11.03.2014 Views

Itinerant Spin Dynamics in Structures of ... - Jacobs University

Itinerant Spin Dynamics in Structures of ... - Jacobs University

Itinerant Spin Dynamics in Structures of ... - Jacobs University

SHOW MORE
SHOW LESS

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

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

<strong>It<strong>in</strong>erant</strong> <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> <strong>in</strong> <strong>Structures</strong> <strong>of</strong> Reduced<br />

Dimensionality<br />

A dissertation presented<br />

by<br />

Paul Thomas Wenk<br />

<strong>in</strong> partial fulfillment <strong>of</strong> the requirements<br />

for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

<strong>in</strong> the subject <strong>of</strong><br />

Physics<br />

Thesis Committee:<br />

Pr<strong>of</strong>. Dr. Stefan Kettemann [thesis advisor] (<strong>Jacobs</strong> <strong>University</strong> Bremen, Germany &<br />

Pohang <strong>University</strong> <strong>of</strong> Science and Technology, South Korea)<br />

Pr<strong>of</strong>. Dr. Georges Bouzerar(Institut Néel, France & <strong>Jacobs</strong> <strong>University</strong> Bremen, Germany)<br />

Pr<strong>of</strong>. Dr. Ulrich Kle<strong>in</strong>ekathöfer(<strong>Jacobs</strong> <strong>University</strong> Bremen, Germany)<br />

<strong>Jacobs</strong> <strong>University</strong> Bremen<br />

Date <strong>of</strong> Defense: August, 29 2011<br />

———————————<br />

School <strong>of</strong> Eng<strong>in</strong>eer<strong>in</strong>g and Science


©2011 - Paul Thomas Wenk<br />

All rights reserved.


<strong>It<strong>in</strong>erant</strong> <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> <strong>in</strong> <strong>Structures</strong> <strong>of</strong> Reduced Dimensionality<br />

Abstract<br />

In the present thesis results <strong>of</strong> the study <strong>of</strong> sp<strong>in</strong> dynamics and quantum transport<br />

<strong>in</strong> disordered semiconductor quantum wires with sp<strong>in</strong>-orbit coupl<strong>in</strong>g are presented.<br />

Start<strong>in</strong>g from basic sp<strong>in</strong> dynamics we derive the dependence <strong>of</strong> the weak localization correction<br />

to the conductance on the strength and the k<strong>in</strong>d <strong>of</strong> sp<strong>in</strong>-orbit <strong>in</strong>teraction (l<strong>in</strong>ear<br />

and cubic Dresselhaus, as well as Rashba coupl<strong>in</strong>g), the width <strong>of</strong> the quantum wires as<br />

well as the mobility, temperature and Zeeman term. Furthermore, we exploit the connection<br />

found between the microscopic picture given by the Cooperon and the sp<strong>in</strong> diffusion<br />

equation to extract the sp<strong>in</strong> relaxation rate which shows the same wire dependencies as<br />

the weak localization correction. We also show how the result depends on the smoothness<br />

and the direction <strong>of</strong> the transverse conf<strong>in</strong>ement <strong>of</strong> the quantum wires. In this context we<br />

have addressed the question concern<strong>in</strong>g long persist<strong>in</strong>g or even persistent sp<strong>in</strong> states <strong>in</strong><br />

sp<strong>in</strong>tronic devices, present<strong>in</strong>g the correspond<strong>in</strong>g optimal adjustment <strong>of</strong> sp<strong>in</strong> orbit coupl<strong>in</strong>gs<br />

<strong>of</strong> different k<strong>in</strong>d and optimal alignment <strong>of</strong> the wire direction <strong>in</strong> semiconductor crystals.<br />

Experiments[HSM + 06, HSM + 07, KKN09, LSK + 07, WGZ + 06, SGB + 09] which report the<br />

dimensional reduction <strong>of</strong> the sp<strong>in</strong> relaxation rate <strong>in</strong> agreement with previous results were<br />

rais<strong>in</strong>g new questions, <strong>in</strong> particular as regard<strong>in</strong>g the crossover from diffusive to ballistic<br />

wires, which we answer us<strong>in</strong>g modified Cooperon equation. In addition, we focus on the<br />

<strong>in</strong>tr<strong>in</strong>sic sp<strong>in</strong> Hall effect, which is only due to sp<strong>in</strong>-orbit coupl<strong>in</strong>g. Hav<strong>in</strong>g shown the basic<br />

features with analytical calculations, we solve the sp<strong>in</strong> Hall conductivity <strong>in</strong> presence <strong>of</strong><br />

b<strong>in</strong>ary and block-distributed impurities (Anderson model). At this we apply the Kernel<br />

Polynomial Method, which allows for a f<strong>in</strong>ite size analysis <strong>of</strong> the metal-<strong>in</strong>sulator transition<br />

and the calculation <strong>of</strong> sp<strong>in</strong> Hall conductivity <strong>in</strong> large systems compared with those<br />

addressable with exact diagonalization.<br />

iii


Contents<br />

Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

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

Table <strong>of</strong> Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

Citations to Previously Published Work . . . . . . . . . . . . . . . . . . . . . . .<br />

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

i<br />

iii<br />

iv<br />

vii<br />

viii<br />

ix<br />

1 Introduction 1<br />

2 <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 6<br />

2.1 Short Rem<strong>in</strong>der on the Orig<strong>in</strong> <strong>of</strong> <strong>Sp<strong>in</strong></strong> Orbit Coupl<strong>in</strong>g . . . . . . . . . . . . 6<br />

2.2 <strong>Dynamics</strong> <strong>of</strong> a Localized <strong>Sp<strong>in</strong></strong> . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.3 <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> <strong>of</strong> <strong>It<strong>in</strong>erant</strong> Electrons . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.3.1 Ballistic <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.3.2 <strong>Sp<strong>in</strong></strong> Diffusion Equation . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.3.3 <strong>Sp<strong>in</strong></strong> Orbit Interaction <strong>in</strong> Semiconductors . . . . . . . . . . . . . . . 11<br />

2.3.4 <strong>Sp<strong>in</strong></strong> Diffusion <strong>in</strong> the Presence <strong>of</strong> <strong>Sp<strong>in</strong></strong>-Orbit Interaction . . . . . . . 15<br />

2.4 <strong>Sp<strong>in</strong></strong> Relaxation Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.4.1 D’yakonov-Perel’ <strong>Sp<strong>in</strong></strong> Relaxation . . . . . . . . . . . . . . . . . . . 19<br />

2.4.2 DP <strong>Sp<strong>in</strong></strong> Relaxation with Electron-Electron and Electron-Phonon<br />

Scatter<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.4.3 Elliott-Yafet <strong>Sp<strong>in</strong></strong> Relaxation . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.4.4 <strong>Sp<strong>in</strong></strong> Relaxation due to <strong>Sp<strong>in</strong></strong>-Orbit Interaction with Impurities . . . 21<br />

2.4.5 Bir-Aronov-Pikus <strong>Sp<strong>in</strong></strong> Relaxation . . . . . . . . . . . . . . . . . . . 22<br />

2.4.6 Magnetic Impurities . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.4.7 Nuclear <strong>Sp<strong>in</strong></strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.4.8 Magnetic Field Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation . . . . . . . . . . . . 23<br />

3 WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 25<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.1.1 One-Dimensional Wires . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.1.2 Wires with W > λ F . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.2 Quantum Transport Corrections . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

3.2.1 Diagrammatic Approach . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

iv


Contents<br />

v<br />

3.2.2 Weak Localization <strong>in</strong> Quantum Wires . . . . . . . . . . . . . . . . . 34<br />

3.3 The Cooperon and <strong>Sp<strong>in</strong></strong> Diffusion <strong>in</strong> 2D . . . . . . . . . . . . . . . . . . . . 37<br />

3.4 Solution <strong>of</strong> the Cooperon Equation <strong>in</strong> Quantum Wires . . . . . . . . . . . . 42<br />

3.4.1 Quantum Wires with <strong>Sp<strong>in</strong></strong>-Conserv<strong>in</strong>g Boundaries . . . . . . . . . . 42<br />

3.4.2 Zero-Mode Approximation . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.4.3 Exact Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.4.4 Other Types <strong>of</strong> Boundary Conditions . . . . . . . . . . . . . . . . . 53<br />

3.5 Magnetoconductivity with Zeeman splitt<strong>in</strong>g . . . . . . . . . . . . . . . . . . 57<br />

3.5.1 2DEG . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.5.2 Quantum Wire with <strong>Sp<strong>in</strong></strong>-Conserv<strong>in</strong>g Boundary Conditions . . . . . 60<br />

3.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

4 Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 67<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.1.1 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.2 <strong>Sp<strong>in</strong></strong> Relaxation anisotropy <strong>in</strong> the (001) system . . . . . . . . . . . . . . . . 70<br />

4.2.1 2D system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2.2 Quasi-1D wire . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.3 <strong>Sp<strong>in</strong></strong> relaxation <strong>in</strong> quasi-1D wire with [110] growth direction . . . . . . . . . 75<br />

4.3.1 Special case: without cubic Dresselhaus SOC . . . . . . . . . . . . . 76<br />

4.3.2 With cubic Dresselhaus SOC . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.4 Weak Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.5 Diffusive-Ballistic Crossover . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

4.5.1 <strong>Sp<strong>in</strong></strong> Relaxation at Q SO W ≪ 1 . . . . . . . . . . . . . . . . . . . . . 80<br />

4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5 <strong>Sp<strong>in</strong></strong> Hall Effect 83<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

5.1.1 About the Def<strong>in</strong>ition <strong>of</strong> <strong>Sp<strong>in</strong></strong> Current . . . . . . . . . . . . . . . . . . 84<br />

5.2 SHE without Impurities: Exact Calculation . . . . . . . . . . . . . . . . . . 85<br />

5.3 Numerical Analysis <strong>of</strong> SHE . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

5.3.1 Exact Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

5.3.2 Kernel Polynomial Method . . . . . . . . . . . . . . . . . . . . . . . 92<br />

5.3.3 SHC calculation us<strong>in</strong>g KPM . . . . . . . . . . . . . . . . . . . . . . . 101<br />

6 Critical Discussion and Future Perspective 106<br />

List <strong>of</strong> Symbols 108<br />

List <strong>of</strong> Figures 110<br />

List <strong>of</strong> Tables 115<br />

Bibliography 116<br />

A SOC Strength <strong>in</strong> the Experiment 129


vi<br />

Contents<br />

B L<strong>in</strong>ear Response 132<br />

B.1 Kubo Formula for Weak Disorder . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

C Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation 137<br />

C.1 Sum Formula for the Cooperon . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

C.2 <strong>Sp<strong>in</strong></strong>-Conserv<strong>in</strong>g Boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

C.3 Relaxation Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140<br />

C.4 Weak Localization Correction <strong>in</strong> 2D . . . . . . . . . . . . . . . . . . . . . . 141<br />

C.5 Exact Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

D Hamiltonian <strong>in</strong> [110] growth direction 146<br />

E Summation over the Fermi Surface 148<br />

F KPM 150


Citations to Previously Published Work<br />

Large portions <strong>of</strong> the chapters have appeared <strong>in</strong> the follow<strong>in</strong>g four papers:<br />

1. Wenk P., Kettemann S. et al. <strong>Sp<strong>in</strong></strong> Polarized Transport and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong><br />

QuantumWires. In: NanoScienceandTechnology. Spr<strong>in</strong>ger-VerlagBerl<strong>in</strong>Heidelberg;<br />

(2010).<br />

2. Wenk P., Kettemann S. <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Quantum Wires. In: Sattler K, Francis&<br />

Taylor Handbook on Nanophysics.; (2010) p.49.<br />

3. Wenk P., Kettemann S. Dimensional Dependence <strong>of</strong> Weak Localization Corrections<br />

and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Quantum Wires with Rashba <strong>Sp<strong>in</strong></strong> Orbit Coupl<strong>in</strong>g. Phys.<br />

Rev. B. 81 125309 (2010).<br />

4. Wenk P., Kettemann S. Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Two-<br />

Dimensional Systems. Phys. Rev. B. 83 115301 (2011).<br />

vii


Acknowledgments<br />

Complet<strong>in</strong>gthis doctoral work has been awonderfulexperience. It is with pleasure<br />

that I acknowledge the advise and support by the people with whom I was work<strong>in</strong>g dur<strong>in</strong>g<br />

the time <strong>in</strong> Hamburg and Bremen.<br />

In particular I owe my <strong>in</strong>terest <strong>in</strong> the fasc<strong>in</strong>at<strong>in</strong>g world <strong>of</strong> condensed matter to my thesis<br />

advisor Pr<strong>of</strong>essor Stefan Kettemann. It started all by a small and seem<strong>in</strong>gly simple task<br />

concern<strong>in</strong>g ballistic sp<strong>in</strong> transport. Dur<strong>in</strong>g the <strong>in</strong>duction <strong>in</strong>to the field <strong>of</strong> sp<strong>in</strong>tronics his<br />

advice helped me to see on the one hand the richness <strong>of</strong> effects which were <strong>of</strong>ten counter<strong>in</strong>tuitive,<br />

and on the other hand he showed me how to get a feel<strong>in</strong>g for the nature <strong>of</strong> the<br />

complex problem we were work<strong>in</strong>g on by application <strong>of</strong> fundamental rules. He gave me<br />

plenty <strong>of</strong> scope to chose my direction <strong>of</strong> <strong>in</strong>terest and always encouraged me <strong>in</strong> many fruitful<br />

discussions.<br />

Pr<strong>of</strong>essor Georges Bouzerar I owe not only the rediscovery <strong>of</strong> my <strong>in</strong>terest <strong>in</strong> programm<strong>in</strong>g<br />

and a boost <strong>in</strong> my rusty C++ skills from school time, but also a better understand<strong>in</strong>g <strong>of</strong><br />

how to squeeze the essential <strong>of</strong> an phenomenon <strong>in</strong>to the computer and extract mean<strong>in</strong>gful<br />

physics out <strong>of</strong> it. His questions furthered my <strong>in</strong>sight <strong>of</strong> the subject matter.<br />

Throughout my years as a PhD student, I was supported for many semesters by the DFG-<br />

SFB508 B9 at the Institute for Theoretical Physics <strong>in</strong> Hamburg and by the DFG CC5120<br />

at the <strong>Jacobs</strong> <strong>University</strong> Bremen, through the generosity <strong>of</strong> my advisor. I thank the Conrad<br />

Naber foundation for provid<strong>in</strong>g my scholarship for my first year at the <strong>Jacobs</strong> <strong>University</strong><br />

Bremen. In addition I had the opportunity to take part <strong>in</strong> many <strong>in</strong>terest<strong>in</strong>g and productive<br />

conferences (some <strong>in</strong>clud<strong>in</strong>g even s<strong>in</strong>g<strong>in</strong>g) where some were founded by the World Class<br />

<strong>University</strong> program through the Korea Science and Eng<strong>in</strong>eer<strong>in</strong>g Foundation.<br />

F<strong>in</strong>ally, I also would like to acknowledge the support <strong>of</strong> my colleagues at the Hamburg and<br />

<strong>Jacobs</strong> <strong>University</strong>, creat<strong>in</strong>g a stimulat<strong>in</strong>g environment.<br />

viii


Dedicated to L<strong>in</strong>da,<br />

my parents<br />

and my brother Daniel.<br />

Stephanie: So, how was your day?<br />

Leonard: Y’know, I’m a physicist - I thought about stuff.<br />

Stephanie: That’s it?<br />

Leonard: I wrote some <strong>of</strong> it down.<br />

From “The Big Bang Theory“ episode The Panty P<strong>in</strong>ata Polarization.<br />

ix


x<br />

Acknowledgments


Chapter 1<br />

Introduction<br />

Structure <strong>of</strong> this thesis<br />

This thesis falls <strong>in</strong>to four parts,<br />

• Fundamentals <strong>of</strong> sp<strong>in</strong> dynamics and sp<strong>in</strong> relaxation mechanisms (Chapter1 and 2),<br />

• <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> <strong>in</strong> Quantum Wires:<br />

– Dimensional dependence <strong>of</strong> weak localization and weak antilocalization and the<br />

relation to sp<strong>in</strong> diffusion (Chapter3)<br />

– Direction dependence <strong>of</strong> sp<strong>in</strong> relaxation and diffusive-ballistic crossover (Chapter4)<br />

• <strong>Sp<strong>in</strong></strong> Hall Effect (Chapter5)<br />

• Critical Discussion and Future Perspective (Chapter6)<br />

The field <strong>of</strong> <strong>in</strong>terest which we are go<strong>in</strong>g to present <strong>in</strong> this PhD thesis is called<br />

sp<strong>in</strong>tronics. Nowadays it is not only a Gedankenexperiment, the emerg<strong>in</strong>g technology <strong>of</strong> it<br />

is already partially <strong>in</strong> use, e.g. <strong>in</strong> Magnetoresistive RAMs[AF07] which use the giant magnetoresistance<br />

discovered by Albert Fert and Peter Grünberg (Nobel Prize 2007)[BBF + 88,<br />

BGSZ89]. But what is actually sp<strong>in</strong>tronics? In a paper from 2004, D’yakonov states that<br />

“What most people apparently mean by sp<strong>in</strong>tronics is the fabrication <strong>of</strong> some useful devices<br />

us<strong>in</strong>g a) creation <strong>of</strong> a non-equilibrium sp<strong>in</strong> density <strong>in</strong> a semiconductor, b) manipulation <strong>of</strong><br />

the sp<strong>in</strong>s by external fields, and c) detection <strong>of</strong> the result<strong>in</strong>g sp<strong>in</strong> state.”[Dya04]. One part<br />

1


2 Chapter 1: Introduction<br />

<strong>of</strong> this work will ma<strong>in</strong>ly focus on the second po<strong>in</strong>t, manipulation <strong>of</strong> the sp<strong>in</strong> and understand<strong>in</strong>g<br />

the limitations <strong>of</strong> sp<strong>in</strong> propagation. The ma<strong>in</strong> objective is to shed light on some<br />

<strong>in</strong>terest<strong>in</strong>g effects <strong>in</strong> the field <strong>of</strong> sp<strong>in</strong>-dependent electronic transport.<br />

Look<strong>in</strong>g <strong>in</strong> the literature one realizes that many proposals for two-dimensional (2D) sp<strong>in</strong>tronic<br />

devices are based on the presence <strong>of</strong> sp<strong>in</strong>-orbit coupl<strong>in</strong>g (SOC) <strong>in</strong> a 2D electron<br />

system (2DES) semiconductor heterostructure. This idea goes back to the sp<strong>in</strong> field-effect<br />

transistor proposed by Datta and Das[DD90] which is schematically plotted <strong>in</strong> Fig.1.1. The<br />

(a)<br />

(b)<br />

Figure 1.1: (a) a) Schematic <strong>of</strong> Datta-Das sp<strong>in</strong> modulator device <strong>in</strong> a cross-section. The 2D<br />

electron gas (2DEG) has a distance <strong>of</strong> L from the emitter (↑) to the collector (↓). Normal to<br />

thiscross-section thereisanadditional conf<strong>in</strong>ement. b)Theconductionbandwhichconf<strong>in</strong>es<br />

the electrons to a 2D system and the electron distribution (dotted) are shown. Taken from<br />

Ref.[NATE97]. (b) Manipulation <strong>of</strong> sp<strong>in</strong> precession due to SOC by gate voltage.<br />

electrons are <strong>in</strong>jected from metallic leads <strong>in</strong>to the 2DES with the sp<strong>in</strong> parallel to the transport<br />

direction. In the nonmagnetic region the sp<strong>in</strong> is changed along the distance L due to


Chapter 1: Introduction 3<br />

the coupl<strong>in</strong>g <strong>of</strong> sp<strong>in</strong> degree <strong>of</strong> freedom and orbital motion <strong>of</strong> the electron. In this model<br />

the electrons are mov<strong>in</strong>g ballistically <strong>in</strong> a quasi one-dimensional channel, due to a second<br />

conf<strong>in</strong>ement. The result for weak SOC <strong>in</strong> first order perturbation theory is a modulation<br />

<strong>of</strong> the sp<strong>in</strong> current with a phase-shift<br />

∆θ = 2αm e<br />

2 L, (1.1)<br />

with α the strength <strong>of</strong> SOC and m e the effective electron mass. This phase-shift ∆θ can be<br />

manipulated by the conf<strong>in</strong><strong>in</strong>g field E z via the gate which <strong>in</strong> turn changes the probability<br />

to f<strong>in</strong>d the sp<strong>in</strong> <strong>in</strong> the ”down” state at the dra<strong>in</strong>. The gate-control makes the sp<strong>in</strong>-FET so<br />

promis<strong>in</strong>g for sp<strong>in</strong>tronic applications. In Sec.2.3.3 we will review this transport <strong>in</strong> ballistic<br />

wires.<br />

Apply<strong>in</strong>g this model <strong>in</strong> an experiment one is, however, confronted with several problems.<br />

<strong>Sp<strong>in</strong></strong>tronic devices which rely on coherent sp<strong>in</strong> precession <strong>of</strong> conduction electrons[DD90,<br />

ZFD04] require a small sp<strong>in</strong> relaxation rate 1/τ s . But as the electron momentum is randomized<br />

due to disorder, the coupl<strong>in</strong>g between sp<strong>in</strong> and orbital degree <strong>of</strong> freedom, the<br />

sp<strong>in</strong>-orbit (SO) <strong>in</strong>teraction, is expected to result not only <strong>in</strong> a sp<strong>in</strong> precession but <strong>in</strong> randomization<br />

<strong>of</strong> the electron sp<strong>in</strong>. This coupl<strong>in</strong>g can lead to counter <strong>in</strong>tuitive effects as the<br />

follow<strong>in</strong>g, described by D’yakonov and Perel’[DP72]: Analyz<strong>in</strong>g the sp<strong>in</strong> transport <strong>in</strong> e.g.<br />

n-type semiconductors at low temperature (T 5K), one f<strong>in</strong>ds that the more the electron<br />

is scattered the longer is the life time <strong>of</strong> the <strong>in</strong>itial sp<strong>in</strong> state. Such experiments are<br />

<strong>of</strong>ten performed <strong>in</strong> devises where the wire width W is several nanometers wide, so that<br />

boundary effects can play an important role: Look<strong>in</strong>g at the extreme situation where W<br />

is <strong>of</strong> the order <strong>of</strong> Fermi wavelength λ F , the D’yakonov-Perel’ sp<strong>in</strong> relaxation is expected<br />

to vanish,[KK00, MFA02] s<strong>in</strong>ce the back scatter<strong>in</strong>g from impurities can <strong>in</strong> one-dimensional<br />

wires only reverse the SO field and thereby the sp<strong>in</strong> precession. Immediately the follow<strong>in</strong>g<br />

question arises: How many channels can be added without enlarg<strong>in</strong>g the sp<strong>in</strong> relaxation rate<br />

significantly? In Ref.[Ket07], which is the start<strong>in</strong>g po<strong>in</strong>t for the analysis presented <strong>in</strong> the<br />

first part <strong>of</strong> the present work, S. Kettemann could show, that 1/τ s is already strongly reduced<br />

<strong>in</strong> much wider wires: as soon as the wirewidth W is smaller than bulksp<strong>in</strong> precession<br />

length L SO , which is the length on which the electron sp<strong>in</strong> precesses a full cycle. S<strong>in</strong>ce L SO<br />

can be several µm and is not changed significantly as the wire width W is reduced, the<br />

reduction <strong>of</strong> sp<strong>in</strong> relaxation can be very useful for applications.<br />

Thestudy<strong>of</strong> thereduction <strong>of</strong> the sp<strong>in</strong>relaxation rate <strong>in</strong> quantum wires for widths exceed<strong>in</strong>g


4 Chapter 1: Introduction<br />

both the elastic mean-free path l e and λ F is <strong>in</strong> addition motivated by the fact that recently<br />

it could be observed with optical[HSM + 06] as well as with measurements <strong>of</strong> the change <strong>in</strong><br />

magnetoconductivity.[DLS + 05, LSK + 07, SGP + 06, WGZ + 06, KKN09]<br />

We have a tool to study such dephas<strong>in</strong>g and symmetry-break<strong>in</strong>g mechanisms <strong>in</strong><br />

conductors[AAKL82, Ber84, CS86]: It is the quantum <strong>in</strong>terference <strong>of</strong> electrons <strong>in</strong> lowdimensional,<br />

disordered conductors. We will show <strong>in</strong> Chapter3 how this effect results <strong>in</strong><br />

corrections to the electrical conductivity ∆σ, which is known as weak localization (WL)<br />

effect. Obviously also here we focus on systems with entanglement <strong>of</strong> sp<strong>in</strong> and charge by<br />

SO <strong>in</strong>teraction. The SO field, which has various forms <strong>in</strong> the semiconductor, makes the<br />

effect richer because it can enhances the conductivity by revers<strong>in</strong>g the effect <strong>of</strong> WL. This<br />

is called weak antilocalization (WAL). We are go<strong>in</strong>g to calculate the correction with the<br />

Cooperon equation. Due to the fact that the orig<strong>in</strong> <strong>of</strong> the <strong>in</strong>terference-suppression is the<br />

randomization <strong>of</strong> the sp<strong>in</strong>, more precise, the D’yakonov and Perel’ sp<strong>in</strong> relaxation <strong>in</strong> this<br />

work, it seems natural to establish a connection between the sp<strong>in</strong> diffusion picture, e.g<br />

derived from the Eilenberger equation by Schwab et al.[SDGR06], and the local correction<br />

described by us<strong>in</strong>g the Cooperon equation, as we are go<strong>in</strong>g to shown <strong>in</strong> Sec.3.3. The l<strong>in</strong>k<br />

to applications <strong>in</strong> the field <strong>of</strong> sp<strong>in</strong>tronics is a better <strong>in</strong>sight <strong>in</strong> what possibilities we have to<br />

create long liv<strong>in</strong>g sp<strong>in</strong> modes which will appear as special cases <strong>in</strong> our calculation.<br />

As mentioned, the addition <strong>of</strong> boundaries can change the sp<strong>in</strong> relaxation significantly.<br />

The change depends <strong>of</strong> k<strong>in</strong>d and direction <strong>of</strong> the wire boundaries. Chapter4 will<br />

focus especially on the latter. The connection between the WL and the sp<strong>in</strong> diffusion will<br />

help us to understand both at the same time: The magnetoconductivity and the existence<br />

<strong>of</strong> persistent sp<strong>in</strong> states for a given wire.<br />

Hav<strong>in</strong>g derived a consistent theory <strong>of</strong> sp<strong>in</strong> relaxation <strong>in</strong> quantum wires, one could wonder<br />

why there are measurements <strong>of</strong> sp<strong>in</strong> lifetime, like done by Kunihashi et al.[KKN09] <strong>in</strong> gatefitted<br />

narrowwiresfrommagnetotransport experiments, whichshowdeviations fromtheory.<br />

This happens although all dom<strong>in</strong>ant SO coupl<strong>in</strong>g (SOC) types, namely l<strong>in</strong>ear Rashba and<br />

l<strong>in</strong>. and cubic Dresselhaus SOC, have been <strong>in</strong>cluded <strong>in</strong> the theory. The crucial po<strong>in</strong>t is: If<br />

a part <strong>of</strong> the SOCs is left out, i.e. the cubic Dresselhaus SOC, one has great accordance between<br />

experiment and theory. In Chapter4 we are go<strong>in</strong>g to show how this puzzle is resolved<br />

by do<strong>in</strong>g a crossover from the diffusive to the ballistic regime and reveal<strong>in</strong>g a suppression<br />

<strong>of</strong> the problematic terms.


Chapter 1: Introduction 5<br />

Throughout this work we set ≡ 1.


