C. Carmo, N. Franco, A. D. Sequeira,<br />

T. M. Burbaev, V. A. Kurbatov, M. M. Rzaev, A. O. Pogosov, N. N. Sibeldin, V. A.<br />

Tsvetkov, H. Lichtenberger <strong>and</strong> F. Schäffler, ”Low-temperature molecular beam epitaxy<br />

<strong>of</strong> Ge on Si”, Mat. Sci. Semicond. Proc. 8, 35–39 (2005)<br />

� H. Lichtenberger, M. Mühlberger, C. Schelling <strong>and</strong> F. Schäffler, ”Ordering <strong>of</strong> self-<br />

assembled Si0.55Ge0.45 isl<strong>and</strong>s on vicinal Si(001) substrates”, J. Cryst. Growth 278,<br />

78–82 (2005)<br />

� T. M. Burbaev, V. A. Kurbatov, M. M. Rzaev, A. O. Pogosov, N. N. Sibel’din, V. A.<br />

Tsvetkov, H. Lichtenberger, F. Schäffler, J. P. Leitão, N. A. Sobolev <strong>and</strong> M. C. Carmo,<br />

”Morphological Transformation <strong>of</strong> a Germanium Layer Grown on a Silicon Surface by<br />

Molecular-Beam Epitaxy at Low Temperatures”, Phys. Solid State 47, 71–75 (2005)<br />

� H. Lichtenberger, M. Mühlberger <strong>and</strong> F. Schäffler, ”Ordering <strong>of</strong> Si0.55Ge0.45 Isl<strong>and</strong>s on<br />

Vicinal Si(001) Substrates: The Interplay between <strong>Kinetic</strong> Step Bunching <strong>and</strong> <strong>Strain</strong>-<br />

Driven Isl<strong>and</strong> Growth”, Appl. Phys. Lett. 86, 131919 (2005)<br />

� W. Jantsch, H. Malissa, Z. Wilamowski, H. Lichtenberger, G. Chen, F. Schäffler <strong>and</strong><br />

G. Bauer, ”Spin Properties <strong>of</strong> Electrons in Low-Dimensional <strong>SiGe</strong> Structures”, Journal<br />

<strong>of</strong> Superconductivity 18, 145–149 (2005)


POSTFACE 161<br />

� D. Gruber, D. Pachinger, H. Malissa, M. Mühlberger, H. Lichtenberger, W. Jantsch<br />

<strong>and</strong> F. Schäffler, ”g-Factor tuning <strong>of</strong> 2D electrons in double-gated Si/<strong>SiGe</strong> quantum<br />

wells”, Physica E 32, 254–257 (2006)<br />

� T. Berer, D. Pachinger, G. Pillwein, M. Mühlberger, H. Lichtenberger, G. Brunthaler<br />

<strong>and</strong> F. Schäffler, ”Single-electron transistor in strained Si/<strong>SiGe</strong> heterostructures”, Phys-<br />

ica E 34, 456–459 (2006)<br />

� Z. Zhong, H. Lichtenberger, G. Chen, M. Mühlberger, C. Schelling, J. Mysliveček, A.<br />

Halilovic, J. Stangl, G. Bauer <strong>and</strong> W. Jantsch <strong>and</strong> F. Schäffler, ”Ordered <strong>SiGe</strong> isl<strong>and</strong>s<br />

on vicinal <strong>and</strong> pre-patterned Si(001) substrates”, Microelectronic Engineering 83, 1730–<br />

1735 (2006)<br />

� G. Chen, H. Lichtenberger, F. Schäffler, G. Bauer <strong>and</strong> W. Jantsch, ”Geometry depen-<br />

dent nucleation mechanism for <strong>SiGe</strong> isl<strong>and</strong>s grown on pit-patterned Si(001) substrates”,<br />

Materials Science <strong>and</strong> Engineering: C 26, 795–799 (2006)<br />

� H. Malissa, W. Jantsch, G. Chen, D. Gruber, H. Lichtenberger, F. Schäffler, Z. Wil-<br />

amowski, A. Tyryshkin <strong>and</strong> S. Lyon, ”Investigation <strong>of</strong> the spin properties <strong>of</strong> electrons<br />

in zero-dimensional <strong>SiGe</strong> structures by electron paramagnetic resonance”, Materials<br />

Science <strong>and</strong> Engineering B 126, 172–175 (2006)<br />

� T. Berer, D. Pachinger, G. Pillwein, M. Mühlberger, H. Lichtenberger, G. Brunthaler<br />

<strong>and</strong> F. Schäffler, ”Lateral quantum dots in Si/<strong>SiGe</strong> realized by a Schottky split-gate<br />

technique”, Appl. Phys. Lett. 88, 162112 (2006)<br />

� G. Chen, H. Lichtenberger, G. Bauer <strong>and</strong> W. Jantsch, F. Schäffler, ”Initial stage <strong>of</strong> the<br />

2D-3D transition <strong>of</strong> a strained <strong>SiGe</strong> layer on a pit-patterned Si(001) template”, Phys.<br />

Rev. B 74, 035302 (2006)


162 POSTFACE<br />

Acknowledgements<br />

I would like to thank all the people who have supported me with my PhD thesis <strong>and</strong> have<br />

contributed a lot to the success <strong>of</strong> this work:<br />

� Pr<strong>of</strong>. Dr. Friedrich Schäffler for the interesting topic, for the productive discussions<br />

<strong>and</strong> for his ongoing support in every respect.<br />

� Dr. Detlev Grützmacher for taking on the second report.<br />

� All other members <strong>of</strong> the institute <strong>and</strong> the people in the <strong>of</strong>fice for the friendly atmo-<br />

sphere <strong>and</strong> the good collegiality. I cannot name all the people personally here, as the<br />

list would be simply too long. However, I am deeply grateful for their assistance in<br />

different respects. They <strong>of</strong>fered technical or administrative support, helped h<strong>and</strong>ling<br />

computer problems or contributed in valuable physical discussions.<br />

� My family for the encouragement <strong>and</strong> making all this possible.


POSTFACE 163<br />

Abbreviations<br />

AFM Atomic Force Microscope<br />

HRXTEM High Resolution CrossSectional Transmission Electron Microscope<br />

MBE Molecular Beam Epitaxy<br />

MODQW Modulation-doped Quantum Well<br />

PL Photoluminescence<br />

SAP Surface Angle Plot<br />

SEM Scanning Electron Microscope<br />

SOM Surface Orientation Map<br />

TEM Transmission Electron Microscope

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!