Chapter 2<br />

<strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and<br />

Analysis <strong>of</strong> 2D Systems<br />

2.1 Short Rem<strong>in</strong>der on the Orig<strong>in</strong> <strong>of</strong> <strong>Sp<strong>in</strong></strong> Orbit Coupl<strong>in</strong>g<br />

The<strong>in</strong>teraction whichmakes sp<strong>in</strong>tronicdevices so<strong>in</strong>terest<strong>in</strong>g is arelativistic effect,<br />

theSOC.The<strong>in</strong>tr<strong>in</strong>sicdegree<strong>of</strong>freedomsp<strong>in</strong>isadirectconsequence<strong>of</strong>theLorentz<strong>in</strong>variant<br />

formulation <strong>of</strong> quantum mechanics. Expand<strong>in</strong>g the relativistic Dirac equation <strong>in</strong> the ratio<br />

<strong>of</strong> the electron velocity and the speed <strong>of</strong> light up to the order (v/c) 2 (the derivation can be<br />

found <strong>in</strong> standard text books like Ref.[Sak67]) one gets<br />

( p<br />

2<br />

2m e0<br />

−eϕ− p4<br />

8m 3 e0 c2 −<br />

e<br />

4m 2 e0<br />

c2σ<br />

·(∇ϕ×p)−<br />

e<br />

)<br />

8m 2 e0 c2∆ϕ ψ = (E −m e0 c 2 )ψ,<br />

with the electrostatic potential ϕ, and the free electron mass m e0 . Our <strong>in</strong>terest concerns<br />

the so called Thomas term −e/(4m 2 e0 c2 )σ·(∇ϕ×p). In atomic physics we assume that the<br />

electric field is a central field, E(r) = −(dϕ/dr)e r which leads to<br />

with ŝ = σ/2.<br />

− e<br />

e<br />

4m 2 ·(∇ϕ×p) = −<br />

e0c2σ (− 1 4m 2 e0 c2 r<br />

e 1<br />

= −<br />

2m 2 e0 c2 r<br />

)<br />

dϕ<br />

σ ·(r×p) (2.1)<br />

dr<br />

dϕ<br />

(2.2)<br />

drŝ·L<br />

≡ λŝ·L, (2.3)<br />

Duetothelattice-periodic potential <strong>in</strong>acrystall<strong>in</strong>esolidthiseffect canhavestrong<strong>in</strong>fluence<br />

6


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 7<br />

Figure 2.1: Schematic representation <strong>of</strong> the band structure <strong>of</strong> GaAs around the Γ po<strong>in</strong>t<br />

(extracted from Ref.[Bur08]).<br />

on the energy band structure. Us<strong>in</strong>g the Kane Model it can be shown that the SOC due<br />

to lattice-periodic potential leads to a lift<strong>in</strong>g <strong>of</strong> degeneracy with a gap ∆ SO = 3/(4λ) (with<br />

λ form Eq.(2.3)) between the Γ 8v states describ<strong>in</strong>g heavy and light holes and Γ 7v states,<br />

as sketched <strong>in</strong> Fig.2.1 where Γ 7v is represented as the split-<strong>of</strong>f band.[DR93, W<strong>in</strong>04a] It is<br />

important to notice that <strong>in</strong> the follow<strong>in</strong>g we do not focus on the SOC which is due to strong<br />

Coulomb potential <strong>of</strong> the atomic core regions, but on the appearance <strong>of</strong> SO effect <strong>in</strong> the<br />

conduction band due to additional external electric field and spatial symmetry break<strong>in</strong>g as<br />

expla<strong>in</strong>ed <strong>in</strong> more detail <strong>in</strong> the next sections. Notice that the effects we present <strong>in</strong> this work<br />

are therefore on a different energy scale: The Pauli splitt<strong>in</strong>g ∆ SO can be large, 0.34 eV <strong>in</strong><br />

GaAs, compared to splitt<strong>in</strong>g at the conduction band which is on the order <strong>of</strong> meV <strong>in</strong> GaAs.<br />

Before we review the sp<strong>in</strong> dynamics <strong>of</strong> conduction electrons and holes <strong>in</strong> semiconductors<br />

and metals, let us first reconsider the sp<strong>in</strong> dynamics <strong>of</strong> a localized sp<strong>in</strong>, as governed by the<br />

Bloch equations.


8 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

2.2 <strong>Dynamics</strong> <strong>of</strong> a Localized <strong>Sp<strong>in</strong></strong><br />

A localized sp<strong>in</strong> ŝ, like a nuclear sp<strong>in</strong>, or the sp<strong>in</strong> <strong>of</strong> a magnetic impurity <strong>in</strong> a solid,<br />

precesses <strong>in</strong> an external magnetic field B due to the Zeeman <strong>in</strong>teraction with Hamiltonian<br />

H Z = −γ g ŝB, where γ g is the correspond<strong>in</strong>g gyromagnetic ratio <strong>of</strong> the nuclear sp<strong>in</strong> or<br />

magnetic impuritysp<strong>in</strong>, respectively, whichwewill set equal toone, unlessneededexplicitly.<br />

This sp<strong>in</strong> dynamics is governed by the Bloch equation <strong>of</strong> a localized sp<strong>in</strong>,<br />

∂ t ŝ = γ g ŝ×B. (2.4)<br />

This equation is identical to the Heisenberg equation ∂ t ŝ = −i[ŝ,H Z ] for the quantum<br />

mechanical sp<strong>in</strong> operator ŝ <strong>of</strong> an S = 1/2-sp<strong>in</strong>, <strong>in</strong>teract<strong>in</strong>g with the external magnetic<br />

field B due to the Zeeman <strong>in</strong>teraction with Hamiltonian H Z . The solution <strong>of</strong> the Bloch<br />

equation for a magnetic field po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the z-direction is ŝ z (t) = ŝ z (0), while the x- and<br />

y- components <strong>of</strong> the sp<strong>in</strong> are precess<strong>in</strong>g with frequency ω 0 = γ g B around the z-axis,<br />

ŝ x (t) = ŝ x (0)cosω 0 t+ŝ y (0)s<strong>in</strong>ω 0 t, ŝ y (t) = −ŝ x (0)s<strong>in</strong>ω 0 t+ŝ y (0)cosω 0 t. S<strong>in</strong>ce a localized<br />

sp<strong>in</strong><strong>in</strong>teracts withits environment by exchange <strong>in</strong>teraction and magnetic dipole<strong>in</strong>teraction,<br />

the precession will dephase after a time τ 2 , and the z-component <strong>of</strong> the sp<strong>in</strong> relaxes to its<br />

equilibrium value s z0 with<strong>in</strong> a relaxation time τ 1 . This modifies the Bloch equations to the<br />

phenomenological equations,<br />

∂ t ŝ x = γ g (ŝ y B z −ŝ z B y )− 1 τ 2<br />

ŝ x<br />

∂ t ŝ y = γ g (ŝ z B x −ŝ x B z )− 1 τ 2<br />

ŝ y<br />

∂ t ŝ z = γ g (ŝ x B y −ŝ y B x )− 1 τ 1<br />

(ŝ z −s z0 ). (2.5)<br />

2.3 <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> <strong>of</strong> <strong>It<strong>in</strong>erant</strong> Electrons<br />

2.3.1 Ballistic <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong><br />

Start<strong>in</strong>g from Dirac equation we have seen that one obta<strong>in</strong>s <strong>in</strong> addition to the<br />

Zeeman term a term which couples the sp<strong>in</strong> s with the momentum p <strong>of</strong> the electrons, the<br />

sp<strong>in</strong>-orbit coupl<strong>in</strong>g<br />

H SO = − µ B<br />

2m e0 c 2ŝ p×E = −ŝB SO(p), (2.6)


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 9<br />

where we set the gyromagnetic ratio γ g = 1. E = −∇ϕ, is an electrical field, and B SO (p) =<br />

µ B /(2m e0 c 2 )p×E. Substitution <strong>in</strong>to the Heisenberg equation yields the Bloch equation <strong>in</strong><br />

the presence <strong>of</strong> sp<strong>in</strong>-orbit <strong>in</strong>teraction:<br />

∂ t ŝ = ŝ×B SO (p), (2.7)<br />

so that the sp<strong>in</strong> performs a precession around the momentum dependent sp<strong>in</strong>-orbit field<br />

B SO (p). It is important to note, that the sp<strong>in</strong>-orbit field does not break the <strong>in</strong>variance<br />

under time reversal ( ŝ → −ŝ,p → −p ), <strong>in</strong> contrast to an external magnetic field B.<br />

Therefore, averag<strong>in</strong>g over all directions <strong>of</strong> momentum, there is no sp<strong>in</strong> polarization <strong>of</strong> the<br />

conduction electrons. However, <strong>in</strong>ject<strong>in</strong>g a sp<strong>in</strong>-polarized electron with given momentum p<br />

<strong>in</strong>to a translationally <strong>in</strong>variant wire, its sp<strong>in</strong> precesses <strong>in</strong> the sp<strong>in</strong>-orbit field as the electron<br />

moves through the wire. The sp<strong>in</strong> will be oriented aga<strong>in</strong> <strong>in</strong> the <strong>in</strong>itial direction after it<br />

moved a length L SO , the sp<strong>in</strong> precession length. The precise magnitude <strong>of</strong> L SO does not<br />

only depend on the strength <strong>of</strong> the sp<strong>in</strong>-orbit <strong>in</strong>teraction but may also depend on the<br />

direction <strong>of</strong> its movement <strong>in</strong> the crystal, as we will discuss below.<br />

2.3.2 <strong>Sp<strong>in</strong></strong> Diffusion Equation<br />

Translational <strong>in</strong>variance isbroken bythepresence<strong>of</strong>disorderdueto impuritiesand<br />

lattice imperfections <strong>in</strong> the conductor. As the electrons scatter from the disorder potential<br />

elastically, their momentum changes <strong>in</strong> a stochastic way, result<strong>in</strong>g <strong>in</strong> diffusive motion. That<br />

results <strong>in</strong> a change <strong>of</strong> the the local electron density ρ(r,t) = ∑ α=± | ψ α(r,t) | 2 , where<br />

α = ± denotes the orientation <strong>of</strong> the electron sp<strong>in</strong>, and ψ α (r,t) is the position and time<br />

dependent electron wave function amplitude. On length scales exceed<strong>in</strong>g the elastic mean<br />

free path l e , that density is governed by the diffusion equation<br />

∂ρ<br />

∂t = D e∇ 2 ρ, (2.8)<br />

where the diffusion constant D e is related to the elastic scatter<strong>in</strong>g time τ by D e = v 2 F τ/d D,<br />

wherev F is the Fermi velocity, and d D the diffusion dimension 1 <strong>of</strong> the electron system. That<br />

diffusion constant is related to the mobility <strong>of</strong> the electrons, µ e = eτ/m e by the E<strong>in</strong>ste<strong>in</strong><br />

relation µ e ρ = e2νD e , where ν is the density <strong>of</strong> states (DOS) per sp<strong>in</strong> at the Fermi energy<br />

E F and m e the effective electron mass.<br />

1 d D can have a fractal value e.g. on quasi-periodic lattices


10 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

Inject<strong>in</strong>g an electron at position r 0 <strong>in</strong>to a conductor with previously constant electrondensity<br />

ρ 0 , thesolution<strong>of</strong> thediffusionequation yieldsthat theelectron density spreads<br />

<strong>in</strong> space accord<strong>in</strong>g to ρ(r,t) = ρ 0 +exp(−(r−r 0 ) 2 /4D e t)/(4πD e t) d D/2 . The dimension d D<br />

is equal to the k<strong>in</strong>etic dimension d, d D = d, if the elastic mean free path l e is smaller than<br />

the size <strong>of</strong> the sample <strong>in</strong> all directions. If the elastic mean free path is larger than the<br />

sample size <strong>in</strong> one direction the diffusion dimension reduces by one, accord<strong>in</strong>gly. Thus, on<br />

average the variance <strong>of</strong> the distance the electron moves after time t is 〈(r−r 0 ) 2 〉 = 2d D D e t.<br />

This <strong>in</strong>troduces a new length scale, the diffusion length L D (t) = √ D e t. We can rewrite the<br />

density as ρ = 〈ψ † (r,t)ψ(r,t)〉, where ψ † = (ψ † + ,ψ† − ) is the two-component vector <strong>of</strong> the<br />

up (+), and down (-) sp<strong>in</strong> fermionic creation operators, and ψ the 2-component vector <strong>of</strong><br />

annihilation operators, respectively, 〈...〉 denotes the expectation value. Accord<strong>in</strong>gly, the<br />

sp<strong>in</strong> density s(r,t) is expected to satisfy a diffusion equation, as well. The sp<strong>in</strong> density is<br />

def<strong>in</strong>ed by<br />

where σ is the vector <strong>of</strong> Pauli matrices,<br />

⎛ ⎞ ⎛<br />

σ x = ⎝ 0 1 ⎠,σ y = ⎝ 0 −i<br />

1 0 i 0<br />

s(r,t) = 1 2 〈ψ† (r,t)σψ(r,t)〉, (2.9)<br />

⎞<br />

⎠, and σ z =<br />

⎛<br />

⎝ 1 0<br />

0 −1<br />

Thus the z-component <strong>of</strong> the sp<strong>in</strong> density is half the difference between the density <strong>of</strong><br />

sp<strong>in</strong> up and down electrons, s z = (ρ + −ρ − )/2, which is the local sp<strong>in</strong> polarization <strong>of</strong> the<br />

electron system. Thus, we can directly <strong>in</strong>fer the diffusion equation for s z , and, similarly, for<br />

the other components <strong>of</strong> the sp<strong>in</strong> density, yield<strong>in</strong>g, without magnetic field and sp<strong>in</strong>-orbit<br />

<strong>in</strong>teraction,[Tor56]<br />

∂s<br />

∂t = D e∇ 2 s− sˆτ . (2.10)<br />

s<br />

Here, <strong>in</strong> the sp<strong>in</strong> relaxation term we <strong>in</strong>troduced the tensor ˆτ s , which can have non-diagonal<br />

matrix elements. In the case <strong>of</strong> a diagonal matrix, τ sxx = τ syy = τ 2 , is the sp<strong>in</strong> dephas<strong>in</strong>g<br />

time, and τ szz = τ 1 the sp<strong>in</strong> relaxation time. The sp<strong>in</strong> diffusion equation can be written as<br />

a cont<strong>in</strong>uity equation for the sp<strong>in</strong> density vector, by def<strong>in</strong><strong>in</strong>g the sp<strong>in</strong> diffusion current <strong>of</strong><br />

the sp<strong>in</strong> components s i ,<br />

⎞<br />

⎠.<br />

J si = −D e ∇s i . (2.11)<br />

Thus, we get the cont<strong>in</strong>uity equation for the sp<strong>in</strong> density components s i ,<br />

∂s i<br />

∂t +∇J s i<br />

= − ∑ j<br />

s j<br />

(ˆτ s ) ij<br />

. (2.12)


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 11<br />

2.3.3 <strong>Sp<strong>in</strong></strong> Orbit Interaction <strong>in</strong> Semiconductors<br />

SOI <strong>in</strong> semiconductors is closely related to break<strong>in</strong>g <strong>of</strong> symmetries which lift sp<strong>in</strong><br />

degeneracy: In the case without magnetic field B we start with a tw<strong>of</strong>old degeneracy: Time<br />

<strong>in</strong>version symmetry, E ↑ (k) = E ↓ (−k) and space <strong>in</strong>version symmetry, E ↑ (k) = E ↑ (−k). As<br />

a consequence we have E ↑ (k) = E ↓ (k). In the follow<strong>in</strong>g we show how the degeneracy is<br />

lifted <strong>in</strong> semiconductor devices <strong>in</strong> case <strong>of</strong> B = 0.<br />

While silicon and germanium have <strong>in</strong> their diamond structure an <strong>in</strong>version symmetry<br />

around every midpo<strong>in</strong>t on each l<strong>in</strong>e connect<strong>in</strong>g nearest neighbor atoms, this is not<br />

the case for III-V-semiconductors like GaAs, InAs, InSb, or ZnS. These have a z<strong>in</strong>c-blende<br />

structure which can be obta<strong>in</strong>ed from a diamond structure with neighbored sites occupied<br />

by the two different elements. Therefore the <strong>in</strong>version symmetry is broken, which results<br />

<strong>in</strong> sp<strong>in</strong>-orbit coupl<strong>in</strong>g. This can be understood by notic<strong>in</strong>g that pairs like Ga-As are local<br />

dipoles whose electric field is responsible for SOC if <strong>in</strong>version symmetry is broken 2 . Similarly,<br />

that symmetry is broken <strong>in</strong> II-VI-semiconductors. This bulk <strong>in</strong>version asymmetry<br />

(BIA) coupl<strong>in</strong>g, or <strong>of</strong>ten so called Dresselhaus-coupl<strong>in</strong>g, is anisotropic, as given by [Dre55]<br />

[<br />

H D = γ D σx k x (ky 2 −k2 z )+σ yk y (kz 2 −k2 x )+σ zk z (kx 2 −k2 y )] , (2.13)<br />

where γ D is the Dresselhaus-sp<strong>in</strong>-orbit coefficient. Band structure calculations yield the<br />

follow<strong>in</strong>g values: γ D = 27.6 eVÅ(GaAs), = 27.2 eVÅ(InAs), = 760.1 eVÅ(InSb) [W<strong>in</strong>03].<br />

Some values extracted <strong>in</strong> experiments are listed <strong>in</strong> Tab.A. Conf<strong>in</strong>ement <strong>in</strong> quantum wells<br />

with width a z on the order <strong>of</strong> the Fermi wave length λ F yields accord<strong>in</strong>gly a sp<strong>in</strong>-orbit<br />

<strong>in</strong>teraction where the momentum <strong>in</strong> growth direction is <strong>of</strong> the order <strong>of</strong> 1/a z . Because<br />

<strong>of</strong> the anisotropy <strong>of</strong> the Dresselhaus term, the sp<strong>in</strong>-orbit <strong>in</strong>teraction depends strongly on<br />

the growth direction <strong>of</strong> the quantum well. Grown <strong>in</strong> [001] direction, one gets, tak<strong>in</strong>g the<br />

expectation value <strong>of</strong> Eq.(2.13) <strong>in</strong> the direction normal to the plane, not<strong>in</strong>g that 〈k z 〉 =<br />

〈kz〉 3 = 0, [Dre55]<br />

H D[001] = α 1 (−σ x k x +σ y k y )+γ D (σ x k x ky 2 −σ y k y kx). 2 (2.14)<br />

where α 1 = γ D 〈kz〉 2 is the l<strong>in</strong>ear Dresselhaus parameter. Thus, <strong>in</strong>sert<strong>in</strong>g an electron with<br />

momentum along the x-direction, with its sp<strong>in</strong> <strong>in</strong>itially polarized <strong>in</strong> z-direction, it will<br />

2 Start<strong>in</strong>g from an extended Kane model where p-like higher energy bands are <strong>in</strong>cluded one gets a Hamiltonian<br />

with matrix elements which are only nonzero if the crystal has no center <strong>of</strong> <strong>in</strong>version.[FMAE + 07]


12 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

precess around the x-axis as it moves along. For narrow quantum wells, where 〈kz 2〉 ∼<br />

1/a 2 z ≥ kF 2 the l<strong>in</strong>ear term exceeds the cubic Dresselhaus terms. For a typical value <strong>of</strong><br />

〈kz 2〉 = 0.036/nm2 , one gets accord<strong>in</strong>gly[W<strong>in</strong>04b], α 1 =0.99 meVnm (GaAs), 0.98 meVnm<br />

(InAs), 27.4 meVnm (InSb). A special situation arises for quantum wells grown <strong>in</strong> the [110]-<br />

direction, where it turns out that the sp<strong>in</strong>-orbit field is po<strong>in</strong>t<strong>in</strong>g normal to the quantum<br />

well, as shown <strong>in</strong> Fig.2.2, so that an electron whose sp<strong>in</strong> is <strong>in</strong>itially polarized along the<br />

normal <strong>of</strong> the plane, rema<strong>in</strong>s polarized as it moves <strong>in</strong> the quantum well.<br />

0 SIA<br />

0<br />

BIA111<br />

0<br />

0<br />

010<br />

0<br />

BIA001<br />

110 0<br />

BIA110<br />

0 001<br />

0<br />

1 10<br />

0<br />

100<br />

Figure 2.2: The sp<strong>in</strong>-orbit vector fields for l<strong>in</strong>ear structure <strong>in</strong>version asymmetry (Rashba)<br />

coupl<strong>in</strong>g, and for l<strong>in</strong>ear bulk <strong>in</strong>version asymmetry (BIA) sp<strong>in</strong>-orbit coupl<strong>in</strong>g for quantum<br />

wells grown <strong>in</strong> [111], [001] and [110] direction, respectively.<br />

In quantum wells with asymmetric electrical conf<strong>in</strong>ement the <strong>in</strong>version symmetry is broken<br />

as well. The result<strong>in</strong>g sp<strong>in</strong>-orbit coupl<strong>in</strong>g, the structural <strong>in</strong>version asymmetry coupl<strong>in</strong>g


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 13<br />

(SIA), also called Rashba-sp<strong>in</strong>-orbit <strong>in</strong>teraction[Ras60] is given by<br />

H R = α 2 (σ x k y −σ y k x ), (2.15)<br />

whereα 2 dependsontheasymmetry<strong>of</strong>theconf<strong>in</strong>ementpotentialV(z)<strong>in</strong>thedirectionz, the<br />

growth direction <strong>of</strong> the quantum well, and can thus be deliberately changed by application<br />

<strong>of</strong> a gate potential. This dependence allows one, <strong>in</strong> pr<strong>in</strong>ciple, to control the electron sp<strong>in</strong><br />

with a gate potential, which can therefore be used as the basis <strong>of</strong> a sp<strong>in</strong> transistor.[DD90]<br />

One should stress that the expectation value <strong>of</strong> the electrical field E c = −∂ z V(z) <strong>in</strong> the<br />

conduction band state vanishes if the effective mass m e is not position dependent. However<br />

it can be shown that the parameter α 2 is modulated by the electric field <strong>in</strong> the valence band<br />

and the z-dependent Pauli splitt<strong>in</strong>g ∆ 0 .[FMAE + 07] Several measured values <strong>of</strong> α 2 are listed<br />

<strong>in</strong> Tab.A and ratios <strong>of</strong> α 2 and α 1 <strong>in</strong> Tab.A.<br />

We can comb<strong>in</strong>e all sp<strong>in</strong>-orbit coupl<strong>in</strong>gs by <strong>in</strong>troduc<strong>in</strong>g the sp<strong>in</strong>-orbit field such that the<br />

Hamiltonian has the form <strong>of</strong> a Zeeman term:<br />

H SO = −sB SO (k), (2.16)<br />

where the sp<strong>in</strong> vector is s = σ/2. But we stress aga<strong>in</strong> that s<strong>in</strong>ce B SO (k) → B SO (−k) =<br />

−B SO (k)underthetimereversaloperation, sp<strong>in</strong>-orbitcoupl<strong>in</strong>gdoesnotbreaktimereversal<br />

symmetry, s<strong>in</strong>ce the time reversal operation also changes the sign <strong>of</strong> the sp<strong>in</strong>, s → −s. Only<br />

an external magnetic field B breaks the time reversal symmetry. Thus, the electron sp<strong>in</strong><br />

operator ŝ is for fixed electron momentum k governed by the Bloch equations with the<br />

sp<strong>in</strong>-orbit field,<br />

∂ŝ<br />

∂t = ŝ×(B+B SO(k))− 1ˆτ s<br />

ŝ. (2.17)<br />

The sp<strong>in</strong> relaxation tensor is no longer necessarily diagonal <strong>in</strong> the presence <strong>of</strong> sp<strong>in</strong>-orbit<br />

<strong>in</strong>teraction as will be shown <strong>in</strong> Sec.2.4.1.<br />

In narrow quantum wells where the cubic Dresselhaus coupl<strong>in</strong>g is weak compared to the<br />

l<strong>in</strong>ear Dresselhaus and Rashba coupl<strong>in</strong>gs, the sp<strong>in</strong>-orbit field is given by<br />

⎛ ⎞<br />

−α 1 k x +α 2 k y<br />

B SO (k) = −2⎜<br />

⎝ α 1 k y −α 2 k x<br />

⎟<br />

⎠ , (2.18)<br />

0<br />

whichchangesbothitsdirectionanditsamplitude| B SO (k) |= 2 √ (α 2 1 +α2 2 )k2 −4α 1 α 2 k x k y ,<br />

as the direction <strong>of</strong> the momentum k is changed. Accord<strong>in</strong>gly, the electron energy dispersion


14 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

close to the Fermi energy is <strong>in</strong> general anisotropic as given by<br />

E ± (k) = 1 √<br />

k 2 ±αk 1−4 α 1α 2<br />

2m e α 2 cosθs<strong>in</strong>θ, (2.19)<br />

where k =| k |, α = √ α 2 1 +α2 2 , and k x = kcosθ. Thus, when an electron is <strong>in</strong>jected with<br />

energy E, with momentum k along the [100]-direction, k x = k,k y = 0, its wave function is a<br />

superposition <strong>of</strong> pla<strong>in</strong> waves with the positive momenta k ± = ∓αm e +m e (α 2 +2E/m e ) 1/2 .<br />

The momentum difference k − − k + = 2m e α causes a rotation <strong>of</strong> the electron eigenstate<br />

<strong>in</strong> the sp<strong>in</strong> subspace. When at x = 0 the electron sp<strong>in</strong> was polarized up sp<strong>in</strong>, with the<br />

eigenvector<br />

⎛<br />

⎞<br />

ψ(x = 0) = ⎝ 1 0<br />

⎠,<br />

then, when its momentum po<strong>in</strong>ts <strong>in</strong> x-direction, at a distance x, it will have rotated the<br />

sp<strong>in</strong> as described by the eigenvector<br />

⎛ ⎞ ⎛<br />

ψ(x) = 1 ⎝ 1 ⎠e ik +x + 1 ⎝<br />

2 α 1 +iα 2 2<br />

α<br />

1<br />

− α 1+iα 2<br />

α<br />

⎞<br />

⎠e ik −x . (2.20)<br />

In Fig.2.3 we plot the correspond<strong>in</strong>g sp<strong>in</strong> density as def<strong>in</strong>ed <strong>in</strong> Eq.(2.9) for pure Rashba<br />

coupl<strong>in</strong>g, α 1 = 0. Thesp<strong>in</strong>will po<strong>in</strong>taga<strong>in</strong> <strong>in</strong>the<strong>in</strong>itial direction, whenthephasedifference<br />

1<br />

sz<br />

0<br />

1<br />

0 L SO 2 L SO<br />

x<br />

Figure 2.3: Precession <strong>of</strong> a sp<strong>in</strong> <strong>in</strong>jected at x = 0, polarized <strong>in</strong> z-direction, as it moves by<br />

one sp<strong>in</strong> precession length L SO = π/m e α through the wire with l<strong>in</strong>ear Rashba sp<strong>in</strong>-orbit<br />

coupl<strong>in</strong>g α 2 .<br />

between the two pla<strong>in</strong> waves is 2π, which gives the condition for sp<strong>in</strong> precession length as


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 15<br />

2π = (k − −k + )L SO , yield<strong>in</strong>g for l<strong>in</strong>ear Rashba and Dresselhaus coupl<strong>in</strong>g, and the electron<br />

mov<strong>in</strong>g <strong>in</strong> [100]- direction,<br />

L SO = π/m e α. (2.21)<br />

We note that the period <strong>of</strong> sp<strong>in</strong> precession changes with the direction <strong>of</strong> the electron momentum<br />

s<strong>in</strong>ce the sp<strong>in</strong>-orbit field, Eq.(2.18), is anisotropic.<br />

2.3.4 <strong>Sp<strong>in</strong></strong> Diffusion <strong>in</strong> the Presence <strong>of</strong> <strong>Sp<strong>in</strong></strong>-Orbit Interaction<br />

As the electrons are scattered by imperfections like impurities and dislocations,<br />

their momentum is changed randomly. Accord<strong>in</strong>gly, the direction <strong>of</strong> the sp<strong>in</strong>-orbit field<br />

B SO (k) changes randomly as the electron moves through the sample. This has two consequences:<br />

the electron sp<strong>in</strong> direction becomes randomized, dephas<strong>in</strong>g the sp<strong>in</strong> precession<br />

and relax<strong>in</strong>g the sp<strong>in</strong> polarization. In addition, the sp<strong>in</strong> precession term is modified, as the<br />

momentum k changes randomly, and has no longer the form given <strong>in</strong> the ballistic Bloch-like<br />

equation, Eq.(2.17). One can derive the diffusion equation for the expectation value <strong>of</strong><br />

the sp<strong>in</strong>, the sp<strong>in</strong> density Eq.(2.9) semiclassically, [MC00, SDGR06] or by diagrammatic<br />

expansion, as will be presented <strong>in</strong> Chapter3. In order to get a better understand<strong>in</strong>g on<br />

the mean<strong>in</strong>g <strong>of</strong> this equation, we will give a simplified classical derivation, <strong>in</strong> the follow<strong>in</strong>g.<br />

The sp<strong>in</strong> density at time t+∆t can be related to the one at the earlier time t. Note that<br />

for ballistic times ∆t ≤ τ, the distance the electron has moved with a probability p ∆x , ∆x,<br />

is related to that time by the ballistic equation, ∆x = k(t)∆t/m when the electron moves<br />

with the momentum k(t). On this time scale the sp<strong>in</strong> evolution is still governed by the<br />

ballistic Bloch equation Eq.(2.17). Thus, we can relate the sp<strong>in</strong> density at the position x<br />

at the time t+∆t, to the one at the earlier time t at position x−∆x:<br />

s(x,t+∆t) = ∑ ∆x<br />

p ∆x<br />

((1− 1ˆτ )<br />

)<br />

∆t s(x−∆x,t)−∆t[B+B SO (k(t))]×s(x−∆x,t) .<br />

s<br />

(2.22)<br />

Now, we can expand <strong>in</strong> ∆t to first order and <strong>in</strong> ∆x to second order. Next, we average over<br />

thedisorderpotential, assum<strong>in</strong>gthat theelectrons arescattered isotropically, andsubstitute<br />

∑<br />

∆x p ∆x... = ∫ (dΩ/Ω)... where Ω is the total angle, and ∫ dΩ denotes the <strong>in</strong>tegral over<br />

all angles with ∫ (dΩ/Ω) = 1. Also, we get (s(x,t+∆t)−s(x,t))/∆t → ∂ t s(x,t) for<br />

∆t → 0, and 〈∆x 2 i 〉 = 2D e∆t, where D e is the diffusion constant. While the disorder<br />

average yields 〈∆x〉 = 0, and 〈B SO (k(t))〉 = 0, separately, for isotropic impurity scatter<strong>in</strong>g,


16 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

averag<strong>in</strong>g their product yields a f<strong>in</strong>ite value, s<strong>in</strong>ce ∆x depends on the momentum at time t,<br />

k(t), yield<strong>in</strong>g 〈∆xB SOi (k(t))〉 = 2∆t〈v F B SOi (k(t))〉, where 〈...〉 denotes the average over<br />

the Fermi surface. This way, we can also evaluate the average <strong>of</strong> the sp<strong>in</strong>-orbit term <strong>in</strong><br />

Eq.(2.22), expanded to first order <strong>in</strong> ∆x, and get, substitut<strong>in</strong>g ∆t → τ the sp<strong>in</strong> diffusion<br />

equation,<br />

∂s<br />

∂t = −B×s+D e∇ 2 s+2τ〈(∇v F )B SO (p)〉×s− 1ˆτ s<br />

s, (2.23)<br />

<strong>Sp<strong>in</strong></strong> polarized electrons <strong>in</strong>jected <strong>in</strong>to the sample spread diffusively, and their sp<strong>in</strong> polarization,<br />

while spread<strong>in</strong>g diffusively as well, decays <strong>in</strong> amplitude exponentially <strong>in</strong> time. S<strong>in</strong>ce,<br />

between scatter<strong>in</strong>g events the sp<strong>in</strong>s precess around the sp<strong>in</strong>-orbit fields, one expects also an<br />

oscillation <strong>of</strong> the polarization amplitude <strong>in</strong> space. One can f<strong>in</strong>d the spatial distribution <strong>of</strong><br />

the sp<strong>in</strong> density which is the solution <strong>of</strong> Eq.(2.23) with the smallest decay rate Γ s . As an<br />

example, the solution for l<strong>in</strong>ear Rashba coupl<strong>in</strong>g is, [SDGR06]<br />

s(x,t) = (ê q cos(qx)+Aê z s<strong>in</strong>(qx))e −t/τs , (2.24)<br />

with 1/τ s = 7/16τ s0 where 1/τ s0 = 2τkF 2α2 2 and where the amplitude <strong>of</strong> the momentum q is<br />

determ<strong>in</strong>ed by D e q 2 = 15/16τ s0 , and A = 3/ √ 15, and ê q = q/q. This solution is plotted <strong>in</strong><br />

Fig.2.4 for ê q = (1,1,0)/ √ 2. We will derive this solution <strong>in</strong> the context <strong>of</strong> local corrections<br />

to the static conductivity, Chapter3. Thereby we show that by add<strong>in</strong>g Dresselhaus SOC it<br />

will be even possible to create persistent solutions.<br />

InFig.2.5 weplotthel<strong>in</strong>early <strong>in</strong>dependentsolution obta<strong>in</strong>ed by <strong>in</strong>terchang<strong>in</strong>g cos ands<strong>in</strong><strong>in</strong><br />

Eq.(2.24), with the sp<strong>in</strong> po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> z-direction, <strong>in</strong>itially. We choose ê q = ê x . Comparison<br />

with the ballistic precession <strong>of</strong> the sp<strong>in</strong>, Fig 2.5 shows that the period <strong>of</strong> precession is<br />

enhanced by the factor 4/ √ 15 <strong>in</strong> the diffusive wire, and that the amplitude <strong>of</strong> the sp<strong>in</strong><br />

density is modulated, chang<strong>in</strong>g from 1 to A = 3/ √ 15.<br />

Inject<strong>in</strong>g a sp<strong>in</strong>-polarized electron at one po<strong>in</strong>t, say x = 0, its density spreads<br />

the same way it does without sp<strong>in</strong>-orbit <strong>in</strong>teraction, ρ(r,t) = exp(−r 2 /4D e t)/(4πD e t) dD/2 ,<br />

where r is the distance to the <strong>in</strong>jection po<strong>in</strong>t. However, the decay <strong>of</strong> the sp<strong>in</strong> density is<br />

periodically modulated as a function <strong>of</strong> 2π √ 15/16r/L SO .[Fro01] The sp<strong>in</strong>-orbit <strong>in</strong>teraction<br />

together with the scatter<strong>in</strong>g from impurities is also a source <strong>of</strong> sp<strong>in</strong> relaxation, as we discuss<br />

<strong>in</strong> the next Section together with other mechanisms <strong>of</strong> sp<strong>in</strong> relaxation. We can f<strong>in</strong>d the<br />

classical sp<strong>in</strong> diffusion current <strong>in</strong> the presence <strong>of</strong> sp<strong>in</strong>-orbit <strong>in</strong>teraction, <strong>in</strong> a similar way<br />

as one can derive the classical diffusion current: The current at the position r is a sum


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 17<br />

S y<br />

0.5 0 0.5<br />

0.5<br />

0.5<br />

0<br />

0<br />

S z<br />

154L SO 2<br />

x<br />

154L SO<br />

Figure 2.4: The sp<strong>in</strong> density for l<strong>in</strong>ear Rashba coupl<strong>in</strong>g which is a solution <strong>of</strong> the sp<strong>in</strong><br />

diffusion equation with the relaxation rate 7/16τ s . The sp<strong>in</strong> po<strong>in</strong>ts <strong>in</strong>itially <strong>in</strong> the x − y-<br />

plane <strong>in</strong> the direction (1,1,0).<br />

over all currents <strong>in</strong> its vic<strong>in</strong>ity which are directed towards that position. Thus, j(r,t) =<br />

〈vρ(r−∆x)〉 where an angular average over all possible directions <strong>of</strong> the velocity v is taken.<br />

Expand<strong>in</strong>g<strong>in</strong>∆x = l e v/v, andnot<strong>in</strong>gthat〈vρ(r)〉 = 0, onegets j(r,t) = 〈v(−∆x)∇ρ(r)〉 =<br />

−(v F l e /2)∇ρ(r) = −D e ∇ρ(r). For the classical sp<strong>in</strong> diffusion current <strong>of</strong> sp<strong>in</strong> component S i ,<br />

as def<strong>in</strong>ed by j Si (r,t) = vS i (r,t), there is the complication that the sp<strong>in</strong> keeps precess<strong>in</strong>g<br />

as it moves from r −∆x to r, and that the sp<strong>in</strong>-orbit field changes its direction with the<br />

direction <strong>of</strong> the electron velocity v. Therefore, the 0-th order term <strong>in</strong> the expansion <strong>in</strong> ∆x<br />

does not vanish, rather, we get<br />

j Si (r,t) = 〈vS k i (r,t)〉−D e∇S i (r,t), (2.25)<br />

where Si k is the part <strong>of</strong> the sp<strong>in</strong> density which evolved from the sp<strong>in</strong> density at r − ∆x<br />

mov<strong>in</strong>g with velocity v and momentum k. Not<strong>in</strong>g that the sp<strong>in</strong> precession on ballistic scales<br />

t ≤ τ is governed by the Bloch equation, Eq.(2.17), we f<strong>in</strong>d by <strong>in</strong>tegration <strong>of</strong> Eq.(2.17),<br />

that Si k = −τ [B SO (k)×S] i<br />

so that we can rewrite the first term yield<strong>in</strong>g the total sp<strong>in</strong><br />

diffusion current as<br />

j Si = −τ〈v F [B SO (k)×S] i<br />

〉−D e ∇S i . (2.26)<br />

Thus, we can rewrite the sp<strong>in</strong> diffusion equation <strong>in</strong> terms <strong>of</strong> this sp<strong>in</strong> diffusion current and


18 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

S<br />

1<br />

1 2<br />

0.89<br />

Sz<br />

0<br />

1 2<br />

0.77<br />

0 L so 2 L so<br />

x<br />

Figure 2.5: The sp<strong>in</strong> density for l<strong>in</strong>ear Rashba coupl<strong>in</strong>g which is a solution <strong>of</strong> the sp<strong>in</strong><br />

diffusion equation with the relaxation rate 1/τ s = 7/16τ s0 . Note that, compared to the<br />

ballistic sp<strong>in</strong> density, Fig.2.3, the period is slightly enhanced by a factor 4/ √ 15. Also, the<br />

amplitude <strong>of</strong> the sp<strong>in</strong> density changes with the position x, <strong>in</strong> contrast to the ballistic case.<br />

The color is chang<strong>in</strong>g <strong>in</strong> proportion to the sp<strong>in</strong> density amplitude.<br />

get the cont<strong>in</strong>uity equation<br />

∂s i<br />

∂t = −∇j S i<br />

+τ〈∇v F (B SO (k)×S) i<br />

〉− 1<br />

(ˆτ s ) ij<br />

s j . (2.27)<br />

It is important to note that <strong>in</strong> contrast to the cont<strong>in</strong>uity equation for the density, there are<br />

two additional terms, due to the sp<strong>in</strong>-orbit <strong>in</strong>teraction. The last one is the sp<strong>in</strong> relaxation<br />

tensor which will be considered <strong>in</strong> detail <strong>in</strong> the next section. The other term arises due<br />

to the fact that Eq.(2.23) conta<strong>in</strong>s a factor 2 <strong>in</strong> front <strong>of</strong> the sp<strong>in</strong>-orbit precession term,<br />

while the sp<strong>in</strong> diffusion current Eq.(2.26) does not conta<strong>in</strong> that factor. This has important<br />

physical consequences, result<strong>in</strong>g <strong>in</strong> the suppression <strong>of</strong> the sp<strong>in</strong> relaxation rate <strong>in</strong> quantum<br />

wires and quantum dots as soon as their lateral extension is smaller than the sp<strong>in</strong> precession<br />

length L SO , as we will see <strong>in</strong> Chapter3.<br />

2.4 <strong>Sp<strong>in</strong></strong> Relaxation Mechanisms<br />

The <strong>in</strong>tr<strong>in</strong>sic sp<strong>in</strong>-orbit <strong>in</strong>teraction itself causes the sp<strong>in</strong> <strong>of</strong> the electrons to precess<br />

coherently, as the electrons move through a conductor, def<strong>in</strong><strong>in</strong>g the sp<strong>in</strong> precession length


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 19<br />

L SO , Eq. (2.21). S<strong>in</strong>ce impurities and dislocations <strong>in</strong> the conductor randomize the electron<br />

momentum, the impurity scatter<strong>in</strong>g is transfered <strong>in</strong>to a randomization <strong>of</strong> the electron sp<strong>in</strong><br />

by the sp<strong>in</strong>-orbit <strong>in</strong>teraction, which thereby results <strong>in</strong> sp<strong>in</strong> dephas<strong>in</strong>g and sp<strong>in</strong> relaxation.<br />

This results <strong>in</strong> a new length scale, the sp<strong>in</strong> relaxation length, L s , which is related to the<br />

sp<strong>in</strong> relaxation rate 1/τ s by<br />

L s = √ D e τ s . (2.28)<br />

2.4.1 D’yakonov-Perel’ <strong>Sp<strong>in</strong></strong> Relaxation<br />

D’yakonov-Perel’ (DP) sp<strong>in</strong> relaxation can be understood qualitatively <strong>in</strong> the follow<strong>in</strong>g<br />

way: The sp<strong>in</strong>-orbit field B SO (k) changes its direction randomly after each elastic<br />

scatter<strong>in</strong>g event from an impurity, that is, after a time <strong>of</strong> the order <strong>of</strong> the elastic scatter<strong>in</strong>g<br />

time τ, when the momentum is changed randomly as sketched <strong>in</strong> Fig. 2.6. Thus, the sp<strong>in</strong><br />

Figure 2.6: Elastic scatter<strong>in</strong>g from impurities changes the direction <strong>of</strong> the sp<strong>in</strong>-orbit field<br />

around which the electron sp<strong>in</strong> is precess<strong>in</strong>g.<br />

has the time τ to perform a precession around the present direction <strong>of</strong> the sp<strong>in</strong>-orbit field,<br />

and can thus change its direction only by an angle <strong>of</strong> the order <strong>of</strong> B SO τ by precession.<br />

After a time t with N t = t/τ scatter<strong>in</strong>g events, the direction <strong>of</strong> the sp<strong>in</strong> will therefore have<br />

changed by an angle <strong>of</strong> the order <strong>of</strong> | B SO |τ √ N t = | B SO | √ τt. Def<strong>in</strong><strong>in</strong>gthe sp<strong>in</strong> relaxation<br />

time τ s as the time by which the sp<strong>in</strong> direction has changed by an angle <strong>of</strong> order one, we<br />

thus f<strong>in</strong>d that 1/τ s ∼ τ〈B SO (k) 2 〉, where the angular brackets denote <strong>in</strong>tegration over all<br />

angles. Remarkably, this sp<strong>in</strong> relaxation rate becomes smaller, the more scatter<strong>in</strong>g events<br />

take place, because the smaller the elastic scatter<strong>in</strong>g time τ is, the less time the sp<strong>in</strong> has<br />

to change its direction by precession. Such a behavior is also well known as motional, or<br />

dynamic narrow<strong>in</strong>g <strong>of</strong> magnetic resonance l<strong>in</strong>es[BPP48].<br />

A more rigorous derivation for the k<strong>in</strong>etic equation <strong>of</strong> the sp<strong>in</strong> density matrix<br />

yields additional <strong>in</strong>terference terms, not taken <strong>in</strong>to account <strong>in</strong> the above argument. It can<br />

be obta<strong>in</strong>ed by iterat<strong>in</strong>g the expansion <strong>of</strong> the sp<strong>in</strong> density Eq. (2.22) once <strong>in</strong> the sp<strong>in</strong>


20 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

precession term, which yields the term<br />

〈<br />

s(x,t)×<br />

∫ ∆t<br />

0<br />

dt ′ B SO (k(t ′ ))×<br />

∫ ∆t<br />

0<br />

〉<br />

dt ′′ B SO (k(t ′′ )) , (2.29)<br />

where 〈...〉 denotes the average over all angles due to the scatter<strong>in</strong>g from impurities. S<strong>in</strong>ce<br />

the electrons move ballistically at times smaller than the elastic scatter<strong>in</strong>g time, the momenta<br />

are correlated only on time scales smaller than τ, yield<strong>in</strong>g<br />

〈k i (t ′ )k j (t ′′ )〉 = (1/2)k 2 δ ij τδ(t ′ −t ′′ ). (2.30)<br />

Not<strong>in</strong>g that (A×B×C) m = ǫ ijk ǫ klm A i B j C l and ∑ ǫ ijk ǫ klm = δ il δ jm −δ im δ jl we f<strong>in</strong>d that<br />

Eq. (2.29) simplifies to − ∑ i (1/τ sij)S j , where the matrix elements <strong>of</strong> the sp<strong>in</strong> relaxation<br />

terms are given by [DP71c],<br />

1<br />

(ˆτ s ) ij<br />

= τ ( 〈B SO (k) 2 〉δ ij −〈B SO (k) i B SO (k) j 〉 ) , (2.31)<br />

where 〈...〉 denotes the average over the direction <strong>of</strong> the momentum k. In Chapter3 we will<br />

focus on this k<strong>in</strong>d <strong>of</strong> sp<strong>in</strong> relaxation and show that these nondiagonal terms can dim<strong>in</strong>ish<br />

the sp<strong>in</strong> relaxation and even result <strong>in</strong> vanish<strong>in</strong>g sp<strong>in</strong> relaxation. In the context <strong>of</strong> weak<br />

localization, which is presented <strong>in</strong> the next Chapter, we will show that the relaxation tensor<br />

Eq.(2.31) can be also derived from Cooperon equation (see AppendixC.3).<br />

2.4.2 DP <strong>Sp<strong>in</strong></strong> Relaxation with Electron-Electron and Electron-Phonon<br />

Scatter<strong>in</strong>g<br />

It has been noted, that the momentum scatter<strong>in</strong>g which limits the D’yakonov-<br />

Perel’ mechanism <strong>of</strong> sp<strong>in</strong> relaxation is not restricted to impurity scatter<strong>in</strong>g, but can also be<br />

due to electron-phonon or electron-electron <strong>in</strong>teractions[GI02, GI04, PF06, DR04]. Thus<br />

the scatter<strong>in</strong>g time τ is the total scatter<strong>in</strong>g time as def<strong>in</strong>ed by, [GI02, GI04], 1/τ = 1/τ 0 +<br />

1/τ ee + 1/τ ep , where 1/τ 0 is the elastic scatter<strong>in</strong>g rate due to scatter<strong>in</strong>g from impurities<br />

∫<br />

with potential V, given by 1/τ 0 = 2πνn i (dθ/2π)(1 − cosθ) | V(k,k ′ ) | 2 , where ν is the<br />

DOS per sp<strong>in</strong> at the Fermi energy, n i is the concentration <strong>of</strong> impurities with potential V,<br />

and kk ′ = kk ′ cos(θ). Concern<strong>in</strong>g the temperature dependence <strong>of</strong> the sp<strong>in</strong> relaxation, for<br />

degenerate electrons <strong>in</strong> semiconductors (E k = E F , with the Fermi energy E F ) it is given<br />

by the temperature dependence <strong>of</strong> τ(T). However, for a non-degenerate statistics one f<strong>in</strong>ds<br />

1/τ s ∼ T 3 τ m (T), where τ m = 〈τ(E k )E k 〉/〈E k 〉.


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 21<br />

2.4.3 Elliott-Yafet <strong>Sp<strong>in</strong></strong> Relaxation<br />

Because <strong>of</strong> the sp<strong>in</strong>-orbit <strong>in</strong>teraction the conduction electron wave functions are<br />

not eigenstates <strong>of</strong> the electron sp<strong>in</strong>, but have an admixture <strong>of</strong> both sp<strong>in</strong> up and sp<strong>in</strong> down<br />

wave functions. Thus, a nonmagnetic impurity potential V can change the electron sp<strong>in</strong>,<br />

by chang<strong>in</strong>g their momentum due to the sp<strong>in</strong>-orbit coupl<strong>in</strong>g. This results <strong>in</strong> another source<br />

<strong>of</strong> sp<strong>in</strong> relaxation which is stronger, the more <strong>of</strong>ten the electrons are scattered, and is<br />

thus proportional to the momentum scatter<strong>in</strong>g rate 1/τ[Ell54, Yaf63]. For degenerate III-V<br />

semiconductors one f<strong>in</strong>ds[Cha75, PT84]<br />

1<br />

τ s<br />

∼<br />

∆ 2 SO<br />

(E G +∆ SO ) 2 E 2 k<br />

E 2 G<br />

1<br />

τ(k) , (2.32)<br />

where E G is the gap between the valence and the conduction band <strong>of</strong> the semiconductor,<br />

E k the energy <strong>of</strong> the conduction electron, and ∆ SO is the sp<strong>in</strong>-orbit splitt<strong>in</strong>g <strong>of</strong> the valence<br />

band. Thus, the Elliott-Yafet sp<strong>in</strong> relaxation (EYS) can be dist<strong>in</strong>guished, be<strong>in</strong>g proportional<br />

to 1/τ, and thereby to the resistivity, <strong>in</strong> contrast to the DP sp<strong>in</strong> scatter<strong>in</strong>g rate, Eq.<br />

(2.31), which is proportional to the conductivity. S<strong>in</strong>ce the EYS decays <strong>in</strong> proportion to<br />

the <strong>in</strong>verse <strong>of</strong> the band gap, it is negligible <strong>in</strong> large band gap semiconductors like Si and<br />

GaAs. The scatter<strong>in</strong>g rate 1/τ is aga<strong>in</strong> the sum <strong>of</strong> the impurity scatter<strong>in</strong>g rate [Ell54], the<br />

electron-phonon scatter<strong>in</strong>g rate [Yaf63, GF97], and electron-electron <strong>in</strong>teraction [Bog80], so<br />

that all these scatter<strong>in</strong>g processes result <strong>in</strong> EY sp<strong>in</strong> relaxation. In degenerate semiconductors<br />

and <strong>in</strong> metals, the electron-electron scatter<strong>in</strong>g rate is given by the Fermi liquid <strong>in</strong>elastic<br />

electron scatter<strong>in</strong>g rate 1/τ ee ∼ T 2 /E F . The electron-phonon scatter<strong>in</strong>g time 1/τ ep ∼ T 5<br />

decays fasterwith temperature. Thus, at low temperatures theElliott-Yafet sp<strong>in</strong>relaxation,<br />

Eq.(2.32), is dom<strong>in</strong>ated by elastic impurity scatter<strong>in</strong>g τ 0 . In non-degenerate semiconductors,<br />

where the Fermi energy is below the conduction band edge, one f<strong>in</strong>ds 1/τ s ∼ T 2 /τ(T).<br />

2.4.4 <strong>Sp<strong>in</strong></strong> Relaxation due to <strong>Sp<strong>in</strong></strong>-Orbit Interaction with Impurities<br />

The sp<strong>in</strong>-orbit <strong>in</strong>teraction, as def<strong>in</strong>ed <strong>in</strong> Eq.(2.6), arises whenever there is a gradient<br />

<strong>in</strong> an electrostatic potential. Thus, the impurity potential gives rise to the sp<strong>in</strong>-orbit<br />

<strong>in</strong>teraction<br />

V SO = 1<br />

2m 2 c2∇V ×k s. (2.33)


22 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

Perturbation theory yields then directly the correspond<strong>in</strong>g sp<strong>in</strong> relaxation rate<br />

1 ∑<br />

∫ dθ<br />

= πνn i<br />

τ s 2π (1−cosθ) | V SO(k,k ′ ) αβ | 2 , (2.34)<br />

α,β<br />

proportional to the concentration <strong>of</strong> impurities n i . Here α,β = ± denotes the sp<strong>in</strong> <strong>in</strong>dices.<br />

S<strong>in</strong>cethesp<strong>in</strong>-orbit<strong>in</strong>teraction <strong>in</strong>creases withtheatomic numberZ <strong>of</strong>theimpurityelement,<br />

this sp<strong>in</strong> relaxation <strong>in</strong>creases as Z 2 , be<strong>in</strong>g stronger for heavier element impurities.<br />

2.4.5 Bir-Aronov-Pikus <strong>Sp<strong>in</strong></strong> Relaxation<br />

Theexchange<strong>in</strong>teraction J betweenelectrons andholes<strong>in</strong>p-dopedsemiconductors<br />

results <strong>in</strong> sp<strong>in</strong> relaxation, as well.[BAP76] Its strength is proportional to the density <strong>of</strong><br />

holes p and depends on their it<strong>in</strong>erancy. If the holes are localized they act like magnetic<br />

impurities. If they are it<strong>in</strong>erant, the sp<strong>in</strong> <strong>of</strong> the conduction electrons is transfered by the<br />

exchange <strong>in</strong>teraction to the holes, where the sp<strong>in</strong>-orbit splitt<strong>in</strong>g <strong>of</strong> the valence bands results<br />

<strong>in</strong> fast sp<strong>in</strong> relaxation <strong>of</strong> the hole sp<strong>in</strong> due to the Elliott-Yafet, or the D’yakonov-Perel’<br />

mechanism.<br />

2.4.6 Magnetic Impurities<br />

Magnetic impurities have a sp<strong>in</strong> S which <strong>in</strong>teracts with the sp<strong>in</strong> <strong>of</strong> the conduction<br />

electrons by the exchange <strong>in</strong>teraction J, result<strong>in</strong>g <strong>in</strong> a spatially and temporarily fluctuat<strong>in</strong>g<br />

local magnetic field<br />

B MI (r) = − ∑ i<br />

Jδ(r−R i )S, (2.35)<br />

where the sum is over the position <strong>of</strong> the magnetic impurities R i . This gives rise to sp<strong>in</strong><br />

relaxation <strong>of</strong> the conduction electrons, with a rate given by<br />

1<br />

τ Ms<br />

= 2πn M νJ 2 S(S +1), (2.36)<br />

where n M is the density <strong>of</strong> magnetic impurities, and ν is the DOS at the Fermi energy.<br />

Here, S is the sp<strong>in</strong> quantum number <strong>of</strong> the magnetic impurity, which can take the values<br />

S = 1/2,1,3/2,2.... Antiferromagnetic exchange <strong>in</strong>teraction between the magnetic impurity<br />

sp<strong>in</strong> and the conduction electrons results <strong>in</strong> a competition between the conduction<br />

electrons to form a s<strong>in</strong>glet with the impurity sp<strong>in</strong>, which results <strong>in</strong> enhanced nonmagnetic<br />

and magnetic scatter<strong>in</strong>g. At low temperatures the magnetic impurity sp<strong>in</strong> is screened by


Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems 23<br />

the conduction electrons result<strong>in</strong>g <strong>in</strong> a vanish<strong>in</strong>g <strong>of</strong> the magnetic scatter<strong>in</strong>g rate. Thus,<br />

the sp<strong>in</strong> scatter<strong>in</strong>g from magnetic impurities has a maximum at a temperature <strong>of</strong> the order<br />

<strong>of</strong> the Kondo temperature T K ∼ E F exp(−1/νJ)[MACR06, MHEZ71, ZBvDA04]. In<br />

semiconductors T K is exponentially small due to the small effective mass m e (see Appendix<br />

A) and the result<strong>in</strong>g small DOS ν. Therefore, the magnetic moments rema<strong>in</strong> free at the<br />

experimentally achievable temperatures. At large concentration <strong>of</strong> magnetic impurities, the<br />

RKKY-exchange <strong>in</strong>teraction between the magnetic impurities quenches however the sp<strong>in</strong><br />

quantum dynamics, so that S(S +1) is replaced by its classical value S 2 . In Mn-p-doped<br />

GaAs, the exchange <strong>in</strong>teraction between the Mn dopants and the holes can result <strong>in</strong> compensation<br />

<strong>of</strong> the hole sp<strong>in</strong>s and therefore a suppression <strong>of</strong> the Bir-Aronov-Pikus (BAP) sp<strong>in</strong><br />

relaxation[ADK + 08].<br />

2.4.7 Nuclear <strong>Sp<strong>in</strong></strong>s<br />

Nuclear sp<strong>in</strong>s <strong>in</strong>teract by the hyperf<strong>in</strong>e <strong>in</strong>teraction with conduction electrons. The<br />

hyperf<strong>in</strong>e <strong>in</strong>teraction between nuclear sp<strong>in</strong>s Î and the conduction electron sp<strong>in</strong>, ŝ, results <strong>in</strong><br />

a local Zeeman field given by [OW53]<br />

ˆB N (r) = − 8π 3<br />

g 0 µ B<br />

∑<br />

γ n Îδ(r−R n ), (2.37)<br />

γ g<br />

whereγ n isthegyromagneticratio<strong>of</strong>thenuclearsp<strong>in</strong>. Thespatialandtemporalfluctuations<br />

<strong>of</strong>thishyperf<strong>in</strong>e<strong>in</strong>teraction fieldresult<strong>in</strong>sp<strong>in</strong>relaxationproportionaltoitsvariance, similar<br />

to the sp<strong>in</strong> relaxation by magnetic impurities.<br />

n<br />

2.4.8 Magnetic Field Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation<br />

The magnetic field changes the electron momentum due to the Lorentz force,<br />

result<strong>in</strong>g <strong>in</strong> a cont<strong>in</strong>uous change <strong>of</strong> the sp<strong>in</strong>-orbit field, which similar to the momentum<br />

scatter<strong>in</strong>g results <strong>in</strong> motional narrow<strong>in</strong>g and thereby a reduction <strong>of</strong> DP sp<strong>in</strong> relaxation:<br />

[Ivc73, PT84, BB04],<br />

1<br />

τ s<br />

∼<br />

τ<br />

1+ωcτ 2 2. (2.38)<br />

Another source <strong>of</strong> a magnetic field dependence is the precession around the external magnetic<br />

field. In bulk semiconductors and for magnetic fields perpendicularto a quantum well,<br />

the orbital mechanism is dom<strong>in</strong>at<strong>in</strong>g, however. This magnetic field dependence can be used<br />

to identify the sp<strong>in</strong> relaxation mechanism, s<strong>in</strong>ce the EYS does have only a weak magnetic


24 Chapter 2: <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong>: Overview and Analysis <strong>of</strong> 2D Systems<br />

field dependence due to the weak Pauli-paramagnetism.<br />

In the next chapter we will also analyze the effect <strong>of</strong> a Zeeman term on the crossover from<br />

weak localization to weak antilocalization. As we will see, this crossover can be controlled<br />

by sp<strong>in</strong> relaxation rate.


Chapter 3<br />

WL/WAL Crossover and <strong>Sp<strong>in</strong></strong><br />

Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

3.1 Introduction<br />

To build sp<strong>in</strong> based devices which rely on coherent sp<strong>in</strong> precession <strong>of</strong> conduction<br />

electrons[DD90, ZFD04], as presented <strong>in</strong> the <strong>in</strong>troduction <strong>of</strong> this work, it has to be analyzed<br />

under which conditions, such as wire geometry, type and <strong>in</strong>tensity <strong>of</strong> SOC, impurity density<br />

etc., sp<strong>in</strong> relaxation rate can be m<strong>in</strong>imized. We have shown <strong>in</strong> Sec.2.4 that if electron<br />

momentum is randomized due to disorder, SO <strong>in</strong>teraction is expected to result not only <strong>in</strong><br />

a sp<strong>in</strong> precession but <strong>in</strong> randomization <strong>of</strong> the electron sp<strong>in</strong> with rate 1/τ s .[DP72] In the<br />

follow<strong>in</strong>g we focus on the D’yakonov-Perel’ sp<strong>in</strong> relaxation mechanism.<br />

3.1.1 One-Dimensional Wires<br />

In one-dimensional wires, whose width W is <strong>of</strong> the order <strong>of</strong> the Fermi wave length<br />

λ F , impurities can only reverse the momentum p → −p. Therefore, the sp<strong>in</strong>-orbit field<br />

can only change its sign, when a scatter<strong>in</strong>g from impurities occurs. B SO (p) → B SO (−p) =<br />

−B SO (p). Therefore, the precession axis and the amplitude <strong>of</strong> the sp<strong>in</strong>-orbit field does not<br />

change, revers<strong>in</strong>g only the sp<strong>in</strong> precession, so that the D’yakonov-Perel’-sp<strong>in</strong> relaxation is<br />

absent <strong>in</strong> one-dimensional wires[KK00, MFA02]. In an external magnetic field, the precession<br />

around the magnetic field axis, due to the Zeeman-<strong>in</strong>teraction is compet<strong>in</strong>g with the<br />

sp<strong>in</strong>-orbit field, however. Then, as the electrons are scattered from impurities, both the<br />

25


26 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

precession axis and the amplitude <strong>of</strong> the total precession field is chang<strong>in</strong>g, s<strong>in</strong>ce<br />

| B+B SO (−p) |=| B−B SO (p) |≠| B+B SO (p) |,<br />

result<strong>in</strong>g <strong>in</strong> sp<strong>in</strong> dephas<strong>in</strong>g and relaxation, as the sign <strong>of</strong> the momentum changes randomly.<br />

3.1.2 Wires with W > λ F<br />

In this Chapter, we show, however, that the condition for a coherent sp<strong>in</strong> precession<br />

is not only the 1d wire, 1/τ s is already strongly reduced <strong>in</strong> much wider wires:<br />

as soon as the wire width W is smaller than bulk sp<strong>in</strong> precession length L SO , which is<br />

the length on which the electron sp<strong>in</strong> precesses a full cycle. This expla<strong>in</strong>s the reduction<br />

<strong>of</strong> the sp<strong>in</strong> relaxation rate <strong>in</strong> quantum wires for widths exceed<strong>in</strong>g both the elastic<br />

mean-free path l e and λ F , as observed with optical[HSM + 06] as well as with WL<br />

measurements[DLS + 05, LSK + 07, SGP + 06, WGZ + 06, KKN09]. As an example we show<br />

two experiments <strong>in</strong> Fig.3.1, where the significant dimensional reduction has been observed.<br />

S<strong>in</strong>ce L SO can be several µm and is not changed significantly as the wire width W is reduced,<br />

such a reduction <strong>of</strong> sp<strong>in</strong> relaxation can be very useful for applications: the sp<strong>in</strong><br />

<strong>of</strong> conduction electrons precesses coherently as it moves along the wire on length scale<br />

L SO . It becomes randomized and relaxes on the longer length scale L s (W) = √ D e τ s only<br />

[D e = v F l e /2 (v F , Fermi velocity) is the 2D diffusion constant].<br />

To understand the connection between the conductivity measurements and sp<strong>in</strong> relaxation<br />

we recall that quantum <strong>in</strong>terference <strong>of</strong> electrons <strong>in</strong> low-dimensional, disordered conductors<br />

is known to result <strong>in</strong> corrections to the electrical conductivity ∆σ. This quantum correction,<br />

the WL effect, is a very sensitive tool to study dephas<strong>in</strong>g and symmetry-break<strong>in</strong>g<br />

mechanisms <strong>in</strong> conductors.[AAKL82, Ber84, CS86] The entanglement <strong>of</strong> sp<strong>in</strong> and charge by<br />

SO <strong>in</strong>teraction reverses the effect <strong>of</strong> WL and thereby enhances the conductivity. This WAL<br />

effect was predicted by Hikami et al.[HLN80] for conductors with impurities <strong>of</strong> heavy elements.<br />

As conduction electrons scatter from such impurities, the SO <strong>in</strong>teraction randomizes<br />

their sp<strong>in</strong>. The result<strong>in</strong>g sp<strong>in</strong> relaxation suppresses <strong>in</strong>terference <strong>in</strong> sp<strong>in</strong> triplet configurations.<br />

S<strong>in</strong>cethetime-reversal operation changes not only thesign <strong>of</strong> momentum but also the<br />

sign <strong>of</strong> the sp<strong>in</strong>, the <strong>in</strong>terference <strong>in</strong> s<strong>in</strong>glet configuration rema<strong>in</strong>s unaffected. S<strong>in</strong>ce s<strong>in</strong>glet<br />

<strong>in</strong>terference reduces the electron’s return probability, it enhances the conductivity, which is<br />

named the WAL effect. In weak magnetic fields, the s<strong>in</strong>glet contributions are suppressed.<br />

Thereby, the conductivity is reduced and the magnetoconductivity becomes negative. The


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 27<br />

magnetoconductivity <strong>of</strong> wires is thus related to the magnitude <strong>of</strong> the sp<strong>in</strong> relaxation rate.<br />

InSec.3.2, we firstderive thequantumcorrections to theconductivity for wires with general<br />

bulk SO <strong>in</strong>teraction and relate it to the Cooperon propagator. In Sec.3.3, we diagonalize<br />

the Cooperon for two-dimensional (2D) electron systems with Rashba SO <strong>in</strong>teraction. We<br />

compare the spectrum <strong>of</strong> the triplet Cooperon with the one <strong>of</strong> the sp<strong>in</strong>-diffusion equation.<br />

InSec.3.4, we presentthesolution <strong>of</strong> the Cooperonequation for awiregeometry. We review<br />

the solutions <strong>of</strong> the sp<strong>in</strong>-diffusion equation <strong>in</strong> the wire geometry and compare the result<strong>in</strong>g<br />

sp<strong>in</strong> relaxation rate with the one extracted from the Cooperon equation. Then we proceed<br />

to calculate the quantum corrections to the conductivity us<strong>in</strong>g the exact diagonalization <strong>of</strong><br />

the Cooperon propagator. In the last part <strong>of</strong> this section, we consider two other k<strong>in</strong>ds <strong>of</strong><br />

boundary conditions. We calculate the sp<strong>in</strong> relaxation rate <strong>in</strong> narrow wires with adiabatic<br />

boundaries, which arise <strong>in</strong> wires with smooth lateral conf<strong>in</strong>ement and regard also tubular<br />

wires. In Sec.3.5, we study the <strong>in</strong>fluence <strong>of</strong> the Zeeman coupl<strong>in</strong>g to a magnetic field perpendicular<br />

to the quantum well <strong>in</strong> a system with sharp boundaries and analyze how the<br />

magnetoconductivity is modified. In Sec.3.6, we draw the conclusions and compare with<br />

experimental results. In AppendixC.2, we give the derivation <strong>of</strong> the non-Abelian Neumann<br />

boundary conditions for the Cooperon propagator. In AppendixC.3, we show the connection<br />

between the effective vector potential A S due to SO coupl<strong>in</strong>g and the sp<strong>in</strong> relaxation<br />

tensor. InAppendixC.4, we give theexact quantumcorrection totheelectrical conductivity<br />

<strong>in</strong> 2D. In AppendixC.5, we detail the diagonalization <strong>of</strong> the Cooperon propagator.<br />

3.2 Quantum Transport Corrections<br />

3.2.1 Diagrammatic Approach<br />

As the temperature is lowered, we expect quantum mechanical coherence to be<br />

more important: The phase coherence length l ϕ <strong>in</strong>creases with decreas<strong>in</strong>g temperature.<br />

If l ϕ is much larger then the elastic scatter<strong>in</strong>g length but smaller then the sample size<br />

one would expect that all <strong>in</strong>terference effects disappear due to self-averag<strong>in</strong>g. However,<br />

it was found that one process seems to survive, as measurements show <strong>in</strong> logarithmically<br />

<strong>in</strong>creas<strong>in</strong>g resistance as temperature decreases, Fig.3.2. In order to <strong>in</strong>troduce the problem<br />

<strong>of</strong> dephas<strong>in</strong>g and WL, we beg<strong>in</strong> with a semiclassical picture <strong>of</strong> how an electron propagates<br />

from a po<strong>in</strong>t r to r ′ : The correspond<strong>in</strong>g probability amplitude P is given as the sum over


28 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

(a)<br />

(b)<br />

Figure 3.1: Two different experimental approaches to extract the wire width dependence<br />

<strong>of</strong> sp<strong>in</strong> relaxation rate. (a) Measurement by Kerr rotation (extracted from [HSM + 06]) and<br />

(b) us<strong>in</strong>g magnetoconductivity experiments (extracted from [LSK + 07]).<br />

all classical paths α with their correspond<strong>in</strong>g actions S α<br />

P(r,r ′ ) ≈ ∑ α<br />

A α e iSα . (3.1)<br />

It is <strong>in</strong>tuitively clear that only this <strong>in</strong>terference processes will survive self-averag<strong>in</strong>g which<br />

are <strong>in</strong>dependent <strong>of</strong> the impurity position. Disadvantageous for conductivity is clearly, if an<br />

electron returns to the po<strong>in</strong>t he started from. This are paths where the relative phase is<br />

<strong>in</strong>dependent <strong>of</strong> the position <strong>of</strong> the impurities:<br />

|P(r,r)| 2 ≈ ∑ α,β<br />

A α A β e i(Sα−S β) . (3.2)<br />

There are two possibilities, which cancel the phase factor: The first one is pure classic,<br />

namely represent<strong>in</strong>g a scatter<strong>in</strong>g that causes the electron to traverse the way α backwards.<br />

The reason for the second one is with the time-reversal-symmetry <strong>of</strong> the system, which<br />

allows to be β the time-reversed path <strong>of</strong> α, more specific<br />

|P(r,r)| 2 ≈ ∑ α<br />

〈|A α| 2 +|A α| 2 〉+2R ∑ α<br />

〈A αA ∗<br />

α 〉 (3.3)<br />

= |P class (r,r)| 2 +|P WL (r,r)| 2 . (3.4)


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 29<br />

Figure 3.2: Measured WL corrections on a th<strong>in</strong> PdAu film, Ref.[DO79]. The resistivity<br />

<strong>in</strong>creases logarithmically as the temperature decreases.<br />

Summ<strong>in</strong>g over all possible paths will generally yield only the classical and the<br />

pure quantum-mechanical term. Other path have <strong>in</strong> general a large phase difference and<br />

will average out. Thus, the quantum-mechanical return-probability is twice the classical<br />

one. This effect is called WL. This effect can be destroyed if we th<strong>in</strong>k about the Aharonov-<br />

Bohm-Effect[AB59]: Switch<strong>in</strong>g on a magnetic field, Fig.(3.3), we can add an additional<br />

phase to the propagat<strong>in</strong>g electron which destroys the constructive <strong>in</strong>terference, lead<strong>in</strong>g to<br />

an enhanced conductivity,<br />

|P(r,r)| 2 ≈ ∑ α<br />

= ∑ α<br />

〈(A αe i2πϕ +A αe −i2πϕ )(A ∗<br />

α e −i2πϕ +A ∗<br />

α e i2πϕ )〉 (3.5)<br />

〈2|A α| 2 (1+cos(4πϕ))〉 (3.6)<br />

with ϕ = φ/φ 0 ,φ 0 = h e .<br />

Because <strong>of</strong> 〈(1 +cos(4πϕ))〉 = 1 we have no quantum corrections on average <strong>in</strong> this case.<br />

The quantum correction to the conductivity appears therefore <strong>in</strong> the difference between the<br />

classical diffusion with the correspond<strong>in</strong>g propagator (Diffuson) and the propagation which<br />

<strong>in</strong>cludes quantum <strong>in</strong>terference between two possibilities for a particle to traverse a closed<br />

loop <strong>in</strong> opposite -time reversed- directions described by the Cooperon. To calculate this<br />

quantum correction at zero frequency ω, one applies Kubo formula (Appendix B.1) which<br />

yields<br />

σ xx (ω = 0) ≡ σ = e2 2π<br />

m 2 e Vol<br />

∫ ∞<br />

0<br />

dE<br />

(<br />

− ∂f(E) )<br />

〈Tr[δ(E −H 0 )p x δ(E −H 0 )p x ]〉<br />

∂E<br />

imp<br />

. (3.7)


30 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

(a)<br />

(b)<br />

Figure 3.3: Exemplification <strong>of</strong> the second term <strong>in</strong> Eq.(3.4): Interference <strong>of</strong> electrons travel<strong>in</strong>g<br />

<strong>in</strong> the opposite direction along the same path causes an enhanced back-scatter<strong>in</strong>g, the<br />

WL effect. (a) Closed electron paths enclose a magnetic flux from an external magnetic<br />

field, <strong>in</strong>dicated as the red arrow, break<strong>in</strong>g time reversal symmetry, break<strong>in</strong>g constructive<br />

<strong>in</strong>terference. (b) The entanglement <strong>of</strong> sp<strong>in</strong> and charge by SO <strong>in</strong>teraction causes the sp<strong>in</strong> to<br />

precess <strong>in</strong>between two scatterers around an axis which changes with the momentum vector<br />

<strong>of</strong> the it<strong>in</strong>erant electron. This effective field can cause WAL.<br />

With the def<strong>in</strong>ition<br />

G R/A<br />

E (p′ ,p) =<br />

〈 ∣ ∣∣∣<br />

p ′ 1<br />

E −H 0 ∓iη<br />

the conductivity can be rewritten to the follow<strong>in</strong>g form<br />

σ =<br />

with the propagator <strong>of</strong> density<br />

〉<br />

∣ p , (3.8)<br />

e 2 ∑<br />

πm 2 p x p ′ x<br />

eVol<br />

×〈GR (p,p ′ )G A (p ′ ,p)〉 imp , (3.9)<br />

p,p ′<br />

Γ(p,p ′ ) = 〈G R (p,p ′ )G A (p ′ ,p)〉 imp , (3.10)<br />

where impurity averag<strong>in</strong>g products <strong>of</strong> Green’s functions <strong>of</strong> the type 〈G R G R 〉 and 〈G A G A 〉<br />

yield small corrections <strong>of</strong> order 1/E F τ and will be neglected (AppendixB.1).<br />

The first approximation one can apply is to assume<br />

〈G R (p,p ′ )G A (p ′ ,p)〉 imp ≈ 〈G R (p)〉 imp 〈G A (p)〉 imp , (3.11)


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 31<br />

where we used the def<strong>in</strong>ition G R/A (p) ≡ G R/A (p,p ′ )δ p,p ′. The average over impurities can<br />

be depicted diagrammatically,<br />

(<br />

〈G〉 imp = + +<br />

) (<br />

+ +<br />

+ +<br />

+ +<br />

)<br />

where the fermion l<strong>in</strong>e<br />

+··· , (3.12)<br />

denotes the unperturbed Green’s function. For uncorrelated<br />

disorder potential, 〈V(x)V(x ′ )〉 = δ(x−x ′ )/2πντ, as we will use <strong>in</strong> the follow<strong>in</strong>g, we perform<br />

the disorder average <strong>in</strong> first-order Born approximation and get<br />

G R E(p) = =<br />

1<br />

E −H 0 (p)+i 1 , (3.13)<br />

2τ<br />

whereG A E (p) is its complex conjugate, respectively. H 0 is the Hamiltonian without disorder<br />

potential V. The impurity vertex (the cross) is given by 1/2πντ. Until now we have<br />

the same <strong>in</strong>formation <strong>in</strong> the scatter<strong>in</strong>g time τ as we would ga<strong>in</strong> from the Drude formula.<br />

Assum<strong>in</strong>g low temperature, we can simplify Eq.(3.9) to<br />

e 2 ∑<br />

σ =<br />

πm 2 e Vol p 2 x ×〈G R (p)〉 imp 〈G A (p)〉 imp (3.14)<br />

p<br />

e 2 ∑<br />

=<br />

πm 2 e Vol p 2 x ×G R (p)G A (p) (3.15)<br />

=<br />

p<br />

e 2 ∑ p 2 x<br />

πm 2 e Vol p (E F −E p ) 2 + ( )<br />

1 2<br />

(3.16)<br />

2τ<br />

which can be simplified <strong>in</strong> the metallic regime, E F ≫ 1/τ, where the dom<strong>in</strong>ant contribution<br />

is given by energies close to E F , to<br />

e 2 ∫ ∞<br />

( ) E2me<br />

≈<br />

πm 2 dE(Volρ(E))<br />

eVol<br />

d<br />

≈ 2 e2<br />

m e<br />

ρ(E F )E F<br />

1<br />

d<br />

0<br />

∫ ∞<br />

∞<br />

dE<br />

1<br />

(E F −E) 2 + ( 1<br />

2τ<br />

1<br />

(E F −E) 2 + ( 1<br />

2τ<br />

) 2<br />

(3.17)<br />

) 2<br />

(3.18)<br />

= e2 nτ<br />

m e<br />

, (3.19)


32 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

where d is the dimension. To get the quantum corrections to the Drude conductivity, we<br />

have to <strong>in</strong>clude the additional contribution by consider<strong>in</strong>g the connection <strong>of</strong>G R andG A<br />

due to the impurity potential V:<br />

〈Γ〉 imp = +<br />

(<br />

+<br />

+<br />

)<br />

+··· (3.20)<br />

This sum can be separated <strong>in</strong>to uncrossed and crossed diagrams. As known from standard<br />

literature both can be calculated <strong>in</strong> an analog way. Summ<strong>in</strong>g up only ladder diagrams will<br />

lead to the Diffuson ˆD<br />

Γ D E,E ′(p,p′ ) =<br />

(<br />

)<br />

δ p,p ′ + + +···<br />

=<br />

1<br />

p+q<br />

(3.21)<br />

1− ∑ q<br />

p ′ +q<br />

= G R E(p)G A E ′(p′ ) 1 τ ˆD E,E ′(p,p ′ ), (3.22)<br />

It is important to notice that each ladder diagram is <strong>of</strong> the same order as the Drude<br />

diagram 1 . In contrast to this classical contribution, the diagrams where the impurity l<strong>in</strong>es,<br />

which connect the advanced and retarded l<strong>in</strong>es, cross are smaller by the factor 1/(p F l)<br />

(see e.g. Ref.[Ram82]). Eq.(3.22) can be solved easily for ˆD if we expand the Diffuson <strong>in</strong><br />

(E ′ −E) and (p ′ −p) (the pole stems from particle conservation):<br />

ˆD E,E ′(p,p ′ ) =<br />

1<br />

i(E −E ′ )+D e (p ′ −p) 2, (3.23)<br />

with the diffusion constant D e = v 2 F τ/d. If Rσ xx is now calculated not only by us<strong>in</strong>g<br />

the bubble diagram, we end up with a correction <strong>of</strong> the momentum relaxation time τ <strong>in</strong><br />

Eq.(3.19) be<strong>in</strong>g replaced by the transport time<br />

∫<br />

τ 0 ∼ dp F |V(p F −p ′ F )|2 (1−p F ·p ′ F ). (3.24)<br />

However, we are <strong>in</strong>terested <strong>in</strong> the calculation which goes beyond this class <strong>of</strong> diagrams.<br />

Time-reversal symmetry helps to sum up the group <strong>of</strong> crossed diagrams via unknott<strong>in</strong>g<br />

1 bubble diagram


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 33<br />

them:<br />

p i<br />

p o<br />

p i<br />

p o<br />

Γ<br />

=<br />

Γ<br />

(3.25)<br />

p o<br />

′<br />

p i<br />

′<br />

p i<br />

′<br />

p o<br />

′<br />

p i<br />

p o<br />

p i<br />

p o<br />

=<br />

Γ<br />

=<br />

Γ<br />

(3.26)<br />

p i<br />

′<br />

p o<br />

′<br />

−p o<br />

′<br />

−p i<br />

′<br />

Exploit<strong>in</strong>g Eq.(3.25) makes the calculation <strong>of</strong> the maximally crossed diagrams easier:<br />

Γ C E,E ′(p,p′ ) = + +<br />

+ ··· (3.27)<br />

=<br />

1<br />

p+q<br />

(3.28)<br />

1− ∑ q<br />

p ′ −q<br />

= G R E (p)G A E ′(p′ ) 1 τĈE,E ′(p,p′ ), (3.29)<br />

with the Cooperon 2 propagator Ĉ for E Fτ ≫ 1 (E F , Fermi energy) given by<br />

Ĉ ω=E−E ′(Q = p+p ′ ) = τ<br />

⎛<br />

⎜<br />

⎝1− ∑ q<br />

E,p + q<br />

E ′ ,p ′ −q<br />

⎞<br />

⎟<br />

⎠<br />

−1<br />

. (3.30)<br />

In contrast to the Diffuson, the <strong>in</strong>frared divergence is now at p = −p ′ , i.e. the correction<br />

to the conductivity for ω = 0,<br />

∆σ = 2 e2<br />

π<br />

1 ∑<br />

m 2 (−p 2 x )G R (p)G A (p)G R (Q−p)G A (Q−p) 1 τĈω=0(Q) (3.31)<br />

p,Q<br />

is due to the factor (−p 2 x), <strong>in</strong> the case without magnetic field and SOC, negative. The<br />

divergent nature expla<strong>in</strong>s post hoc the choice <strong>of</strong> the maximally crossed diagrams.<br />

Notice that one obta<strong>in</strong>s the Cooperon us<strong>in</strong>g time-reversal symmetry from the Diffuson. We<br />

will use this later to map the Cooperon equation onto the sp<strong>in</strong> diffusion equation.<br />

2 The name stems from the s<strong>in</strong>gularity at total momentum be<strong>in</strong>g zero, as <strong>in</strong> the case <strong>of</strong> a Cooper pair<br />

where the consequence is superconductivity.


34 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

3.2.2 Weak Localization <strong>in</strong> Quantum Wires<br />

If the host lattice <strong>of</strong> the electrons provides SO <strong>in</strong>teraction, quantum corrections to<br />

the conductivity have to be calculated <strong>in</strong> the basis <strong>of</strong> eigenstates <strong>of</strong> the Hamiltonian with<br />

SO <strong>in</strong>teraction<br />

H = 1<br />

2m e<br />

(p+eA) 2 +V(x)− 1 2 γ gσ(B+B SO (p)), (3.32)<br />

where m e is the effective electron mass (see AppendixA for examples <strong>in</strong> semiconductors).<br />

A is the vector potential due to the external magnetic field B. B T SO<br />

= (B SOx ,B SOy ) is<br />

the momentum dependent SO field. σ is a vector, with components σ i , i = x,y,z, the<br />

Pauli matrices, γ g is the gyromagnetic ratio with γ g = gµ B with the effective g factor <strong>of</strong><br />

the material, and µ B = e/2m e is the Bohr magneton constant. In Sec.2.3.3 we presented<br />

the dom<strong>in</strong>ant SO <strong>in</strong>teractions <strong>in</strong> semiconductors: For example, the break<strong>in</strong>g <strong>of</strong> <strong>in</strong>version<br />

symmetry <strong>in</strong> III-V semiconductors causes a SO coupl<strong>in</strong>g, which for quantum wells grown<br />

<strong>in</strong> the [001] direction is given by [Dre55]<br />

− 1 2 γ gB SO,D = α 1 (−ê x p x +ê y p y )+γ D (ê x p x p 2 y −ê yp y p 2 x ). (3.33)<br />

Here, α 1 = γ D 〈p 2 z〉 is the l<strong>in</strong>ear Dresselhaus parameter, which measures the strength <strong>of</strong> the<br />

term l<strong>in</strong>ear <strong>in</strong> momenta p x ,p y <strong>in</strong> the plane <strong>of</strong> the 2DES. When 〈p 2 z 〉 ∼ 1/a2 z ≥ k2 F (a z is<br />

the thickness <strong>of</strong> the 2DES and k F is the Fermi wavenumber), that term exceeds the cubic<br />

Dresselhaus terms which have coupl<strong>in</strong>g strength γ D . Asymmetric conf<strong>in</strong>ement <strong>of</strong> the 2DES<br />

yields the Rashba term which does not depend on the growth direction<br />

− 1 2 γ gB SO,R = α 2 (ê x p y −ê y p x ), (3.34)<br />

with α 2 the Rashba parameter.[BR84, Ras60] We consider the standard white-noise model<br />

for the impurity potential, V(x), which vanishes on average 〈V(x)〉 = 0, is uncorrelated,<br />

〈V(x)V(x ′ )〉 = δ(x−x ′ )/2πντ, and weak, E F τ ≫ 1. Go<strong>in</strong>g to momentum (Q) and frequency<br />

(ω) representation, and proceed<strong>in</strong>g as presented Sec.3.2.1 but now tak<strong>in</strong>g <strong>in</strong>to account<br />

the sp<strong>in</strong> degree <strong>of</strong> freedom for a two electron <strong>in</strong>terference, we yield the quantum<br />

correction to the static conductivity as [HLN80]<br />

∆σ = −2 e2 D e<br />

∑<br />

2π Vol<br />

Q<br />

∑<br />

α,β=±<br />

C αββα,ω=0 (Q), (3.35)<br />

where α,β = ± are the sp<strong>in</strong> <strong>in</strong>dices, and the Cooperon propagator Ĉ is for E Fτ ≫ 1 (E F ,<br />

Fermi energy), given by Eq.(3.30). The summation over the sp<strong>in</strong>s is described <strong>in</strong> more


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 35<br />

detail <strong>in</strong> AppendixC.1. Thus, the problem reduces to the calculation <strong>in</strong> presence <strong>of</strong> SOC<br />

<strong>of</strong> the correlation function<br />

∑<br />

q<br />

E,p + q<br />

E ′ ,p ′ −q<br />

= 1<br />

2πντ<br />

∑<br />

GE,σ R (p+q)G E A ′ ,σ ′(p′ −q), (3.36)<br />

q<br />

which simplifies for weak disorder ǫ F τ ≫ 1 to<br />

∫ dΩ 1<br />

≈<br />

2π 1−iτˆΣ , (3.37)<br />

where<br />

ˆΣ = ǫ p ′ +q,σ ′ −ǫ p−q,σ. (3.38)<br />

For diffusive wires, for which the elastic mean-free path l e is smaller than the wire width<br />

W, the <strong>in</strong>tegral is over all angles <strong>of</strong> velocity v on the Fermi surface. Us<strong>in</strong>g<br />

ǫ p = (p+eA)2 − 1 2m e 2 γ gσ(B+B SO (p)),<br />

v = p−q+eA ,<br />

m e<br />

S = 1 2 (σ +σ′ ),<br />

Q = p+p ′ ,<br />

we obta<strong>in</strong> to lowest order <strong>in</strong> Q,<br />

ˆΣ = −v(Q+2eA+2m e âS)+(Q+2eA)âσ ′ + 1 2 γ g(σ ′ −σ)B. (3.39)<br />

Here, the SO coupl<strong>in</strong>gs are comb<strong>in</strong>ed <strong>in</strong> the matrix<br />

⎛<br />

Thus, the Cooperon becomes<br />

Ĉ(Q) −1 = 1 τ<br />

â =<br />

( ∫ dΩ<br />

1−<br />

2π<br />

⎝ −α 1 +γ D ky 2 −α 2<br />

α 2 α 1 −γ D kx<br />

2<br />

1<br />

1+iτ(v(Q+2eA+2m e âS)+H σ ′ +H Z )<br />

⎞<br />

⎠. (3.40)<br />

)<br />

, (3.41)<br />

where H σ ′ = −(Q+2eA)âσ ′ and the Zeeman coupl<strong>in</strong>g to the external magnetic field yields<br />

H Z = − 1 2 γ g(σ ′ −σ)B. (3.42)


36 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

It follows that for weak disorder and without Zeeman coupl<strong>in</strong>g, the Cooperon depends only<br />

on the total momentum Q and the total sp<strong>in</strong> S. Expand<strong>in</strong>g the Cooperon to second order<br />

<strong>in</strong> (Q+2eA+2m e âS) and perform<strong>in</strong>g the angular <strong>in</strong>tegral which is for 2D diffusion (elastic<br />

mean-free path l e smaller than wire width W) cont<strong>in</strong>uous from 0 to 2π and yields<br />

Ĉ(Q) =<br />

1<br />

D e (Q+2eA+2eA S ) 2 +H γD<br />

. (3.43)<br />

The effective vector potential due to SO <strong>in</strong>teraction, A S = m eˆαS/e (where ˆα = 〈â〉 denotes<br />

the matrix Eq.(3.40), as averaged over angle), couples to total sp<strong>in</strong> vector S whose components<br />

are four by four matrices. The cubic Dresselhaus coupl<strong>in</strong>g is found to reduce the<br />

effect <strong>of</strong> the l<strong>in</strong>ear one to<br />

˜α 1 := α 1 −m e γ D E F /2. (3.44)<br />

Furthermore, it gives rise to the sp<strong>in</strong> relaxation term <strong>in</strong> Eq.(3.43),<br />

H γD = D e (m 2 e E Fγ D ) 2 (Sx 2 +S2 y ). (3.45)<br />

In the representation <strong>of</strong> the s<strong>in</strong>glet, |⇄〉 and triplet states |⇉〉,|⇈〉,|〉 (Tab.3.1), Ĉ destate<br />

(<strong>in</strong>dex: electron-number) m s S<br />

|⇄〉 := 1 √<br />

2<br />

(|↑〉 1<br />

|↓〉 2<br />

−|↑〉 2<br />

|↓〉 1<br />

) 0 0<br />

|⇈〉 := |↑〉 1<br />

|↑〉 2<br />

1 1<br />

|⇉〉 := 1 √<br />

2<br />

(|↑〉 1<br />

|↓〉 2<br />

+|↑〉 2<br />

|↓〉 1<br />

) 0 1<br />

|〉 := |↓〉 1<br />

|↓〉 2<br />

−1 1<br />

Table 3.1: S<strong>in</strong>glet and triplet states<br />

couples <strong>in</strong>to a s<strong>in</strong>glet and a triplet sector. Thus, the quantum conductivity is a sum <strong>of</strong><br />

s<strong>in</strong>glet and triplet terms<br />

⎛<br />

∆σ = −2 e2 D e<br />

2π Vol<br />

∑<br />

1<br />

−<br />

⎜ D<br />

Q ⎝ e (Q+2eA)<br />

} {{ 2 + ∑<br />

}<br />

s<strong>in</strong>glet contribution<br />

〈<br />

∣ 〉<br />

∣ ∣∣S<br />

S = 1,m ∣Ĉ(Q) = 1,m . (3.46)<br />

⎟<br />

m=0,±1<br />

⎠<br />

} {{ }<br />

triplet contribution<br />


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 37<br />

With the cut<strong>of</strong>fs due to dephas<strong>in</strong>g 1/τ ϕ and elastic scatter<strong>in</strong>g 1/τ, we can <strong>in</strong>tegrate over all<br />

possible wave vectors Q <strong>in</strong> the 2D case analytically (Appendix C.4).<br />

In 2D, one can treat the magnetic field nonperturbatively us<strong>in</strong>g the basis <strong>of</strong> Landau bands.-<br />

[HLN80, KSZ + 96, MZM + 03, AF01, LG98, Gol05] In wires with widths smaller than cyclotron<br />

length k F l 2 B (l B, the magnetic length, def<strong>in</strong>ed by Bl 2 B<br />

= 1/e), the Landau basis is<br />

not suitable. There is another way to treat magnetic fields: quantum corrections are due<br />

to the <strong>in</strong>terference between closed time-reversed paths. In magnetic fields, the electrons<br />

acquire a magnetic phase, which breaks time-reversal <strong>in</strong>variance. Averag<strong>in</strong>g over all closed<br />

paths, one obta<strong>in</strong>s a rate with which the magnetic field breaks the time-reversal <strong>in</strong>variance,<br />

1/τ B . Like the dephas<strong>in</strong>g rate 1/τ ϕ , it cuts <strong>of</strong>f the divergence aris<strong>in</strong>g from quantum corrections<br />

with small wave vectors Q 2 < 1/D e τ B . In 2D systems, τ B is the time an electron<br />

diffuses along a closed path enclos<strong>in</strong>g one magnetic flux quantum, D e τ B = lB 2 . In wires <strong>of</strong><br />

f<strong>in</strong>ite width W the area which the electron path encloses <strong>in</strong> a time τ B is W √ D e τ B . Requir<strong>in</strong>g<br />

that this encloses one flux quantum gives 1/τ B = D e e 2 W 2 B 2 /3. For arbitrary magnetic<br />

field, the relation<br />

1<br />

τ B<br />

= D e (2e) 2 B 2 〈y 2 〉, (3.47)<br />

with the expectation value <strong>of</strong> the square <strong>of</strong> the transverse position 〈y 2 〉, yields 1/τ B =<br />

(<br />

1−1/(1+W 2 /3lB 2 )) D e /lB 2 . Thus, it is sufficient to diagonalize the Cooperon propagator<br />

as given by Eq.(3.43) without magnetic field, as we will do <strong>in</strong> the next chapters, and to<br />

add the magnetic rate 1/τ B together with dephas<strong>in</strong>g rate 1/τ ϕ to the denom<strong>in</strong>ator <strong>of</strong> Ĉ(Q)<br />

when calculat<strong>in</strong>g the conductivity correction, Eq.(3.46).<br />

3.3 The Cooperon and <strong>Sp<strong>in</strong></strong> Diffusion <strong>in</strong> 2D<br />

The Cooperon can be diagonalized analytically <strong>in</strong> 2D for pure Rashba coupl<strong>in</strong>g,<br />

α 1 = 0,γ D = 0. For this case, we def<strong>in</strong>e the Cooperon Hamilton operator as<br />

H c := Ĉ−1<br />

= Q 2 +2Q SO (Q y S x −Q x S y )+Q 2 SO<br />

D (S2 y +S2 x ), (3.48)<br />

e


38 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

with Q SO = 2m e α 2 = 2π/L SO , where L SO is the sp<strong>in</strong> precession length. In the representation<br />

<strong>of</strong> the s<strong>in</strong>glet |⇄〉 and triplet modes, {|⇈〉,|⇉〉,|〉} it becomes<br />

⎛<br />

H c =<br />

⎜<br />

⎝<br />

⎞<br />

Q 2 0 0 0<br />

√ 0 Q 2 SO +Q2 2Q SO Q + 0<br />

√ √ , (3.49)<br />

0 2Q SO Q − 2Q 2 SO<br />

+Q 2 2Q SO Q +<br />

⎟<br />

⎠<br />

√<br />

0 0 2Q SO Q − Q 2 SO<br />

+Q 2<br />

with Q ± = Q y ±iQ x . Diagonalization yields the gapless s<strong>in</strong>glet eigenvalues and the three<br />

triplet Cooperon eigenvalues with a gap due to the SO coupl<strong>in</strong>g (see Fig.3.4),<br />

(a)<br />

(b)<br />

Figure 3.4: 2D spectrum <strong>of</strong> H c , K i = Q i /Q SO . The physical mean<strong>in</strong>g <strong>of</strong> the gaps <strong>in</strong> the<br />

triplet modes is more comprehensible if H c is related to sp<strong>in</strong> diffusion, where the gaps<br />

appear as sp<strong>in</strong> relaxation rates.<br />

E S (Q)/D e = Q 2 , (3.50)<br />

E T0 (Q)/D e = Q 2 +Q 2 , (3.51)<br />

SO<br />

√<br />

E T± (Q)/D e = Q 2 + 3 2 Q2 SO ± Q2 SO<br />

2<br />

1+16 Q2<br />

Q 2 , (3.52)<br />

SO<br />

where E S denotes the s<strong>in</strong>glet eigenvalue and E T0 ,E T± the three triplet eigenvalues. Notice<br />

that the two m<strong>in</strong>ima <strong>of</strong> the lowest triplet eigenmode are shifted to Q = ±( √ 15/4)Q SO with<br />

a m<strong>in</strong>imal eigenvalue <strong>of</strong> E/D e = (7/16)Q 2 SO<br />

. As we show <strong>in</strong> the follow<strong>in</strong>g, this gap <strong>in</strong> the<br />

triplet modes is directly related to the D’yakonov-Perel’ sp<strong>in</strong> relaxation rate 1/τ s .


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 39<br />

<strong>Sp<strong>in</strong></strong> Diffusion<br />

We can get a better understand<strong>in</strong>g <strong>of</strong> the sp<strong>in</strong> relaxation <strong>in</strong>duced by the SO<br />

coupl<strong>in</strong>g and impurity scatter<strong>in</strong>g by consider<strong>in</strong>g directly the sp<strong>in</strong>-diffusion equation for the<br />

expectation value <strong>of</strong> the electron-sp<strong>in</strong> vector [MC00]<br />

s(r,t) = 1 2 〈ψ† (r,t)σψ(r,t)〉, (3.53)<br />

whereψ † = (ψ † + ,ψ† − )isthetwo-component vector <strong>of</strong> theup(+), anddown(-) sp<strong>in</strong>fermionic<br />

creation operators and ψ the two-component vector <strong>of</strong> annihilation operators, respectively.<br />

In the presence <strong>of</strong> SO coupl<strong>in</strong>g, the sp<strong>in</strong>-diffusion equation becomes for v F | ∇ r s |≪ 1/τ,<br />

0 = ∂ t s+ 1ˆτ s<br />

s−D e ∇ 2 s+γ g (B−2τ〈(∇v F )B SO (p)〉)×s (3.54)<br />

and we def<strong>in</strong>e accord<strong>in</strong>gly the sp<strong>in</strong>-diffusion Hamiltonian H SD<br />

0 = ∂ t s+D e H SD s, (3.55)<br />

where the matrix elements <strong>of</strong> the sp<strong>in</strong> relaxation terms are given by [DP71b, DP71c]<br />

(Appendix C.3)<br />

1<br />

= τγ 2 (<br />

g 〈B SO (k) 2 〉δ ij −〈B SO (k) i B SO (k) j 〉 ) . (3.56)<br />

(ˆτ s ) ij<br />

For pure Rashba SO <strong>in</strong>teraction, the sp<strong>in</strong>-diffusion operator H SD is <strong>in</strong> momentum representation[SDGR06]<br />

⎛<br />

⎞<br />

1<br />

D eτ s<br />

+k 2 0 −i2Q SO k x<br />

H SD = ⎜ 1<br />

⎝ 0<br />

D eτ s<br />

+k 2 −i2Q SO k y<br />

⎟<br />

⎠ , (3.57)<br />

2<br />

i2Q SO k x i2Q SO k y D eτ s<br />

+k 2<br />

with 1/D e τ s = Q 2 SO. In the 2D case, diagonalization yields the eigenvalues<br />

E 0 (k) = k 2 + 1 , (3.58)<br />

D e τ s<br />

√<br />

E ± (k) = k 2 + 3 1<br />

± 1 1+16 k2<br />

2D e τ s 2D e τ s Q 2 . (3.59)<br />

SO<br />

Thus, we f<strong>in</strong>d that the spectrum <strong>of</strong> the sp<strong>in</strong>-diffusion operator and the one <strong>of</strong> the triplet<br />

Cooperon Hamiltonian are identical <strong>in</strong> 2D (Ref. [MCW97]) as long as time-reversal symmetry<br />

is not broken. This confirms that antilocalization <strong>in</strong> the presence <strong>of</strong> SO <strong>in</strong>teraction,


40 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

which has its cause <strong>in</strong> the suppression <strong>of</strong> the triplet modes <strong>in</strong> Eq.(3.46), is <strong>in</strong>deed a direct<br />

measure <strong>of</strong> the sp<strong>in</strong> relaxation. Mathematically, there exists a unitary transformation<br />

H c = U CD H SD U † CD, (3.60)<br />

⎛ ⎞<br />

U CD =<br />

⎜<br />

⎝<br />

− 1 √<br />

2<br />

i √<br />

2<br />

0<br />

0 0 1<br />

1√<br />

2<br />

i √<br />

2<br />

0<br />

⎟<br />

⎠ , (3.61)<br />

with the accord<strong>in</strong>g transformation between sp<strong>in</strong>-density components s i and the triplet components<br />

<strong>of</strong> the Cooperon density ˜s,<br />

1<br />

√<br />

2<br />

(−s x +is y ) = ˜s ⇈ , (3.62)<br />

s z = ˜s ⇉ , (3.63)<br />

1<br />

√ (s x +is y ) = ˜s . 2<br />

(3.64)<br />

This is a consequence <strong>of</strong> the fact that the four-component vector <strong>of</strong> charge density ρ =<br />

(ρ + + ρ − )/2 and sp<strong>in</strong>-density vector S are related to the density vector ˆρ with the four<br />

components 〈ψ † αψ β<br />

〉/ √ 2, where α,β = ±, by a unitary transformation.<br />

Relation to the Diffuson<br />

The classical evolution <strong>of</strong> the four-component density vector ˆρ is by def<strong>in</strong>ition<br />

governed by the diffusion operator, the Diffuson. The Diffuson is related to the Cooperon<br />

<strong>in</strong> momentum space by substitut<strong>in</strong>g Q → p−p ′ and the sum <strong>of</strong> the sp<strong>in</strong>s <strong>of</strong> the retarded<br />

and advanced parts, σ and σ ′ , by their difference. Us<strong>in</strong>g this substitution, Eq.(3.48) leads<br />

thus to the <strong>in</strong>verse <strong>of</strong> the Diffuson propagator<br />

H d :=<br />

ˆD<br />

−1<br />

= Q 2 +2Q SO (Q y˜Sx −Q x˜Sy )+Q 2<br />

D (˜S SO y 2 + ˜S x 2 ), (3.65)<br />

e<br />

with ˜S = (σ ′ −σ)/2, which has the same spectrum as the Cooperon, as long as the timereversal<br />

symmetry is not broken. In the representation <strong>of</strong> s<strong>in</strong>glet and triplet modes the<br />

diffusion Hamiltonian becomes<br />

⎛ √<br />

2Q 2 SO<br />

+Q 2 2Q SO Q − 0 − √ ⎞<br />

2Q SO Q +<br />

√ 2Q SO Q + Q 2 SO<br />

H d =<br />

+Q2 0 0<br />

. (3.66)<br />

⎜<br />

⎝ 0 0 Q 2 0 ⎟<br />

− √ ⎠<br />

2Q SO Q − 0 0 Q 2 SO<br />

+Q 2


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 41<br />

Compar<strong>in</strong>g Eqs. (3.49) and Eq.(3.66), we see that diagonalization leads to Eqs. (3.50)-<br />

(3.52).<br />

Influence <strong>of</strong> Dresselhaus SOC<br />

It can be seen from Eqs. (3.58) and (3.59) that <strong>in</strong> the case <strong>of</strong> a homogeneous<br />

Rashba field, the sp<strong>in</strong>-density has a f<strong>in</strong>ite decay rate as we po<strong>in</strong>ted out <strong>in</strong> Sec.2.3.4. However,<br />

if we go beyond the pureRashba system and <strong>in</strong>clude a l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g, the<br />

first term <strong>in</strong> Eq.(3.33), we can f<strong>in</strong>d sp<strong>in</strong> states which do not relax and are thus persistent.<br />

The sp<strong>in</strong> relaxation tensor, Eq.(3.56), acquires nondiagonal elements and changes to<br />

⎛ ⎞<br />

1<br />

2<br />

1<br />

(k F ) = 4τkF<br />

2 α2 −α 1 α 2 0<br />

⎜ 1<br />

ˆτ s<br />

⎝ −α 1 α 2 2 α2 0 ⎟<br />

⎠ , (3.67)<br />

0 0 α 2<br />

with α = √ α 2 1 +α2 2 . For Q = 0 and α 1 = α 2 = α 0 , we f<strong>in</strong>d <strong>in</strong>deed a vanish<strong>in</strong>g eigenvalue<br />

with a sp<strong>in</strong>-density vector parallel to the sp<strong>in</strong>-orbit field, s = s 0 (1,1,0) T . Moreover there<br />

are two additional modes which do not decay <strong>in</strong> time but are <strong>in</strong>homogeneous <strong>in</strong> space: the<br />

persistent sp<strong>in</strong> helices,[BOZ06, LCCC06, OTA + 99, WOB + 07, KWO + 09]<br />

⎛ ⎞<br />

1 ( )<br />

S = S 0<br />

⎜<br />

⎝ −1 ⎟ 2π<br />

⎠ s<strong>in</strong> (x−y)<br />

L SO<br />

0<br />

⎛ ⎞<br />

0 ( )<br />

√ +S 0 2 ⎜<br />

⎝ 0 ⎟ 2π<br />

⎠ cos (x−y) , (3.68)<br />

L SO<br />

1<br />

(Fig.3.5) and the l<strong>in</strong>early <strong>in</strong>dependent solution, obta<strong>in</strong>ed by <strong>in</strong>terchang<strong>in</strong>g cos and s<strong>in</strong>.<br />

Here, L SO = π/m e<br />

√<br />

2α0 . One has to keep <strong>in</strong> m<strong>in</strong>d that this solution is not an eigenstate<br />

anymore <strong>in</strong> a quantum wire. However, we will show that there exist also long persist<strong>in</strong>g<br />

solutions <strong>in</strong> a quasi-1D case.<br />

It is worth to mention that <strong>in</strong> the case where cubic Dresselhaus coupl<strong>in</strong>g <strong>in</strong> Eq.(3.33)<br />

cannot be neglected, the strength <strong>of</strong> l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g α 1 is shifted[Ket07] to<br />

˜α 1 = α 1 −m e γ D E F /2, as we noted before, Eq.(3.44), and, e.g., <strong>in</strong> the Q = 0 case, the sp<strong>in</strong><br />

relaxation rate becomes<br />

(<br />

1 α<br />

= 2p 2 2<br />

2 − ˜α 2 2<br />

1)<br />

F<br />

τ s α 2 2 + τ +D e (m 2 eE F γ D ) 2 . (3.69)<br />

˜α2 1


42 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

x<br />

0 L SO 2 L SO<br />

S<br />

〈S〉 0<br />

y<br />

0<br />

−S 0<br />

√<br />

2S0<br />

0 〈S〉 z<br />

− √ 2S 0<br />

Figure 3.5: Persistent sp<strong>in</strong> helix solution <strong>of</strong> the sp<strong>in</strong>-diffusion equation for equal magnitude<br />

<strong>of</strong> l<strong>in</strong>ear Rashba and l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g, Eq.(3.68).<br />

The condition for persistence is thus rather ˜α 1 = α 2 . This has been confirmed <strong>in</strong> a recent<br />

measurement (Ref. [KWO + 09]). The existence <strong>of</strong> such long-liv<strong>in</strong>g modes has an effect on<br />

the quantum corrections to the conductivity. In this case, ˜α 1 = α 2 = α 0 , there is only<br />

WL <strong>in</strong> 2D.[PP95, SKK + 08] In the next sections we will make use <strong>of</strong> the equivalence <strong>of</strong> the<br />

triplet sector <strong>of</strong> the Cooperon propagator and the sp<strong>in</strong>-diffusion propagator <strong>in</strong> quantum<br />

wires with appropriate boundary conditions and show how long-liv<strong>in</strong>g modes may change<br />

the quantum corrections to the conductivity.<br />

3.4 Solution <strong>of</strong> the Cooperon Equation <strong>in</strong> Quantum Wires<br />

3.4.1 Quantum Wires with <strong>Sp<strong>in</strong></strong>-Conserv<strong>in</strong>g Boundaries<br />

The conductivity <strong>of</strong> quantum wires with width W < L ϕ = √ D e τ ϕ is without SO<br />

<strong>in</strong>teraction dom<strong>in</strong>ated by the transverse zero-mode Q y = 0. This yields the quasi-1D WL<br />

correction.[KCCC92] However, <strong>in</strong> the presence <strong>of</strong> SO <strong>in</strong>teraction, sett<strong>in</strong>g simply Q y = 0 is<br />

not correct. If we consider sp<strong>in</strong>-conserv<strong>in</strong>g boundaries, rather one has to solve the Cooperon<br />

equation with the follow<strong>in</strong>g modifiedboundaryconditions as derived <strong>in</strong> AppendixC.2(Refs.<br />

[AF01, MFA02]):<br />

(− τ D e<br />

n·〈v F [γ g B SO (k)·S]〉−i∂ n<br />

)<br />

C| ∂S = 0, (3.70)


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 43<br />

where 〈...〉 denotes the average over the direction <strong>of</strong> v F and k which we rewrite us<strong>in</strong>g<br />

Eq.(C.28) for the given geometry as<br />

(−i∂ y +2e(A S ) y<br />

)C<br />

(<br />

x,y = ± W )<br />

= 0, ∀x, (3.71)<br />

2<br />

where n is the unit vector normal to the boundary ∂S and x is the coord<strong>in</strong>ate along the<br />

wire. The transverse zero-mode Q y = 0 does not satisfy this condition. Therefore, it<br />

is convenient to perform a non-Abelian gauge transformation,[AF01, MC00] so that the<br />

transformed problem has Neumann boundary conditions, and the transformed Cooperon<br />

Hamiltonian can therefore be diagonalized <strong>in</strong> zero-mode approximation for quantum wires.<br />

S<strong>in</strong>ce <strong>in</strong> quantum wires these boundary conditions apply only <strong>in</strong> the transverse direction,<br />

a transformation act<strong>in</strong>g <strong>in</strong> the transverse direction is needed: Ĉ → ˜Ĉ = U A ĈU † A , with<br />

U A = exp(i2e(A S ) y<br />

y). Then, the boundary condition simplifies to −i∂ y ˜C(x,y = ±W/2) =<br />

0, ∀x, and the Hamiltonian changes to<br />

˜H c = Q 2 −2Q SO Q x [cos(Q SO y)S y −s<strong>in</strong>(Q SO y)S z ]<br />

+Q 2 SO [cos2 (Q SO y)Sy 2 +s<strong>in</strong>2 (Q SO y)Sz<br />

2<br />

−s<strong>in</strong>(Q SO y)cos(Q SO y)(S y S z +S z S y )] (3.72)<br />

= (Q+2eàs ) 2 . (3.73)<br />

where the effective vector potential A S , as <strong>in</strong>troduced <strong>in</strong> Eq.(3.43),<br />

A S = m e<br />

e ˆαS = m e<br />

e<br />

⎛<br />

⎝ 0 −α 2 0<br />

α 2 0 0<br />

⎛ ⎞<br />

⎞ S x<br />

⎠⎜<br />

⎝ S y<br />

S z<br />

⎟<br />

⎠ , (3.74)<br />

is transformed to the effective vector potential às after the transformation U A has been<br />

applied to the Hamiltonian<br />

à s ≡ m e<br />

e ˜ˆα(y)S<br />

⎛<br />

= m e<br />

e<br />

⎝ 0 −α 2cos(Q SO y) −α 2 s<strong>in</strong>(Q SO y)<br />

0 0 0<br />

⎛<br />

⎞<br />

⎠⎜<br />

⎝<br />

⎞<br />

S x<br />

S y<br />

⎟<br />

⎠ , (3.75)<br />

S z<br />

which varies with the transverse coord<strong>in</strong>ate y on the length scale <strong>of</strong> L SO . Now, we can<br />

see already that for narrow wires W < L SO , this vector potential varies l<strong>in</strong>early with y,


44 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

à s ∼ −m e α 2 Q SO y/e, like the vector potential <strong>of</strong> the external magnetic field B. Thus,<br />

it follows, that for W < L SO , the sp<strong>in</strong> relaxation rate is 1/τ s ∼ Q 2 SO〈y 2 〉 ∼ Q 2 SOW 2 /12,<br />

vanish<strong>in</strong>g for small wire widths. As announced at the beg<strong>in</strong>n<strong>in</strong>g, we thus see that the<br />

presence <strong>of</strong> boundaries dim<strong>in</strong>ishes the sp<strong>in</strong> relaxation already at wire widths <strong>of</strong> the order <strong>of</strong><br />

L SO . If we <strong>in</strong>clude only pure Neumann boundaries to the Hamiltonian H c , i.e., us<strong>in</strong>g the<br />

wrong covariant derivative, this would not affect the absolute sp<strong>in</strong> relaxation m<strong>in</strong>imum and<br />

it would be equal to the nonzero one <strong>in</strong> the 2D case. We give a more precise answer <strong>in</strong> the<br />

follow<strong>in</strong>g.<br />

3.4.2 Zero-Mode Approximation<br />

For W < L ϕ , we can usethe fact that thenthtransversenonzero-modes contribute<br />

terms to the conductivity which are by a factor W/nL ϕ smaller than the 0-mode term, with<br />

n a nonzero <strong>in</strong>teger number. Therefore, it should be a good approximation to diagonalize<br />

the effective quasi-one-dimensional Cooperon propagator, which is the transverse 0-mode<br />

expectation value <strong>of</strong> the transformed <strong>in</strong>verse Cooperon propagator, Eq.(3.73), ˜H1D = 〈0 |<br />

˜H c | 0〉. It is crucial to note that ˜H 1D conta<strong>in</strong>s additional terms, created by the non-Abelian<br />

transformation, whichshowsthat tak<strong>in</strong>gjustthetransversezero-modeapproximation <strong>of</strong> the<br />

untransformed Eq.(3.48) would yield a different, <strong>in</strong>correct result. We can now diagonalize<br />

˜H 1D and f<strong>in</strong>ally f<strong>in</strong>d the dispersion <strong>of</strong> quasi-1D triplet modes<br />

E t0<br />

= Q 2 x + 1 D e 2 Q2 SOt SO , (3.76)<br />

( √<br />

)<br />

E t±<br />

= Q 2 x<br />

D + 1 e 4 Q2 SO<br />

4−t SO ± t 2 +64 Q2 x<br />

SO<br />

Q 2 (1+c SO (c SO −2)) , (3.77)<br />

SO<br />

where c SO and t SO are functions <strong>of</strong> the wire width W as given by<br />

c SO = 1− 2s<strong>in</strong>(Q SOW/2)<br />

, t SO = 1− s<strong>in</strong>(Q SOW)<br />

Q SO W Q SO W . (3.78)<br />

One notices that <strong>in</strong> the limit <strong>of</strong> Q SO W → ∞ we do not recover the previous 2D solution.<br />

This boundary effect will be clarified later on.<br />

Insert<strong>in</strong>g Eq.(3.76) and Eq.(3.77) <strong>in</strong>to the expression for the quantum correction to the<br />

conductivity Eq.(3.46), tak<strong>in</strong>g<strong>in</strong>toaccount themagneticfieldby<strong>in</strong>sert<strong>in</strong>gthemagneticrate<br />

1/τ B (W) and the f<strong>in</strong>ite temperature by <strong>in</strong>sert<strong>in</strong>g the dephas<strong>in</strong>g rate 1/τ ϕ (T), it rema<strong>in</strong>s to<br />

perform the <strong>in</strong>tegral over momentum Q x , as has been done <strong>in</strong> Ref. [Ket07]. For Q SO W < 1,


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 45<br />

the WL correction can then be written as<br />

√ √<br />

HW<br />

∆σ = √<br />

Hϕ +B ∗ (W)/4 −<br />

HW<br />

√<br />

Hϕ +B ∗ (W)/4+H s (W)<br />

√<br />

HW<br />

−2√ Hϕ +B ∗ (W)/4+H s (W)/2<br />

(3.79)<br />

<strong>in</strong> units <strong>of</strong> e 2 /2π. We def<strong>in</strong>ed H W = 1/4eW 2 and the effective external magnetic field<br />

( ) −1<br />

)<br />

B ∗ (W) = 1−<br />

(1+ W2<br />

B. (3.80)<br />

The sp<strong>in</strong> relaxation field H s (W) is for Q SO W < 1,<br />

3l 2 B<br />

H s (W) = 1<br />

12 (Q SOW) 2 H s , (3.81)<br />

suppressed <strong>in</strong> proportion to (W/L SO ) 2 similar to B ∗ (W), Eq.(3.80). Here, H s = 1/4eD e τ s ,<br />

with 1/τ s = 2p 2 F α2 2τ. As mentioned above, the analogy to the suppression <strong>of</strong> the effective<br />

magnetic field, Eq.(3.80), is expected, s<strong>in</strong>ce the SO coupl<strong>in</strong>g enters the transformed<br />

Cooperon, Eq.(3.73), like an effective magnetic vector potential.[Fal03]<br />

Cubic Dresselhaus coupl<strong>in</strong>g, however, would give rise to an additional sp<strong>in</strong> relaxation term,<br />

see Eqs. (3.45) and (C.35), which has no analogy to a magnetic field and is therefore not<br />

suppressed <strong>in</strong> diffusive wires. In Chapter4 we will show that this additional term, though it<br />

cannot vanishfor Q SO W ≪ 1, is widthdependent, s<strong>in</strong>ce theterm Eq.(3.45) <strong>in</strong> theCooperon<br />

Hamiltonian H c is also transformed to obta<strong>in</strong> the modified Neumann boundaries by apply<strong>in</strong>g<br />

the transformation U A , which is W dependent.<br />

When W is larger than SO length L SO , coupl<strong>in</strong>g to higher transverse modes may become<br />

relevant even if W < L ϕ is still satisfied, s<strong>in</strong>ce the SO <strong>in</strong>teraction may <strong>in</strong>troduce coupl<strong>in</strong>g to<br />

highertransversemodes.[Ale06]Wewillstudythesecorrectionsbynumericalexactdiagonalization<strong>in</strong>thenextsection.<br />

Onecanexpectthat<strong>in</strong>ballisticwires,l e > W,thesp<strong>in</strong>relaxation<br />

rate is suppressed <strong>in</strong> analogy to the flux cancellation effect, which yields the weaker rate,<br />

1/τ s (W) = (W/Cl e )(D e W 2 /12L 4 S), where C = 10.8.[BvH88a, DK84, KM02] Before we <strong>in</strong>vestigate<br />

the exact diagonalization <strong>in</strong> the pure Rashba case, we consider an anisotropic field<br />

with l<strong>in</strong>ear Rashba and Dresselhaus SO coupl<strong>in</strong>g to see which form the long persist<strong>in</strong>g sp<strong>in</strong>diffusion<br />

modes have <strong>in</strong> narrow wires. Also, here, we can take advantage <strong>of</strong> the equivalence<br />

<strong>of</strong> Cooperon and sp<strong>in</strong>-diffusion equation as far as time-reversal symmetry is not violated.<br />

We f<strong>in</strong>dthreesolutions whosesp<strong>in</strong>relaxation ratedecay proportional toW 2 forα 2 ≠ α 1 and


46 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

which are persistent for α 2 = α 1 . The first solution is s = s 0 (α 1 ,α 2 ,0) T for Q x = 0 which<br />

is aligned with the effective SO field B SO (k) = −2γ g k x (α 1 ,α 2 ,0) T . In this case, we have<br />

accord<strong>in</strong>g to Eq.(3.81) H s,RD1 (W) = (1/12)( ˜Q SO W) 2 H s , with ˜Q 2 SO = (2m e(α 2 1 −α2 2 ))2 /α 2<br />

and 1/τ s = 2p 2 F α2 τ, α = √ α 2 1 +α2 2 . As mentioned above by transform<strong>in</strong>g the vector potential<br />

A S , Eq.(3.75), this alignment occurs due to the constra<strong>in</strong>t on the sp<strong>in</strong>-dynamics<br />

imposed by the boundary condition as soon as the wire width W is smaller than the sp<strong>in</strong><br />

precession length L SO . In addition, we f<strong>in</strong>d two sp<strong>in</strong> helix solutions <strong>in</strong> narrow wires,<br />

⎛ ⎞ ⎛ ⎞<br />

− α 2<br />

α ( ) 0 ( )<br />

s = s 0<br />

⎜ α 1 ⎟ 2π<br />

⎝ α ⎠ s<strong>in</strong> x +s 0<br />

⎜<br />

L SO<br />

⎝ 0 ⎟ 2π<br />

⎠ cos x , (3.82)<br />

L SO<br />

0<br />

1<br />

and the l<strong>in</strong>early <strong>in</strong>dependent solution, obta<strong>in</strong>ed by <strong>in</strong>terchang<strong>in</strong>g cos and s<strong>in</strong> <strong>in</strong> Eq.(3.82).<br />

The form <strong>of</strong> this long persist<strong>in</strong>g sp<strong>in</strong> helix depends therefore on the ratio <strong>of</strong> l<strong>in</strong>ear Rashba<br />

and l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g strength, Fig.3.6, and its sp<strong>in</strong> relaxation rate is dim<strong>in</strong>ished<br />

as H s,RD2/3 = (1/2)H s,RD1 .<br />

Figure 3.6: Long persist<strong>in</strong>g sp<strong>in</strong> helix solution <strong>of</strong> the sp<strong>in</strong>-diffusion equation <strong>in</strong> a quantum<br />

wire whose width W is smaller than the sp<strong>in</strong> precession length L SO for vary<strong>in</strong>g ratio <strong>of</strong><br />

l<strong>in</strong>ear Rashba α 2 = αs<strong>in</strong>ϕ and l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g, α 1 = αcosϕ, Eq.(3.82), for<br />

fixed α and L SO = π/m e α.<br />

3.4.3 Exact Diagonalization<br />

The exact diagonalization <strong>of</strong> the <strong>in</strong>verse Cooperon propagator, as obta<strong>in</strong>ed after<br />

the non-Abelian transformation, Eq.(3.73), is performed <strong>in</strong> the basis <strong>of</strong> transverse stand<strong>in</strong>g<br />

{<br />

waves, satisfy<strong>in</strong>gNeumannboundaryconditions, 1/ √ W, √ 2/ √ }<br />

W cos((nπ/W)(y −W/2))


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 47<br />

with n ∈ N ∗ , and the plane waves exp(iQ x x) with momentum Q x along the wire. The results<br />

<strong>of</strong> this calculation for different values <strong>of</strong> the dimensionless wire width Q SO W are<br />

shown <strong>in</strong> Fig.3.7. The numerical data po<strong>in</strong>ts are attributed to the different branches <strong>of</strong><br />

the eigenenergy dispersion by compar<strong>in</strong>g their eigenvectors. For small Q SO W, the result is<br />

<strong>in</strong> accordance with the 0-mode approximation: For small wire widths W, the z-component<br />

<strong>of</strong> the total sp<strong>in</strong>, S z , is a good quantum number, as can be seen by expand<strong>in</strong>g Eq.(3.75)<br />

<strong>in</strong> Q SO y. Thus, one can identify the lowest modes with the transverse zero-modes <strong>of</strong> the<br />

triplet modes correspond<strong>in</strong>g to the eigenvalues <strong>of</strong> S z , m = 0,±1, <strong>in</strong> the rotated sp<strong>in</strong> axis<br />

frame, denot<strong>in</strong>g them as E {t0,n=0} and E {t±,n=0} . The m<strong>in</strong>imum <strong>of</strong> the E {t0,n=0} mode is<br />

located at Q x = 0. The m<strong>in</strong>imum <strong>of</strong> E {t−,0} is located at f<strong>in</strong>ite Q x > 0, <strong>in</strong> agreement<br />

with the 0-mode approximation. For larger Q SO W, the modes mix with respect to the sp<strong>in</strong><br />

quantum number and the transverse quantization modes. As a consequence, energy level<br />

cross<strong>in</strong>gs which are present at small wire widths are lifted at larger widths, s<strong>in</strong>ce the mix<strong>in</strong>g<br />

<strong>of</strong> sp<strong>in</strong> and transverse quantization modes results <strong>in</strong> level repulsion be<strong>in</strong>g seen <strong>in</strong> Fig.3.7 as<br />

avoided level cross<strong>in</strong>gs. The branches, E {t0,0} and E {t−,0} , evolve <strong>in</strong>to two modes which become<br />

degenerate at large values <strong>of</strong> Q SO W. These two modes are the only ones whose energy<br />

lies below the energy m<strong>in</strong>imum which we obta<strong>in</strong>ed for the 2D modes, E/D e = (7/16)Q 2 , SO<br />

for a f<strong>in</strong>ite K x -<strong>in</strong>terval around K x = 0. Therefore, we can identify these modes with edge<br />

modes which are created by the Neumann boundary conditions. We can confirm that these<br />

are edge modes by consider<strong>in</strong>g their spatial distribution, shown <strong>in</strong> Fig.3.8. Therefore, even<br />

<strong>in</strong> the limit <strong>of</strong> large widths W, we do get <strong>in</strong> addition to the spectrum obta<strong>in</strong>ed for the 2D<br />

system with open boundary conditions case the edge modes, whose energy is lowered as<br />

seen <strong>in</strong> Fig.3.7. The presence <strong>of</strong> these edge states and the difference to the 2D system with<br />

open boundary conditions can be seen <strong>in</strong> Appendix C.5 <strong>in</strong> the nondiagonal elements which<br />

are proportional to the width times the functions Eqs.(C.65) and (C.66). Even <strong>in</strong> the limit<br />

<strong>of</strong> wide wires there are nondiagonal matrix elements which give a significant contribution<br />

which cannot be neglected. The modes above E/D e = (7/16)Q 2 SO<br />

are extended over the<br />

whole wire system and can thus be characterized as bulk-states, as seen <strong>in</strong> Figs.3.8 (c) and<br />

(d).[Wen07] In Fig.3.9, we compare the results which we obta<strong>in</strong>ed <strong>in</strong> the 0-mode approximation<br />

with the results <strong>of</strong> the exact diagonalization. We plot the absolute m<strong>in</strong>ima <strong>of</strong> the<br />

spectra as function <strong>of</strong> the dimensionless wire width parameter Q SO W/π = 2W/L SO . We<br />

confirm the parabolic suppression <strong>of</strong> the lowest eigenvalues for narrow wires ∼ W 2 /L 2 , SO<br />

obta<strong>in</strong>ed earlier.[MC00, Ket07] We note that the oscillatory behavior <strong>of</strong> the triplet eigen-


48 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

(a)<br />

(b)<br />

5<br />

1.6<br />

4<br />

1.4<br />

1.2<br />

E/DQ 2 SO<br />

3<br />

2<br />

E/DQ 2 SO<br />

1.0<br />

0.8<br />

E {t0,0}<br />

E {t−,0}<br />

1<br />

0.6<br />

0.4<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

K x<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

K x<br />

(c)<br />

(d)<br />

1.0<br />

1.0<br />

0.9<br />

0.9<br />

0.8<br />

0.8<br />

E/DQ 2 SO<br />

0.7<br />

0.6<br />

E/DQ 2 SO<br />

0.7<br />

0.6<br />

0.5<br />

0.5<br />

0.4<br />

0.4<br />

0.3<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

K x<br />

0.3<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

K x<br />

Figure 3.7: Dispersion <strong>of</strong> the triplet Cooperon modes for different dimensionless wire units<br />

Q SO W: (a) Q SO W = 2, (b) Q SO W = 8, (c) Q SO W = 12, (d) Q SO W = 30, plotted as<br />

function <strong>of</strong> K x = Q x /Q SO . For Q SO W ≫ 3, E {t0,0} and E {t−,0} evolve <strong>in</strong>to degenerate<br />

branches for large K x . (For Q SO W = 30, not all high-energy branches are shown.)


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 49<br />

Figure 3.8: Probability density <strong>of</strong> the Cooperon eigenmodes <strong>in</strong> the wire for Q SO W/π = 30.<br />

(a) 3D plot, (b) density plot for one <strong>of</strong> the two lowest branches, show<strong>in</strong>g their edge mode<br />

character. (c) 3D plot and (d) density plot <strong>of</strong> the density <strong>of</strong> the third lowest mode, which<br />

shows bulk character.


50 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

0. 6<br />

E {t−,n>0}<br />

E t0<br />

E {t−,0}<br />

0. 5<br />

E {t+,0}<br />

E/DeQ 2 SO<br />

0. 4<br />

0. 3<br />

0. 2<br />

E {t0,0}<br />

0.1<br />

1<br />

2<br />

3<br />

Q SO W/π<br />

4<br />

5<br />

6<br />

Figure 3.9: Absolute m<strong>in</strong>ima <strong>of</strong> the lowest eigenmodes E {t0,0} , E {t−,0} , and E {t+,0} plotted<br />

as function <strong>of</strong> Q SO W/π = 2W/L SO . We note that the m<strong>in</strong>imum <strong>of</strong> E {t−,0} is located at<br />

±K x ≠ 0. For comparison, the solution <strong>of</strong> the zero-mode approximation E t0 is shown.<br />

values as function <strong>of</strong> W, obta<strong>in</strong>ed <strong>in</strong> the 0-mode approximation,[Ket07] is dim<strong>in</strong>ished accord<strong>in</strong>g<br />

to the exact diagonalization. However, there rema<strong>in</strong>s a sharp maximum <strong>of</strong> E t0 at<br />

Q SO W/π ≈ 1.2 and a shallow maximum <strong>of</strong> E t− at Q SO W/π ≈ 2.5. As noted above, the<br />

values <strong>of</strong> the energy m<strong>in</strong>ima <strong>of</strong> E t0 and E t− at larger widths W are furthermore dim<strong>in</strong>ished<br />

as a result <strong>of</strong> the edge mode character <strong>of</strong> these modes.<br />

Comparison to Solution <strong>of</strong> <strong>Sp<strong>in</strong></strong> Diffusion Equation <strong>in</strong> Quantum Wires<br />

As shown above, the sp<strong>in</strong>-diffusion operator and the triplet Cooperon propagator<br />

have the same eigenvalue spectrum as soon as time-symmetry is not broken. Therefore, the<br />

m<strong>in</strong>ima <strong>of</strong> the sp<strong>in</strong>-diffusion modes, which yield <strong>in</strong>formation on the sp<strong>in</strong> relaxation rate,<br />

must be the same as the one <strong>of</strong> the triplet Cooperon propagator as plotted <strong>in</strong> Fig.3.9. In<br />

Ref. [SDGR06], the value at K x = 0, with K i = Q i /Q SO , has been plotted, as shown <strong>in</strong><br />

Fig.3.10. We note, however, that this does not correspond to the global m<strong>in</strong>imum plotted<br />

<strong>in</strong> Fig.3.9. The two lowest states exhibit two m<strong>in</strong>ima as can be seen <strong>in</strong> Fig.3.7: one local<br />

at K x = 0 and one global, which is for large Q SO W at K x ≈ 0.88. The first one is equal to<br />

the results given by Ref. [SDGR06]. For the WL correction to the conductivity, however,<br />

it is important to reta<strong>in</strong> the global m<strong>in</strong>imum, which is dom<strong>in</strong>ant <strong>in</strong> the <strong>in</strong>tegral over the<br />

longitud<strong>in</strong>al momenta.


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 51<br />

1<br />

0.8<br />

F 1<br />

E/DeQ 2 SO<br />

0.6<br />

0.4<br />

F 2<br />

F 3<br />

0.2<br />

0<br />

0<br />

1 2 3 4 5<br />

Q SO W/π<br />

Figure 3.10: Lowest eigenvalues at K x = 0 plotted aga<strong>in</strong>st Q SO W/π. For comparison, the<br />

global m<strong>in</strong>imum <strong>of</strong> the Cooperon spectrum for Q SO W 9 is plotted, F 3 . Curves F 1 [n]<br />

are given by 7/16 + ( (n/(Q SO W/π)) √ 15/4 ) 2<br />

, n ∈ N. F2 shows the energy m<strong>in</strong>imum <strong>of</strong><br />

the 2D case, F 2 ≡ F 1 [n = 0]. Vertical dotted l<strong>in</strong>es <strong>in</strong>dicate the widths at which the lowest<br />

two branches degenerate at K x = 0. They are given by n/( √ 15/4); consider that the wave<br />

vector for the m<strong>in</strong>imum <strong>of</strong> the E T− mode is ( √ 15/4)Q SO .<br />

Magnetoconductivity<br />

Now, we can proceed to calculate the quantum corrections to the conductivity<br />

us<strong>in</strong>g the exact diagonalization <strong>of</strong> the Cooperon propagator. In Fig.3.11, we show the<br />

result<strong>in</strong>g conductivity as function <strong>of</strong> magnetic field and as function <strong>of</strong> the wire width W.<br />

Here, we have <strong>in</strong>cluded for all wire widths the lowest seven s<strong>in</strong>glet modes and the lowest<br />

21 triplet modes. We choose this number <strong>of</strong> modes so that we <strong>in</strong>cluded sufficient modes to<br />

describe correctly the widest wires considered with Q SO W = 10. Thus, for the considered<br />

low-energy cut<strong>of</strong>f, due to electron dephas<strong>in</strong>g rate 1/τ ϕ <strong>of</strong> 1/D e Q 2 τ SO ϕ = 0.08 and the high<br />

energy cut<strong>of</strong>f 1/D e Q 2 SOτ = 4 due to the elastic scatter<strong>in</strong>g rate, we estimate that seven s<strong>in</strong>glet<br />

modes fall <strong>in</strong> this energy range. S<strong>in</strong>ce for every transverse mode there are one s<strong>in</strong>glet<br />

and three triplet modes, we therefore have to <strong>in</strong>clude 21 triplet modes, accord<strong>in</strong>gly. We<br />

note a change from positive to negative magnetoconductivity as the wire width becomes<br />

smaller than the sp<strong>in</strong> precession length L SO , <strong>in</strong> agreement with the results obta<strong>in</strong>ed with<strong>in</strong><br />

the 0-mode approximation, as reported earlier,[Ket07] plotted for comparison <strong>in</strong> Fig.3.11<br />

(without shad<strong>in</strong>g). At the width, where the crossover occurs, there is a very weak magnetoconductance.<br />

This crossover width W c does depend on the lower cut<strong>of</strong>f, provided by<br />

the temperature-dependent dephas<strong>in</strong>g rate 1/τ ϕ . To estimate the dependence <strong>of</strong> W c on the<br />

dephas<strong>in</strong>g rate, we have to analyze the contribution <strong>of</strong> each term <strong>in</strong> the denom<strong>in</strong>ator <strong>of</strong>


52 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

-1<br />

B/H S<br />

0<br />

1<br />

0.0<br />

-0.5<br />

∆σ ”<br />

-1.0<br />

“<br />

2e 2<br />

2π<br />

2<br />

4<br />

6<br />

Q SO W<br />

8<br />

-<br />

10 1.5<br />

Figure3.11: Thequantumconductivity correction <strong>in</strong>units<strong>of</strong> 2e 2 /2π asfunction<strong>of</strong>magnetic<br />

field B (scaled with bulk relaxation field H s ), and the wire width W scaled with 1/Q SO for<br />

pure Rashba coupl<strong>in</strong>g and cut<strong>of</strong>fs 1/D e Q 2 SOτ ϕ = 0.08, 1/D e Q 2 SOτ = 4: Comparison <strong>of</strong> the<br />

zero-mode calculation (grid without shad<strong>in</strong>g) to the exact diagonalization where the lowest<br />

21 triplet branches and seven s<strong>in</strong>glet branches were taken <strong>in</strong>to account.<br />

s<strong>in</strong>glet and triplet terms <strong>of</strong> the Cooperon. A significant change should arise if<br />

1<br />

(W = W c ) = 1 . (3.83)<br />

τ s τ ϕ<br />

Assum<strong>in</strong>g that this occurs for small wire widths, Q SO W < 1, as confirmed for the parameters<br />

we used, we apply Eq.(3.81) to Eq.(3.83) and conclude that<br />

W c ∼ 1 √ τϕ<br />

. (3.84)<br />

If we calculate the crossover numerically <strong>in</strong> the 0-mode approximation we get the relation<br />

plotted <strong>in</strong> Fig.3.13 which co<strong>in</strong>cides with Eq.(3.84). We note that the change from WAL<br />

to WL may occur at a different width W c than the change <strong>of</strong> sign <strong>in</strong> the correction to the<br />

electrical conductivity ∆σ(B = 0) occurs, W WL . However, we f<strong>in</strong>d that the ratio W c /W WL<br />

is <strong>in</strong>dependent <strong>of</strong> the dephas<strong>in</strong>g rate and the sp<strong>in</strong>-orbit coupl<strong>in</strong>g strength Q SO .<br />

Furthermore while there is quantitative agreement with the 0-mode approximation <strong>in</strong> the<br />

magnitude <strong>of</strong> the magnetoconductivity for all magnetic fields for small wire widths W


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 53<br />

L SO , there is only qualitative agreement at larger wire widths. In particular, the total<br />

magnitude <strong>of</strong> the conductivity is reduced considerably <strong>in</strong> comparison with the 0-mode approximation.<br />

We can attribute this to the reduction <strong>of</strong> the energy <strong>of</strong> the lowest Cooperon<br />

triplet modes due to the emergence <strong>of</strong> edge modes, which is not taken <strong>in</strong>to account when<br />

neglect<strong>in</strong>g transversal spatial variations, as is done <strong>in</strong> the 0-mode approximation. Therefore,<br />

the 0-mode approximation overestimates the suppression <strong>of</strong> the triple modes, result<strong>in</strong>g<br />

<strong>in</strong> an overestimate <strong>of</strong> the conductivity. Similarly, the magnetic field at which the magnetoconductivity<br />

changes its sign from negative to positive is already at a smaller magnetic<br />

field, as seen by the shift <strong>in</strong> the m<strong>in</strong>imum <strong>of</strong> the conductivity towards smaller magnetic<br />

fields (Fig.3.12), <strong>in</strong> comparison to the 0-mode approximation (unshaded) <strong>in</strong> Fig.3.11. This<br />

is <strong>in</strong> accordance with experimental observations, which showed clear deviations from the<br />

0-mode approximation for larger wire widths, with a stronger magnetic field dependence<br />

than obta<strong>in</strong>ed <strong>in</strong> 0-mode approximation.[DLS + 05, LSK + 07, SGP + 06, WGZ + 06] Note that<br />

the nonmonotonous behavior <strong>of</strong> the triplet modes as function <strong>of</strong> the wire width, seen <strong>in</strong><br />

Fig.3.7, cannot be resolved <strong>in</strong> the width dependence <strong>of</strong> the conductivity.<br />

∆σ−∆σ0<br />

2e 2 /2π<br />

0.05<br />

0.00<br />

1<br />

5<br />

Q SO W c<br />

10<br />

1<br />

0<br />

B/H S<br />

Figure 3.12: The relative magnetoconductivity ∆σ(B) − ∆σ(B = 0) <strong>in</strong> units <strong>of</strong> 2e 2 /2π,<br />

with the same parameters and number <strong>of</strong> modes as <strong>in</strong> Fig.3.11.<br />

3.4.4 Other Types <strong>of</strong> Boundary Conditions<br />

Adiabatic Boundary Conditions<br />

When the lateral conf<strong>in</strong>ement potential V is smooth compared to the SO splitt<strong>in</strong>g,<br />

that is, if λ F ∂ y V ≪ ∆ α = 2k F α 2 , where λ F is the Fermi wavelength, the boundaries do<br />

not preserve the sp<strong>in</strong>, s <strong>in</strong> ≠ s out , Eq.(3.70), s<strong>in</strong>ce the sp<strong>in</strong> may adiabatically evolve as the


54 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

Q SO Wc<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.00 0.02 0.04 0.06 0.08 0.10 0.12<br />

1/ ( )<br />

D e Q 2 SOτ ϕ<br />

Figure 3.13: Width <strong>of</strong> wire WQ SO at which there is a crossover from negative to positive<br />

magnetoconductivity as function <strong>of</strong> the lower cut<strong>of</strong>f 1/D e Q 2 SOτ ϕ .<br />

electron is scattered from such a smooth boundary.[GKD04] If this applies, the potential<br />

is adiabatic and the sp<strong>in</strong> <strong>of</strong> the scattered electron stays parallel to the field B SO as its<br />

momentum is changed. This leads to the boundary condition for the sp<strong>in</strong>-density[SDGR06]<br />

s x | y=±W/2 = 0, (3.85)<br />

s y | y=±W/2 = 0, (3.86)<br />

∂ y s z | y=±W/2 = 0. (3.87)<br />

We can transform this boundary condition to the one <strong>of</strong> the triplet Cooperon by us<strong>in</strong>g the<br />

unitary rotation between the sp<strong>in</strong> density <strong>in</strong> the s i representation and the triplet representation<br />

<strong>of</strong> the Cooperon, ˜s i , Eq.(3.62), which leads to the boundary condition<br />

1<br />

√<br />

2<br />

(−s x +is y )| y=±W/2 = ˜s ⇈ | y=±W/2 = 0, (3.88)<br />

∂ y s z | y=±W/2 = ∂ y˜s ⇉ | y=±W/2 = 0, (3.89)<br />

1<br />

√ (s x +is y )| y=±W/2 = ˜s | y=±W/2 = 0.<br />

2<br />

(3.90)<br />

Now, if werequirevanish<strong>in</strong>gmagnetization for the1Dcase, then thediagonalization is done<br />

straightforwardly, plotted <strong>in</strong> Fig.3.14 (see also Ref. [SDGR06]). We use a basis which satisfies<br />

the boundary conditions and therefore consists <strong>of</strong> ∼s<strong>in</strong>(qy)(1,0,0) T , ∼cos(qy)(0,1,0) T ,<br />

and ∼s<strong>in</strong>(qy)(0,0,1) T , with q = nπ/W, n ∈ N ∗ . However, look<strong>in</strong>g at the sp<strong>in</strong>-diffusion<br />

operator [Eq.(3.57)], we see immediately that if we set k to zero and use the fact that<br />

s x,y must vanish at the boundary and s z has to be constant for the chosen k, we receive


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 55<br />

1.0<br />

0.8<br />

EDQ 2 SO<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 5 10 15 20<br />

WQ SO<br />

Figure 3.14: Spectrum <strong>in</strong> the case <strong>of</strong> adiabatic boundaries. Except for states which are<br />

polarized <strong>in</strong> z direction, the relaxation rates for all states diverge at small wire widths,<br />

Q SO W ≪ 1.<br />

a polarized mode. Although this mode is a trivial solution, it differs from all other due to<br />

the fact that it has a f<strong>in</strong>ite sp<strong>in</strong> relaxation time as Q SO W vanishes. For the choice <strong>of</strong> basis<br />

for diagonalization, this means: We set q i = n i π/W for respective ˜s i therewith one state is<br />

described by {n 1 ,n 2 ,n 3 } = {n,n+p 2 ,n +p 3 }, n ∈ N ∗ , p 2 ∈ {−1,0,1,···}, p 3 ∈ N. In<br />

the case {n 1 ,n 2 ,n 3 } = {n,n,n} all branches diverge with reference to the eigenvalues <strong>in</strong> the<br />

limes <strong>of</strong> Q SO W → 0, so that the sp<strong>in</strong> relaxation time goes to zero for small wires.[SDGR06]<br />

In contrast, there is an additional branch <strong>in</strong> the case <strong>of</strong> p 2 = −1,p 3 = 0 which has a f<strong>in</strong>ite<br />

eigenvalue and therefore f<strong>in</strong>ite sp<strong>in</strong> relaxation time for small wire widths<br />

E<br />

D e Q 2 SO<br />

= 2+Kx<br />

(1− 2 32(Q SOW) 2 )<br />

π 4 +O[(Q SO W) 4 ]. (3.91)<br />

The smallest sp<strong>in</strong> relaxation rate for vanish<strong>in</strong>g Q SO W, 1/τ s,1D , which is given for k x = 0, is<br />

foundto bean eigenstate polarized <strong>in</strong> z direction which relaxes with the rate 1/τ s,1D = 2/τ s .<br />

It shows compared with the other modes a monotonous behavior as function <strong>of</strong> Q SO W. If<br />

we allow magnetization for the 1D case, then the comb<strong>in</strong>ation p 1 = −1,p 2 = 0 leads to a<br />

valid solution. For wide wires the smallest absolute m<strong>in</strong>imum is the 2D m<strong>in</strong>imum E/D e =<br />

(7/16)Q 2 ; there are no edge modes. But already at a width <strong>of</strong> Q SO SOW = π/ √ 5 ≈ 1.4 all<br />

modes except the z-polarized exceed the rate 1/τ s,1D .


56 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

Tubular Wires<br />

In tubular wires, such as carbon nanotubes, and InN nanowires <strong>in</strong> which only surface<br />

electrons conduct,[PHC + 09] and radial core-shell InO nanowires,[JLS + 08] the tubular<br />

topology <strong>of</strong> the electron system can be taken <strong>in</strong>to account by periodic boundary conditions.<br />

Inthefollow<strong>in</strong>g, wefocusonwireswherethedom<strong>in</strong>ant SOcoupl<strong>in</strong>gis<strong>of</strong>Rashbatype. Ifone<br />

requires furthermore that this SO-coupl<strong>in</strong>g strength is uniform and the wire curvature can<br />

be neglected,[PHC + 09] the spectrum <strong>of</strong> the Cooperon propagator can be obta<strong>in</strong>ed by substitut<strong>in</strong>g<br />

<strong>in</strong> Eq.(3.52) the transverse momentum Q y by the quantized values Q y = n2π/W,<br />

n is an <strong>in</strong>teger, when W is the circumference <strong>of</strong> the tubular wire. Thus, the sp<strong>in</strong> relaxation<br />

rate rema<strong>in</strong>s unchanged, 1/τ s = (7/16)D e Q 2 SO. If then a magnetic field perpendicular to<br />

the cyl<strong>in</strong>der axis is applied as done <strong>in</strong> Ref. [PHC + 09], there rema<strong>in</strong>s a negative magnetoconductivity<br />

due to the WAL, which is enhanced due to the dimensional crossover from the<br />

2D correction to the conductivity Eq.(C.39) to the quasi-one-dimensional behavior <strong>of</strong> the<br />

quantum correction to the conductivity [Eq.(3.79)]. In tubular wires <strong>in</strong> which the circumference<br />

fulfills the quasi-one-dimensional condition W < L ϕ , the WL correction can then<br />

be written as<br />

∆σ =<br />

√ √<br />

HW<br />

√<br />

Hϕ +B ∗ (W)/4 −<br />

HW<br />

√<br />

Hϕ +B ∗ (W)/4+H s (W)<br />

√<br />

HW<br />

−2√ Hϕ +B ∗ (W)/4+7H s (W)/16<br />

(3.92)<br />

<strong>in</strong> units <strong>of</strong> e 2 /2π. As <strong>in</strong> Eq.(3.79), we def<strong>in</strong>ed H W = 1/4eW 2 , but the effective external<br />

magnetic field differs due to the different geometry: Assum<strong>in</strong>g that W < l B , we have[Ket07]<br />

1<br />

= D e (2e) 2 B 2 〈y 2 〉 (3.93)<br />

τ B<br />

= D e (2e) 21 ( ) BW 2<br />

(3.94)<br />

2 2π<br />

and the effective external magnetic field yields<br />

( ) BW 2<br />

B ∗ (W) = (2e)<br />

(3.95)<br />

2π<br />

= (2e)(Br tube ) 2 , (3.96)


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 57<br />

with the tube radius r tube . The sp<strong>in</strong> relaxation field H s is H s = 1/4eD e τ s , with 1/τ s =<br />

2p 2 F α2 2 τ, or <strong>in</strong> terms <strong>of</strong> the effective Zeeman field B SO,<br />

H s = gγ g<br />

16<br />

B SO (ǫ F ) 2<br />

ǫ F<br />

. (3.97)<br />

Thus the geometrical aspect, 〈y 2 〉 tube /〈y 2 〉 planar ≈ 6.6, might resolve the difference between<br />

measured and calculated SO coupl<strong>in</strong>g strength <strong>in</strong> Ref. [PHC + 09] where a planar geometry<br />

has been assumed to fit the data. This assumption leads <strong>in</strong> a tubular geometry to an<br />

underestimation <strong>of</strong> H s (W). The flux cancellation effect is as long as we are <strong>in</strong> the diffusive<br />

regime, l e ≪ W, negligible.<br />

3.5 Magnetoconductivity with Zeeman splitt<strong>in</strong>g<br />

In the follow<strong>in</strong>g, we want to study if the Zeeman term, Eq.(3.42), is modify<strong>in</strong>g<br />

the magnetoconductivity. Accord<strong>in</strong>gly, we assume that the magnetic field is perpendicular<br />

to the 2DES. Tak<strong>in</strong>g <strong>in</strong>to account the Zeeman term to first order <strong>in</strong> the external magnetic<br />

field B = (0,0,B) T , the Cooperon is accord<strong>in</strong>g to Eq.(3.43) given by<br />

Ĉ(Q) =<br />

1<br />

D e (Q+2eA+2eA S ) 2 +i 1 2 γ g(σ ′ −σ)B . (3.98)<br />

This is valid for magnetic fields γ g B ≪ 1/τ. Due to the term proportional to (σ ′ −σ), the<br />

s<strong>in</strong>gletsector<strong>of</strong>theCooperonmixeswiththetripletone. Wecanf<strong>in</strong>dtheeigenstates <strong>of</strong>C −1 ,<br />

|i〉 with the eigenvalues 1/λ i . Thus, the sum over all sp<strong>in</strong> up and down comb<strong>in</strong>ations αβ,βα<br />

<strong>in</strong> Eq.(3.35) for the conductance correction simplifies <strong>in</strong> the s<strong>in</strong>glet-triplet representation<br />

to (AppendixC.1)<br />

∑<br />

C αββα = ∑ i<br />

αβ<br />

(−〈⇄ |i〉〈i|⇄〉+〈⇈ |i〉〈i|⇈〉<br />

+〈⇉ |i〉〈i|⇉〉+〈 |i〉〈i|〉)λ i . (3.99)<br />

3.5.1 2DEG<br />

The coupl<strong>in</strong>g <strong>of</strong> the s<strong>in</strong>glet to the triplet sector lifts the energy level cross<strong>in</strong>gs at<br />

K = ±1/ √ 2 <strong>of</strong> the s<strong>in</strong>glet E S and the triplet branch E T− as can be seen <strong>in</strong> Fig.3.15 for a<br />

nonvanish<strong>in</strong>g Zeeman coupl<strong>in</strong>g. The spectrum, which is not positive def<strong>in</strong>ite anymore for


58 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

all wave vectors, K = Q/Q SO , is given by<br />

with<br />

E B,2D,1 /D e Q 2 SO<br />

= 1+K 2 x, (3.100)<br />

E B,2D,2 /D e Q 2 SO<br />

= E B,2D,1 + f 1<br />

E B,2D,3/4 /D e Q 2 SO = E B,2D,1 − 1 2<br />

+ 1 6<br />

f 1/3<br />

2<br />

− 1 3 f1/3 2 , (3.101)<br />

(<br />

1±i<br />

√<br />

3<br />

)<br />

f1<br />

f 1/3<br />

2<br />

(<br />

1∓i √ 3<br />

)<br />

f 1/3<br />

2<br />

, (3.102)<br />

f 1 = ˜B 2 −4Kx 2 −1, (3.103)<br />

( √3 √ )<br />

f 2 = 3 108Kx 4 +f3 1 −18K2 x , (3.104)<br />

˜B = gµ B B/D e Q 2 SO, (3.105)<br />

Thus, there are sp<strong>in</strong> states with the same real part <strong>of</strong> the Cooperon energy, so that they<br />

decay equally <strong>in</strong> time, but the imag<strong>in</strong>ary part is different, so that they precess with different<br />

frequencies around the magnetic field axis. A significant change <strong>of</strong> the Cooperon spectrum<br />

appears when gµ B B/D e exceeds Q 2 SO<br />

, as can be seen <strong>in</strong> Fig.3.16(b). All states with a low<br />

decay rate do precess now, due to a f<strong>in</strong>ite imag<strong>in</strong>ary value <strong>of</strong> their eigenvalue. Associated<br />

with this change is also a change <strong>of</strong> the dispersion <strong>of</strong> the real part <strong>of</strong> E B,2D,3 which changes<br />

for K y = 0 from a nearly quadratic dispersion <strong>in</strong> K x , a 0 + a 1 K 2 x for ˜B < 1 to one which<br />

changes more slowly as a 0 +a 1 K 2/3<br />

x +a 2 K 4/3<br />

x +a 3 K 2 x for ˜B ≥ 1 [see Fig.3.16 (a)].<br />

Weak Field<br />

In the case <strong>of</strong> a weak Zeeman field, ˜B ≪ 1, the s<strong>in</strong>glet and triplet sectors are<br />

still approximately separated. A f<strong>in</strong>ite ˜B ≪ 1 lifts however the energy <strong>of</strong> the s<strong>in</strong>glet mode<br />

to E B,2D,2 (K = 0)/D e Q 2 SO<br />

= ˜B 2 /2 + O(˜B 4 ), thus the s<strong>in</strong>glet mode atta<strong>in</strong>s a f<strong>in</strong>ite gap,<br />

correspond<strong>in</strong>g to a f<strong>in</strong>ite relaxation rate. The absolute m<strong>in</strong>imum <strong>of</strong> two <strong>of</strong> the triplet<br />

modes is also lifted by E B,2D,2 (K = ± √ 15/4)/D e Q 2 = 7/16 + (3/4) ˜B 2 SO<br />

+ O(˜B 4 ), while<br />

their value is <strong>in</strong>dependent <strong>of</strong> ˜B at K = 0. In contrast, the m<strong>in</strong>imum <strong>of</strong> the triplet mode<br />

E B,2D,3 , whichapproachesE T+ <strong>in</strong>thelimit<strong>of</strong>nomagneticfield(seeFig.3.4)isdim<strong>in</strong>ishedto<br />

E B,2D,3 (K = 0)/D e Q 2 = 2− ˜B 2 SO<br />

/2+O(˜B 4 ). So, <strong>in</strong> summary, a weak Zeeman field renders<br />

all four Cooperon modes gapfull and that gap can be <strong>in</strong>terpreted as a f<strong>in</strong>ite relaxation rate


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 59<br />

3.0<br />

2.5<br />

R[E]/DeQ 2 SO<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

2 1 0 1 2<br />

0.3<br />

0.2<br />

(a)<br />

K x<br />

I[E]/DeQ 2 SO<br />

0.1<br />

0.0<br />

0.1<br />

0.2<br />

0.3<br />

2 1 0 1 2<br />

(b)<br />

K x<br />

Figure 3.15: (a) Real and (b) imag<strong>in</strong>ary parts <strong>of</strong> the spectrum <strong>of</strong> the 2D Cooperon with<br />

Zeeman term <strong>of</strong> strength gµ B B/D e Q 2 SO<br />

= 0.4. E B,2D,1 (black), E B,2D,2 (red dashed),<br />

E B,2D,3 (green), E B,2D,4 (blue dashed). Dashed vertical l<strong>in</strong>es are located at K x = ±1/ √ 2,<br />

the wave vector where the triplet mode E T− and the s<strong>in</strong>glet mode E S are cross<strong>in</strong>g each<br />

other (without loss <strong>of</strong> generality K y = 0).


60 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

or dephas<strong>in</strong>g rate as the Zeeman coupl<strong>in</strong>g mixes all the sp<strong>in</strong> states, break<strong>in</strong>g time-reversal<br />

<strong>in</strong>variance.<br />

Strong Field<br />

If we expand the spectrum <strong>in</strong> 1/˜B ≪ 1, we f<strong>in</strong>d that all modes have the same gap<br />

proportional to the strength <strong>of</strong> the SO coupl<strong>in</strong>g, D e Q 2 SO<br />

, while two modes atta<strong>in</strong> a f<strong>in</strong>ite<br />

imag<strong>in</strong>ary part with opposite sign<br />

E B,2D,1 /D e Q 2 SO<br />

= 1+K 2 x, (3.106)<br />

E B,2D,2 /D e Q 2 SO = 1+K2 x +O(1/˜B), (3.107)<br />

E B,2D,3/4 /D e Q 2 SO<br />

= 1+K 2 x ∓i ˜B +O(1/˜B). (3.108)<br />

Thus, a strong Zeeman field polarizes the sp<strong>in</strong>s and leads to their precession. The SO<br />

<strong>in</strong>teraction, which is too weak to flip the sp<strong>in</strong>s, merely results <strong>in</strong> a relaxation <strong>of</strong> all modes,<br />

correspond<strong>in</strong>g to a dephas<strong>in</strong>g <strong>of</strong> the sp<strong>in</strong> precession.<br />

3.5.2 Quantum Wire with <strong>Sp<strong>in</strong></strong>-Conserv<strong>in</strong>g Boundary Conditions<br />

In the follow<strong>in</strong>g, we want to study if a Zeeman field modifies the magnetoconductivity<br />

and can shift the crossover from positive to negative Magnetoconductivity as function<br />

<strong>of</strong> wire width W. We have seen that for appropriate parameters the critical width W c is<br />

small compared with L SO . Therefore we stay <strong>in</strong> the 0-mode approximation to get a better<br />

overview <strong>of</strong> the physics. To do so, we first analyze the spectrum.<br />

The modes with low decay rates are situated at K x = 0 and K x ≈ ±1 for small widths and<br />

small enough Zeeman field, ˜B 1, as can be seen <strong>in</strong> Fig.3.17. For Kx = 0, we have<br />

(<br />

E m<strong>in</strong>,0 /D e Q 2 = ˜B 2 SO<br />

+ ˜B 4 1+ (Q SOW) 2 )<br />

(3.109)<br />

12<br />

and for K x = ±1,<br />

E m<strong>in</strong>,±1 /D e Q 2 SO<br />

= (Q SOW) 2 (<br />

+<br />

24<br />

˜B 1 2 2 − (Q SOW) 2 )<br />

96<br />

(<br />

+ ˜B 3 4<br />

16 − (Q SOW) 2<br />

384<br />

)<br />

. (3.110)<br />

As <strong>in</strong> the 2D case, we have a mode which is <strong>in</strong>dependent <strong>of</strong> the Zeeman field and the<br />

spectrum is equal to E t0 with the eigenvector (0,1,0,1) T . Us<strong>in</strong>g this spectrum, we estimate


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 61<br />

4<br />

3<br />

R[E]/DeQ 2 SO<br />

2<br />

1<br />

0<br />

2 1 0 1 2<br />

(a)<br />

K x<br />

1.5<br />

1.0<br />

I[E]/DeQ 2 SO<br />

0.5<br />

0.0<br />

0.5<br />

1.0<br />

1.5<br />

2 1 0 1 2<br />

K x<br />

(b)<br />

˜B<br />

0 1 2<br />

Figure 3.16: (a) Real and (b) imag<strong>in</strong>ary parts <strong>of</strong> the spectrum <strong>of</strong> the 2D Cooperon with<br />

Zeeman term <strong>of</strong> the strength gµ B B/D e Q 2 SO<br />

= 0...2 <strong>in</strong> steps <strong>of</strong> 0.25 (w.l.o.g K y = 0). The<br />

B <strong>in</strong>dependent mode E T0 is not shown.


62 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

R[E]/DeQ 2 SO<br />

I[E]/DeQ 2 SO<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

2 1 0 1 2<br />

K x<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.2<br />

0.4<br />

0.6<br />

(a)<br />

2 1 0 1 2<br />

K x<br />

(b)<br />

˜B<br />

0 0.5 0.8<br />

Figure 3.17: (a) Real and(b) imag<strong>in</strong>ary parts <strong>of</strong> thespectrum<strong>of</strong> the Cooperonwith Zeeman<br />

term <strong>of</strong> the strength gµ B B/D e Q 2 SO<br />

= 0...0.8 <strong>in</strong> steps <strong>of</strong> 0.1 <strong>in</strong> a f<strong>in</strong>ite wire <strong>of</strong> the width<br />

Q SO W = 0.5. The B <strong>in</strong>dependent mode E t0 is not shown.


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 63<br />

the correction to the static conductivity <strong>in</strong> the case <strong>of</strong> a magnetic field which we <strong>in</strong>clude by<br />

means <strong>of</strong> a Zeeman term together with an effective magnetic field appear<strong>in</strong>g <strong>in</strong> the cut<strong>of</strong>f<br />

1/τ B as described <strong>in</strong> Sec.3.2. The ˜g = g/8m e D e factor is used as a material-dependent<br />

parameter. In Fig. 3.18, we see that for large enough ˜g factor, the system changes from<br />

positive magnetoconductivity—<strong>in</strong> the case without Zeeman field and a small-enough wire<br />

width—to negative magnetoconductivity at a f<strong>in</strong>ite Zeeman field for the same wire. Hence,<br />

the ratio W c /W WL changes and one has to be careful not to confuse the crossover def<strong>in</strong>ed<br />

by a change <strong>of</strong> the sign <strong>of</strong> the quantum correction, WL→WAL, and the crossover <strong>in</strong> the<br />

magnetoconductivity. To give an idea how the crossover W c depends on ˜g and the strength<br />

<strong>of</strong> the Zeeman field we analyze two different systems as plotted <strong>in</strong> Fig.3.19: The first<br />

one, plot (a), shows the drop <strong>of</strong> W c <strong>in</strong> a system as just described where we have one<br />

magnetic field which we <strong>in</strong>clude with an orbital and a Zeeman part. For small ˜g we have<br />

Q SO W c (˜g) = Q SO W c (˜g = 0)−const˜g 2 , where const is about 1 <strong>in</strong> the considered parameter<br />

space. Inthe second system [Fig.3.19(b)], we assumethat we can change theorbital and the<br />

Zeeman field separately. The critical width is plotted aga<strong>in</strong>st the Zeeman field. To calculate<br />

W c , we fix the Zeeman field to a certa<strong>in</strong> value, horizontal axis <strong>in</strong> plot (b), while we vary<br />

the effective field and calculate if negative or positive magnetoconductivity is present. For<br />

differentZeemanfieldsB Z /H s wegetdifferentW c . WeseethatW c isshiftedtolargerwidths<br />

as the Zeeman field is <strong>in</strong>creased, Q SO W c (B Z /H s ) = Q SO W c (B Z = 0) + const(B Z /H s ) 2 ,<br />

whereconst is about1<strong>in</strong> the considered parameter space, while∆σ(1/τ B = 0) (not plotted)<br />

is lowered as long as we assume small Zeeman fields. If we notice that B Z mixes s<strong>in</strong>glet and<br />

triplet states it is understood that there is no gapless s<strong>in</strong>glet mode anymore and therefore<br />

∆σ(1/τ B = 0) must decrease for low Zeeman fields.<br />

To estimate ˜g, we take typical values for a GaAs/AlGaAs system and assume the electron<br />

densitytoben s = 1.11×10 11 cm −2 , theeffective massm e /m e0 = 0.063, theLandéfactorg =<br />

0.75 and an elastic mean-free path <strong>of</strong> l e = 10 nm <strong>in</strong> a wire with Q SO W = 1, correspond<strong>in</strong>g<br />

to W = 1.2µm, if we assume a Rashba sp<strong>in</strong>-orbit coupl<strong>in</strong>g strength <strong>of</strong> α 2 = 5 meVÅ. We<br />

thus get ˜g ≈ 0.1 and f<strong>in</strong>d that the Zeeman coupl<strong>in</strong>g due to the perpendicular magnetic field<br />

can have a measurable, albeit small effect on the magnetoconductance <strong>in</strong> GaAs/AlGaAs<br />

systems.


64 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

1.4<br />

1.6<br />

∆σ/( 2e2<br />

2π )<br />

1.8<br />

2.0<br />

2.2<br />

0.0 0.5 1.0 1.5 2.0<br />

B/H S<br />

Figure 3.18: The magnetoconductance ∆σ(B) <strong>in</strong> a magnetic field perpendicular to the<br />

quantumwell, whereitscoupl<strong>in</strong>gviatheZeemantermisconsideredbyexact diagonalization<br />

while its effect on the orbital motion is considered effectively by the magnetic phase-shift<strong>in</strong>g<br />

rate 1/τ B (W) for wire width Q SO W = 1, dephas<strong>in</strong>g rate 1/τ ϕ = 0.06D e Q 2 SO<br />

, and elasticscatter<strong>in</strong>g<br />

rate 1/τ = 4D e Q 2 SO<br />

. The strength <strong>of</strong> the contribution <strong>of</strong> the Zeeman term is<br />

varied by the material-dependent factor ˜g = gµ B H s /D e Q 2 SO<br />

<strong>in</strong> the range ˜g = 0...1.5 <strong>in</strong><br />

steps <strong>of</strong> 0.5: The system changes from positive magnetoconductivity (˜g = 0, green) to<br />

negative one (˜g = 0.5...1.5, cont<strong>in</strong>uous decrease <strong>of</strong> the absolute m<strong>in</strong>imum).<br />

3.6 Conclusions<br />

In conclusion, <strong>in</strong> wires with sp<strong>in</strong>-conserv<strong>in</strong>g boundaries and a width W smaller<br />

than bulk sp<strong>in</strong> precession length L SO , the sp<strong>in</strong> relaxation due to l<strong>in</strong>ear Rashba SO coupl<strong>in</strong>g<br />

is suppressed accord<strong>in</strong>g to the sp<strong>in</strong> relaxation rate (1/τ s )(W) = (π 2 /3)(W/L SO ) 2 (1/τ s ),<br />

where 1/τ s = 2p 2 F α2 2 τ. The enhancement <strong>of</strong> sp<strong>in</strong> relaxation length L s = √ D e τ s (W)<br />

can be understood as follows: The area an electron covers by diffusion <strong>in</strong> time τ s is<br />

WL s . This sp<strong>in</strong> relaxation occurs if that area is equal L 2 SO,[Fal03] which yields 1/L 2 s ∼<br />

1/D e τ s ∼ (W/L SO ) 2 /L 2 SO, <strong>in</strong> agreement with Eq.(3.81). For larger wire widths, the exact<br />

diagonalization reveals a nonmonotonic behavior <strong>of</strong> the sp<strong>in</strong> relaxation as function<br />

<strong>of</strong> the wire width <strong>of</strong> the long-liv<strong>in</strong>g eigenstates. The sp<strong>in</strong> relaxation rate is first enhanced<br />

before it is suppressed as the widths W is decreased. The longest liv<strong>in</strong>g modes<br />

are found to exist at the boundary <strong>of</strong> wide wires. S<strong>in</strong>ce we identified a direct transformation<br />

from the sp<strong>in</strong>-diffusion equation to the Cooperon equation, we could show that<br />

these edge modes affect the conductivity: the 0-mode approximation overestimates the<br />

conductivity for larger wire widths Q SO W > 1 s<strong>in</strong>ce it does not take <strong>in</strong>to account these<br />

edge modes. They add a larger contribution to the negative triplet term <strong>of</strong> the quantum


Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems 65<br />

1.2<br />

Q SO Wc<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

g/8m e D e<br />

(a)<br />

3.5<br />

3.0<br />

Q SO Wc<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

B Z /H s<br />

(b)<br />

Figure 3.19: (a) Change <strong>of</strong> crossover width W c with g factor: The magnetic field is <strong>in</strong>cluded<br />

as an effective field 1/τ B and <strong>in</strong> the Zeeman term. The strength <strong>of</strong> the contribution <strong>of</strong> the<br />

Zeeman term is varied by the material dependent factor ˜g = gµ B H s /D e Q 2 SO. (b) Change<br />

<strong>of</strong> crossover width W c with Zeeman field: To calculate W c , we fix the Zeeman field to a<br />

certa<strong>in</strong> value, horizontal axis, while we vary the effective field <strong>in</strong>dependently and calculate<br />

if negative or positive magnetoconductivity is present. For different Zeeman fields B Z /H s ,<br />

we f<strong>in</strong>d thereby a different width W c . Here, we set g/8m e D e = 1. In (a) and (b), the cut<strong>of</strong>f<br />

due to dephas<strong>in</strong>g is varied: 1/D e Q 2 SOτ ϕ = 0.04,0.06,0.08 (lowest first).


66 Chapter 3: WL/WAL Crossover and <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Conf<strong>in</strong>ed Systems<br />

correction than the bulk modes do, s<strong>in</strong>ce they relax more slowly. This also results <strong>in</strong> a<br />

shift <strong>in</strong> the m<strong>in</strong>imum <strong>of</strong> the conductivity towards smaller magnetic fields <strong>in</strong> comparison to<br />

the 0-mode approximation. The reduction <strong>of</strong> sp<strong>in</strong> relaxation has recently been observed<br />

<strong>in</strong> optical measurements <strong>of</strong> n-doped InGaAs quantum wires[HSM + 06] and <strong>in</strong> transport<br />

measurements.[DLS + 05, LSK + 07, SGP + 06, WGZ + 06, Mei05] Recently <strong>in</strong> Ref.[KKN09],<br />

the enhancement <strong>of</strong> sp<strong>in</strong> lifetime due to dimensional conf<strong>in</strong>ement <strong>in</strong> gated InGaAs wires<br />

with gate controlled SO coupl<strong>in</strong>g was reported. Reference [HSM + 06] reports saturation <strong>of</strong><br />

sp<strong>in</strong>relaxation <strong>in</strong>narrowwires, W ≪ L SO , attributed tocubicDresselhaus coupl<strong>in</strong>g.[Ket07]<br />

The contribution <strong>of</strong> the l<strong>in</strong>ear and cubic Dresselhaus SO <strong>in</strong>teraction to the sp<strong>in</strong> relaxation<br />

turns out to depend strongly on growth direction and will be studied <strong>in</strong> more detail <strong>in</strong><br />

Chapter4. Includ<strong>in</strong>g both the l<strong>in</strong>ear Rashba and Dresselhaus SO coupl<strong>in</strong>g we have shown<br />

that there exist two long persist<strong>in</strong>g sp<strong>in</strong> helix solutions <strong>in</strong> narrow wires even for arbitrary<br />

strength <strong>of</strong> both SO coupl<strong>in</strong>g effects. This is <strong>in</strong> contrast to the 2D case, where the condition<br />

α 1 = α 2 , respectively <strong>in</strong> the case where the cubic Dresselhaus term cannot be neglected,<br />

α 1 −m e γ g E F /2 = α 2 , is required to f<strong>in</strong>d persistent sp<strong>in</strong> helices [BOZ06, LCCC06] as it was<br />

measured recently (Ref. [KWO + 09]). Regard<strong>in</strong>g the type <strong>of</strong> boundary, we found that the<br />

<strong>in</strong>jection <strong>of</strong> polarized sp<strong>in</strong>s <strong>in</strong>to nonmagnetic material is favorable for wires with a smooth<br />

conf<strong>in</strong>ement, λ F ∂ y V < ∆ α = 2k F α 2 . With such adiabatic boundary conditions, states<br />

which are polarized <strong>in</strong> z direction relax with a f<strong>in</strong>ite rate for wires with widths Q SO W ≪ 1,<br />

while the sp<strong>in</strong> relaxation rate <strong>of</strong> all other states diverges <strong>in</strong> that limit.<br />

In tubular wires with periodic boundary conditions, the sp<strong>in</strong> relaxation is found<br />

to rema<strong>in</strong> constant as the wire circumference is reduced. F<strong>in</strong>ally, by <strong>in</strong>clud<strong>in</strong>g the Zeeman<br />

coupl<strong>in</strong>g to the perpendicular magnetic field, we have shown that for sp<strong>in</strong>-conserv<strong>in</strong>g<br />

boundary condition the critical wire width, W c , where the crossover from negative to positive<br />

magnetoconductivity occurs, depends not only on the dephas<strong>in</strong>g rate but also depends<br />

on the g factor <strong>of</strong> the material.


Chapter 4<br />

Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong><br />

Relaxation and Diffusive-Ballistic<br />

Crossover<br />

4.1 Introduction<br />

In Chapter3 it was shown how the sp<strong>in</strong> relaxes <strong>in</strong> a quasi 1D electron system <strong>in</strong><br />

a quantum well grown <strong>in</strong> the [001] direction, depend<strong>in</strong>g on the width <strong>of</strong> the wire, where<br />

the normal <strong>of</strong> the boundary was po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the [010] direction. In a system with only<br />

Rashba SOC or l<strong>in</strong>ear Dresselhaus SOC <strong>in</strong> a (111) quantum well it is clear from the vector<br />

fields plotted <strong>in</strong> Fig.2.2 that a rotation <strong>in</strong> the pla<strong>in</strong> should not effect the physics,<br />

i.e. the m<strong>in</strong>imal sp<strong>in</strong> relaxation rate will not reduce. The l<strong>in</strong>ear Rashba SOC does<br />

not depend on the growth direction at all. It is already known that on the other hand<br />

<strong>in</strong> a (001) 2D system with both BIA[Dre55] and SIA[BR84] we get an anisotropic sp<strong>in</strong>relaxation.[KRW03,<br />

CWd07, WJW10] This has also been studied numerically <strong>in</strong> quasi-1D<br />

GaAs wires[LLK + 10]. In this Chapter, Sec.4.2, we present analytical results concern<strong>in</strong>g<br />

this anisotropy for the 2D case as well as the case <strong>of</strong> quantum wire with sp<strong>in</strong> and charge<br />

conserv<strong>in</strong>g boundaries. As <strong>in</strong> the previous chapter, also here we focus on materials where<br />

the dom<strong>in</strong>ant mechanism for sp<strong>in</strong> relaxation is governed by the DP sp<strong>in</strong> relaxation.<br />

Furthermore we extend our analysis to other growth directions, Sec.4.3: Search<strong>in</strong>g for long<br />

sp<strong>in</strong> decoherence times at room temperature, the (110) quantum wire attracted attention.<br />

67


68 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

[AOMO01, DHR + 04] The properties <strong>of</strong> sp<strong>in</strong> relaxation <strong>in</strong> systems with this growth direction<br />

have also been related to WL measurements, Ref.[HPB + 97]. We present analytical<br />

explanations for dimensional sp<strong>in</strong> relaxation reduction and discuss the crossover from WL<br />

to WAL, Sec.4.4.<br />

It was shown <strong>in</strong> the previous chapter that the cubic Dresselhaus SOC leads to a term,<br />

Eq.(3.45), which h<strong>in</strong>ders the sp<strong>in</strong> relaxation to vanish for small wire widths Q SO W ≪ 1<br />

but W ≫ l e , with the wire width W and the elastic mean free path l e . As we will show<br />

<strong>in</strong> the follow<strong>in</strong>g sections, this term is width dependent but the f<strong>in</strong>ite sp<strong>in</strong> relaxation rate<br />

is not reduced if the wire is rotated <strong>in</strong> the (001) plane. However some <strong>of</strong> the experiments<br />

are done on ballistic wires, i.e. <strong>in</strong> the regime where W ≫ l e does not hold, and we need<br />

to modify the theory used <strong>in</strong> Ref.[Ket07] and presented <strong>in</strong> the previous chapter to enable<br />

us to study the crossover from diffusive to ballistic wires. In Sec.4.5 we show how the sp<strong>in</strong><br />

relaxation which is due to cubic Dresselhaus SOC reduces with the number <strong>of</strong> channels <strong>in</strong><br />

the quantum wire.<br />

We consider aga<strong>in</strong> the Hamiltonian with SOC, Eq.(3.32)<br />

H = 1<br />

2m e<br />

(p+eA) 2 +V(x)− 1 2 γ gσ(B+B SO (p)), (4.1)<br />

where m e is the effective electron mass. A is the vector potential due to the external<br />

magnetic field B. B T SO<br />

= (B SOx ,B SOy ) is the momentum dependent SO field. σ is a<br />

vector, with components σ i , i = x,y,z, the Pauli matrices, γ g is the gyromagnetic ratio<br />

with γ g = gµ B with the effective g factor <strong>of</strong> the material, and µ B = e/2m e is the Bohr<br />

magneton constant. To analyze the sp<strong>in</strong> relaxation for different wire directions we use for<br />

the SO <strong>in</strong>teraction which is caused by BIA to lowest order <strong>in</strong> the wave vector k the general<br />

form, Eq.(2.13),<br />

− 1 2 γ ∑<br />

gB SO,D = γ D ê i p i (p 2 i+1 −p 2 i+2) (4.2)<br />

where the pr<strong>in</strong>cipal crystal axes are given by i ∈ {x,y,z},i → ((i − 1) mod 3) + 1 and<br />

the sp<strong>in</strong>-orbit coefficient for the bulk semiconductor γ D . We consider the standard whitenoise<br />

model for the impurity potential as <strong>in</strong> the previous chapter, V(x), which vanishes on<br />

average 〈V(x)〉 = 0, is uncorrelated, 〈V(x)V(x ′ )〉 = δ(x−x ′ )/2πντ, and weak, E F τ ≫ 1.<br />

To address both, the WL corrections as well as the sp<strong>in</strong> relaxation rates <strong>in</strong> the system, our<br />

start<strong>in</strong>g po<strong>in</strong>t is also here the Cooperon[HLN80]<br />

Ĉ(Q) −1 = 1 ( ∫ dϕ<br />

1−<br />

τ 2π<br />

i<br />

1<br />

1+iτ(v(Q+2eA+2m e âS)+H σ ′ +H Z )<br />

)<br />

, (4.3)


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 69<br />

as presented <strong>in</strong> Eq.(3.41). Expand<strong>in</strong>g the Cooperon to second order <strong>in</strong> (Q+2eA+2m e âS)<br />

and perform<strong>in</strong>g the angular <strong>in</strong>tegral which is for 2D diffusion (elastic mean free path l e<br />

smaller than wire width W) cont<strong>in</strong>uous from 0 to 2π, yields:<br />

Ĉ(Q) =<br />

1<br />

D e (Q+2eA+2eA S ) 2 +H γD<br />

. (4.4)<br />

The effective vector potential due to SO <strong>in</strong>teraction is A S = m eˆαS/e, where ˆα = 〈â〉 is<br />

averaged over angle. The SO term H γD , which cannot be rewritten as a vector potential, is<br />

<strong>in</strong> our case due to the appearance <strong>of</strong> cubic Dresselhaus SOC.<br />

4.1.1 Example<br />

Toget anidea<strong>of</strong>theprocedurewerecall thesituation presented<strong>in</strong>Ref.[Ket07] and<br />

Chapter3 and take up the remark about the additional term due to cubic Dresselhaus SOC,<br />

Eq.(3.45) concern<strong>in</strong>g the width dependence: Start<strong>in</strong>g with the Cooperon Hamiltonian, <strong>in</strong><br />

the case <strong>of</strong> Rashba and l<strong>in</strong>. and cubic Dresselhaus SOC,<br />

with the effective vector potential<br />

with ˜α 1 = α 1 −m e γ D E F /2,<br />

H c := Ĉ−1<br />

= (Q+2eA S ) 2 +(m 2 e<br />

D E Fγ D ) 2 (Sx 2 +S2 y ), (4.5)<br />

e<br />

A S = m e<br />

e ˆαS = m e<br />

e<br />

⎛<br />

⎝ −˜α 1 −α 2 0<br />

α 2 ˜α 1 0<br />

⎛ ⎞<br />

⎞ S x<br />

⎠⎜<br />

⎝ S y<br />

⎟<br />

⎠ , (4.6)<br />

S z<br />

it can be easily shown that the Hamiltonian Eq.(4.5) has only non vanish<strong>in</strong>g eigenvalues<br />

due to the last term <strong>in</strong> Eq.(4.5), which is due to cubic Dresselhaus SOC, if we assume no<br />

boundaries except the lateral conf<strong>in</strong>ement. This term is neither reduced by reason <strong>of</strong> the<br />

boundary<strong>in</strong> the diffusive case, as has been shown <strong>in</strong> the previous chapter for quantum wires<br />

<strong>in</strong> [100] direction and will be extended now to other directions, too. However two triplet<br />

eigenvalues <strong>of</strong> this term depend on the wire width,<br />

E QD1 = q2 s3<br />

2 , (4.7)<br />

(<br />

E QD2,3 = q2 s3 3<br />

2 2 ± s<strong>in</strong>(Q )<br />

SOW)<br />

, (4.8)<br />

2Q SO W<br />

with q 2 s3 /2 = (m2 e E Fγ D ) 2 . In the follow<strong>in</strong>g we are go<strong>in</strong>g to diagonalize the whole Hamiltonian<br />

and change the direction <strong>of</strong> the wire <strong>in</strong> the (001) plane.


70 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

4.2 <strong>Sp<strong>in</strong></strong> Relaxation anisotropy <strong>in</strong> the (001) system<br />

4.2.1 2D system<br />

We rotate the system <strong>in</strong>-plane through the angle θ (the angle θ = π/4 is equivalent<br />

to [110]). This does not effect the Rashba term but changes the Dresselhaus one to[CWd07,<br />

WJW10]<br />

1<br />

γ D<br />

H D[001] = σ y k y cos(2θ)(〈k 2 z 〉−k2 x )−σ xk x cos(2θ)(〈k 2 z 〉−k2 y )<br />

−σ y k x<br />

1<br />

2 s<strong>in</strong>(2θ)(k2 x −k2 y −2〈k2 z 〉)<br />

1<br />

+σ x k y<br />

2 s<strong>in</strong>(2θ)(k2 x −k2 y +2〈k2 z 〉), (4.9)<br />

with the wave vectors k i . The result<strong>in</strong>g Cooperon Hamiltonian, <strong>in</strong>clud<strong>in</strong>g Rashba and<br />

Dresselhaus SOC, reads then<br />

H c = (Q x +α x1 S x +(α x2 −q 2 )S y ) 2 +(Q y +(α x2 +q 2 )S x −α x1 S y ) 2 + q2 s3<br />

2 (S2 x +S 2 y),<br />

where we set<br />

(4.10)<br />

q 2 s3<br />

2 = ( m 2 eE F γ D<br />

) 2, (4.11)<br />

α x1 = 1 2 m eγ D cos(2θ)((m e v) 2 −4〈kz 2 〉), (4.12)<br />

α x2 = − 1 2 m eγ D s<strong>in</strong>(2θ)((m e v) 2 −4〈kz〉) 2 (4.13)<br />

( )<br />

=<br />

q 1 −√<br />

q<br />

2<br />

s3<br />

2<br />

s<strong>in</strong>(2θ) (4.14)<br />

= 2m e˜α 1 s<strong>in</strong>(2θ), (4.15)<br />

with q 1 = 2m e α 1 , q 2 = 2m e α 2 . We see that the part <strong>of</strong> the Hamiltonian which cannot be<br />

written as a vector field and is due to cubic Dresselhaus SOC does not depend on the wire<br />

direction <strong>in</strong> the (001) plane.<br />

Special case: Only l<strong>in</strong>. Dresselhaus SOC equal to Rashba SOC<br />

As a special example for the 2D case we set q s3 = 0 and q 1 = q 2 . To simplify<br />

the search for vanish<strong>in</strong>g sp<strong>in</strong> relaxation we go to polar coord<strong>in</strong>ates. Apply<strong>in</strong>g free wave


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 71<br />

functions (with k x ,k y ) to H c , Eq.(4.10), we end up with (s<strong>in</strong>glet part left out)<br />

H c<br />

q 2 2<br />

=<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

2+Q 2 f θφ −2i exp(2iθ)<br />

4+Q 2 f ⎟ θφ ⎠ (4.16)<br />

c.c. 2+Q 2<br />

with k x /q 2 = Qcos(φ), k y /q 2 = Qs<strong>in</strong>(φ) and<br />

f θφ = (i −1) √ 2exp(iθ)(cos(φ+θ)−s<strong>in</strong>(φ+θ))Q. (4.17)<br />

Vanish<strong>in</strong>g sp<strong>in</strong>relaxation is foundat Q = 0 for arbitrary values <strong>of</strong> θ (the sp<strong>in</strong>with vanish<strong>in</strong>g<br />

sp<strong>in</strong> relaxation is po<strong>in</strong>t<strong>in</strong>g along the [110] direction[SEL03]). Another solution is found at<br />

Q = 2 with the condition θ + φ = 3π/4, which is equivalent to the [110] crystallographic<br />

direction.[CWd07]<br />

4.2.2 Quasi-1D wire<br />

In the follow<strong>in</strong>g we consider sp<strong>in</strong> and charge conserv<strong>in</strong>g boundaries. Due to the<br />

SOC we have, as <strong>in</strong> the previous chapter, Eq.(3.70), modified Neumann condition with the<br />

slight change which is the angle dependency <strong>of</strong> the vector potential:<br />

(− τ D e<br />

n·〈v F [γ g B SO (k)·S]〉−i∂ n<br />

)<br />

C| ∂S = 0, (4.18)<br />

where 〈...〉 denotes the average over the direction <strong>of</strong> v F and k which we rewrite for the<br />

rotated x-y system<br />

(−i∂ y +2e(A S ) y<br />

)C<br />

(<br />

x,y = ± W )<br />

= 0, ∀x, (4.19)<br />

2<br />

where n is the unit vector normal to the boundary ∂S and x is the coord<strong>in</strong>ate along the<br />

wire. In order to do a diagonalization tak<strong>in</strong>g only the zero-mode <strong>in</strong>to account, we have<br />

to simplify the boundary condition. A transformation act<strong>in</strong>g <strong>in</strong> the transverse direction is<br />

needed accord<strong>in</strong>g to Eq.(4.10): Ĉ → ˜Ĉ = U A ĈU † A<br />

, by us<strong>in</strong>g the transformation<br />

U =½4 −i s<strong>in</strong>(q s y) 1 q s<br />

A y +(cos(q s y)−1) 1 q 2 sA 2 y (4.20)<br />

with A y = (α x2 +q 2 )S x −α x1 S y and q s = √ (α x2 +q 2 ) 2 +α 2 x1 .


72 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

<strong>Sp<strong>in</strong></strong> relaxation<br />

We diagonalize the Hamiltonian, Eq.(4.10), after apply<strong>in</strong>g the transformation U,<br />

tak<strong>in</strong>g only the lowest mode <strong>in</strong>to account similar to the calculation presented <strong>in</strong> Sec.3.4.2.<br />

The spectrum <strong>of</strong> the Hamiltonian for small wire width, Wq s < 1, is given by<br />

(<br />

α<br />

E 1/2 (k x > 0) = kx 2 2<br />

±k x<br />

(2q sm − x1 +α 2 2<br />

)<br />

x2 2) −q2 W 2<br />

12q sm<br />

+ 3 (<br />

qs3<br />

2<br />

2 2 +q2 sm ∓ q2 s3 α<br />

2<br />

x1 +α 2 ) 2W<br />

x2 −q2 2<br />

2<br />

2k x 96q<br />

( sm<br />

q 2<br />

s3<br />

(α<br />

2 sm) +q2 2<br />

x1 +α 2 ) 2<br />

x2 −q2 2<br />

−<br />

W 2 , (4.21)<br />

24q 2 sm<br />

E 1 (k x = 0) = q 2 s3 +q2 sm − (α2 x1 +α2 x2 −q2 2 )2 + q2 s3<br />

E 2 (k x = 0) = q2 s3<br />

2 +q2 sm + q2 s3<br />

2<br />

2 q2 s<br />

W 2 , (4.22)<br />

12<br />

q2 2α2 x1<br />

3qsm<br />

2 W 2 , (4.23)<br />

( q 2<br />

s3<br />

(α<br />

E 3 = kx 2 + q2 s3<br />

2 + 2 sm) +q2 2<br />

x1 +α 2 x2 −q2 2<br />

12q 2 sm<br />

) 2<br />

W 2 , (4.24)<br />

with q sm = √ (α x2 −q 2 ) 2 +α 2 x1 . First we notice that the only θ dependence is <strong>in</strong> the<br />

term q sm , which disappears if the Dresselhaus SOC strength ˜α 1 , which is shifted due to<br />

the cubic term, equals the Rashba SOC strength α 2 and the angle <strong>of</strong> the boundary is<br />

θ = (1/4 + n)π, n ∈. Assum<strong>in</strong>g the term proportional to W 2 /k x to be small, the<br />

absolute m<strong>in</strong>imum can be found at<br />

( ) (α<br />

E 1/2,m<strong>in</strong> = 3 qs3<br />

2 q 2<br />

2 2 + sm − q2 s3 2<br />

2 x1 +α 2 ) 2<br />

x2 −q2 2<br />

W 2 (4.25)<br />

24q 2 sm<br />

which is <strong>in</strong>dependent <strong>of</strong> the width W if α x1 (θ = 0) = −q 2 and/or the direction <strong>of</strong> the wire<br />

is po<strong>in</strong>t<strong>in</strong>g <strong>in</strong><br />

θ = 1 ( 2〈k<br />

2<br />

2 arcs<strong>in</strong> z 〉(m e γ D ) 2 ((m e v) 2 −2〈kz〉)−q 2 2<br />

2 )<br />

(m 3 ev 2 γ D −4〈kz〉m 2 . (4.26)<br />

e γ D )q 2<br />

The second possible absolute m<strong>in</strong>imum, which dom<strong>in</strong>ates for sufficient small width W and<br />

q sm ≠ 0 (compare with E 2 (k x = 0)), is found at<br />

( q 2 (α<br />

s3<br />

E 3,m<strong>in</strong> = q2 s3<br />

2 + 2 sm) +q2 2<br />

x1 +α 2 ) 2<br />

x2 −q2 2<br />

12qsm<br />

2 W 2 . (4.27)


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 73<br />

The m<strong>in</strong>imal sp<strong>in</strong>-relaxation rate is found by analyz<strong>in</strong>g the prefactor <strong>of</strong> W 2 <strong>in</strong> Eq.(4.27),<br />

Fig.(4.1). We see immediately that <strong>in</strong> the case <strong>of</strong> vanish<strong>in</strong>g cubic Dresselhaus or <strong>in</strong> the case<br />

where α x1 (θ = 0) = −q 2 we have no direction dependence <strong>of</strong> the m<strong>in</strong>imal sp<strong>in</strong> relaxation.<br />

Notice the shift <strong>of</strong> the absolute m<strong>in</strong>imum away from q 1 = q 2 due to q s3 ≠ 0. In the case <strong>of</strong><br />

q 1 < (q s3 / √ 2)wef<strong>in</strong>dthem<strong>in</strong>imumatθ = (1/4+n)π, n ∈, elseatθ = (3/4+n)π, n ∈,<br />

which is <strong>in</strong>dicated by the dashed l<strong>in</strong>e <strong>in</strong> Fig.(4.1).<br />

Π<br />

0 <br />

4<br />

Π<br />

<br />

2<br />

3Π<br />

<br />

4 Π<br />

2.0<br />

1.5<br />

1<br />

q1/q2<br />

1.0<br />

0.5<br />

0<br />

0.0<br />

Π<br />

0 <br />

4<br />

Π<br />

<br />

3Π<br />

2 <br />

θ<br />

4<br />

Π<br />

Figure 4.1: Dependence <strong>of</strong> the W 2 coefficient <strong>in</strong> Eq.(4.27) on the lateral rotation (θ). The<br />

absolute m<strong>in</strong>imum is found for α x1 (θ = 0) = −q 2 (here: q 1 /q 2 = 1.63) and for different SO<br />

strength we f<strong>in</strong>d the m<strong>in</strong>imum at θ = (1/4 + n)π, n ∈if q 1 < (q s3 / √ 2) (dashed l<strong>in</strong>e:<br />

q 1 = (q s3 / √ 2)) and at θ = (3/4 +n)π, n ∈else. Here we set q s3 = 0.9. The scal<strong>in</strong>g is<br />

arbitrary.<br />

<strong>Sp<strong>in</strong></strong> dephas<strong>in</strong>g<br />

Concern<strong>in</strong>g sp<strong>in</strong>tronic devices it is <strong>in</strong>terest<strong>in</strong>g to know how an ensemble <strong>of</strong> sp<strong>in</strong>s<br />

<strong>in</strong>itially oriented along the [001] direction dephases <strong>in</strong> the wire <strong>of</strong> different orientation θ.<br />

Todothisanalysisweonlyhavetoknowthattheeigenvector fortheeigenvalueE 1 atk x = 0,<br />

Eq.(4.22), is the triplet state |S = 1;m = 0〉 = (|↑↓〉 + |↓↑〉)/ √ 2 ≡<br />

|⇉〉 ˆ=(0,1,0) T , Eqs.(3.62). This is equal to the z-component <strong>of</strong> the sp<strong>in</strong> density whose<br />

evolution is described by the sp<strong>in</strong> diffusion equation, Eq.(3.54). As an example we assume<br />

the case where cubic Dresselhaus term can be neglected and where the Rashba and<br />

l<strong>in</strong>. Dresselhaus SOC are equal. We notice that the dephas<strong>in</strong>g is then width <strong>in</strong>dependent.


74 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

At def<strong>in</strong>ite angles the dephas<strong>in</strong>g time diverges - as for the <strong>in</strong>-plane polarized states with<br />

eigenvalue E 2 (k x = 0) - ,<br />

1<br />

τ s (W) = 2D eq 2 2(1−s<strong>in</strong>(2θ)) (4.28)<br />

whichisplotted<strong>in</strong>Fig.(4.2). Wehavelongest sp<strong>in</strong>dephas<strong>in</strong>gtimeatθ = (1/4+n)π, n ∈.<br />

For θ = (3/4+n)π, n ∈we get the 2D result T 2 = 1/(4q 2 2 D e). We notice that this is the<br />

eigenvalue to the triplet state |S = 1;m = 0〉 <strong>of</strong> the sp<strong>in</strong> relaxation tensor [DP71b, DP71c],<br />

which we derived <strong>in</strong> Sec.3.3.<br />

1<br />

= τγ 2 (<br />

g 〈B SO (k) 2 〉δ ij −〈B SO (k) i B SO (k) j 〉 ) (4.29)<br />

(ˆτ s ) ij<br />

This gives an analytical description <strong>of</strong> numerical calculation done by J.Liu et al., Ref.-<br />

[LLK + 10].<br />

Switch<strong>in</strong>g on cubic Dresselhaus SOC leads to f<strong>in</strong>ite sp<strong>in</strong> dephas<strong>in</strong>g time for all angles θ. In<br />

addition T 2 is then width dependent. In the case <strong>of</strong> strong cubic Dresselhaus SOC where<br />

q 2 s3 /2 = q2 1 = q2 2 , the dephas<strong>in</strong>g time T 2 is angle <strong>in</strong>dependent and for q 2 s3 /2 > q2 1 = q2 2 the<br />

m<strong>in</strong>ima <strong>in</strong> T 2 (θ) change to maxima and vice versa.<br />

50.0<br />

0.<br />

0.1<br />

T2Deq 2 2<br />

10.0<br />

5.0<br />

1.0<br />

0.2<br />

0.3<br />

0.4<br />

0.5<br />

∈<br />

0 Π Π 3Π<br />

<br />

Π 5Π 3Π 7Π<br />

<br />

2Π<br />

4 2 4<br />

4 2 4<br />

θ<br />

Figure4.2: Thesp<strong>in</strong>dephas<strong>in</strong>gtimeT 2 <strong>of</strong>asp<strong>in</strong><strong>in</strong>itially orientedalongthe[001] direction<strong>in</strong><br />

units <strong>of</strong> (D e q2 2 ) for the special case <strong>of</strong> equal Rashba and l<strong>in</strong>. Dresselhaus SOC. The different<br />

curves show different strength <strong>of</strong> cubic Dresselhaus <strong>in</strong> units <strong>of</strong> q s3 /q 2 . In the case <strong>of</strong> f<strong>in</strong>ite<br />

cubic Dresselhaus SOC we set W = 0.4/q 2 . If q s3 = 0: T 2 diverges at θ = (1/4+n)π, n<br />

(dashed vertical l<strong>in</strong>es). The horizontal dashed l<strong>in</strong>e <strong>in</strong>dicated the 2D sp<strong>in</strong> dephas<strong>in</strong>g time,<br />

T 2 = 1/(4q 2 2 D e).


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 75<br />

Special case: θ = 0<br />

In this case the longitud<strong>in</strong>al direction <strong>of</strong> the wire is [100].<br />

If we neglect the term proportional to W 2 /k x <strong>in</strong> Eq.(4.21) the lowest sp<strong>in</strong> relaxation is<br />

found to be<br />

or<br />

depend<strong>in</strong>g whether<br />

(<br />

1 α<br />

= q2 2<br />

s3 x1 −q 2 ) ( 2<br />

D e τ s 2 + 2<br />

12q 2 s<br />

q 2 s + q2 s3<br />

2<br />

(<br />

1 α<br />

= 3q2 2<br />

s3 x1 −q 2 ) ( ) 2<br />

D e τ s 4 + 2 qs 2 − q2 s3<br />

2<br />

24qs<br />

2<br />

(<br />

α<br />

− q2 2<br />

s3 x1 −q 2 ) ( 2<br />

4 + 2<br />

q 2 s +3 q2 s3<br />

2<br />

)<br />

W 2<br />

)<br />

W 2<br />

W 2<br />

(4.30)<br />

, (4.31)<br />

24qs<br />

2 (4.32)<br />

is negative or positive. This shows that the cubic Dresselhaus term adds not only to<br />

the relaxation rate by a constant term but is also width dependent. However, this width<br />

dependence does not reduce the sp<strong>in</strong> relaxation rate below q 2 s3 /2.<br />

4.3 <strong>Sp<strong>in</strong></strong> relaxation <strong>in</strong> quasi-1D wire with [110] growth direction<br />

To get the sp<strong>in</strong>-relaxation <strong>in</strong> a [110] quantum wire with Rashba and Dresselhaus<br />

SOC aga<strong>in</strong> we have to rotate the spacial coord<strong>in</strong>ate system <strong>of</strong> the Dresselhaus Hamiltonian<br />

Eq.(4.2) but now with the rotation matrix<br />

We get<br />

R =<br />

⎛<br />

⎜<br />

⎝<br />

√1<br />

1<br />

2<br />

0 √2<br />

− 1 √<br />

2<br />

0<br />

1 √2<br />

0 1 0<br />

⎞<br />

H D[110]<br />

γ D<br />

= σ x (−k 2 x k z −2k 2 y k z +k 3 z )<br />

+σ y (4k x k y k z )<br />

⎟<br />

⎠ . (4.33)<br />

+σ z (kx 3 −2k xky 2 −k xkz 2 ). (4.34)<br />

The conf<strong>in</strong>ement <strong>in</strong> z-direction (z ≡[110]) leads to 〈k z 〉 = 〈k 3 z 〉 = 0, and 〈k2 z 〉 = ∫ |∇φ| 2 dz.<br />

The Hamiltonian for the quantum wire <strong>in</strong> [110] direction has then the follow<strong>in</strong>g form


76 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

[HPB + 97]<br />

( 1<br />

H [110] = −γ D σ z k x<br />

2 〈k2 z〉− 1 )<br />

2 (k2 x −2ky)<br />

2 . (4.35)<br />

Includ<strong>in</strong>g the Rashba SOC (q 2 ), not<strong>in</strong>g that its Hamiltonian does not depend on the orientation<br />

<strong>of</strong> the wire,[HPB + 97] we end up with the follow<strong>in</strong>g Cooperon Hamiltonian<br />

C −1<br />

D e<br />

= (Q x − ˜q 1 S z −q 2 S y ) 2 +(Q y +q 2 S x ) 2 + ˜q2 3<br />

2 S2 z , (4.36)<br />

with<br />

˜q 1 = 2m e<br />

γ D<br />

2 〈k2 z〉− γ D<br />

2<br />

m e E F<br />

2<br />

, (4.37)<br />

q 2 = 2m e α 2 , (4.38)<br />

and ˜q 3 = (3m e E 2 F(γ D /2)). (4.39)<br />

We see immediately that <strong>in</strong> the 2D case states polarized <strong>in</strong> the z-direction have vanish<strong>in</strong>g<br />

sp<strong>in</strong> relaxation as long as we have no Rashba SOC. Compared with the (001) system the<br />

constant term due to cubic Dresselhaus does not mix sp<strong>in</strong> directions. Here we set the<br />

appropriate Neumann boundary condition as follows:<br />

(<br />

(−i∂ y +2m e α 2 S x )C x,y = ± W )<br />

= 0, ∀x. (4.40)<br />

2<br />

The presence <strong>of</strong> Rashba SOC adds a vector potential proportional to S x . Apply<strong>in</strong>g a nonabelian<br />

gauge transformation as before to simplify the boundary condition, we diagonalize<br />

the transformed Hamiltonian (App.(D.1)) up to second order <strong>in</strong> q 2 W <strong>in</strong> the 0-mode approximation.<br />

4.3.1 Special case: without cubic Dresselhaus SOC<br />

The spectrum is found to be<br />

E 1 = k 2 x + 1<br />

12 ∆2 (q 2 W) 2 , (4.41)<br />

E 2,3 = k 2 x + 1<br />

24 ∆2( 24−(q 2 W) 2)<br />

± ∆ 24√<br />

∆ 2 (q 2 W) 4 +4k 2 x (24−(q 2W) 2 ) 2 , (4.42)<br />

with the lowest sp<strong>in</strong> relaxation rate found at f<strong>in</strong>ite wave vectors k x m<strong>in</strong><br />

= ± ∆ 24 (24−(q 2W) 2 ),<br />

We set ∆ = √˜q 2 1 +q2 2 .<br />

1<br />

D e τ s<br />

= ∆2<br />

24 (q 2W) 2 . (4.43)


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 77<br />

4.3.2 With cubic Dresselhaus SOC<br />

If cubic Dresselhaus SOC cannot be neglected, the absolute m<strong>in</strong>imum <strong>of</strong> sp<strong>in</strong><br />

relaxation can also shift to k x m<strong>in</strong><br />

= 0. This depends on the ratio <strong>of</strong> Rashba and l<strong>in</strong>.<br />

Dresselhaus SOC:<br />

If q 2 /q 1 ≪ 1, we f<strong>in</strong>d the absolute m<strong>in</strong>imum at k x m<strong>in</strong><br />

= 0,<br />

E m<strong>in</strong>1 = ˜q 3 + ˜q 2 1 +q2 2<br />

2<br />

−∆ c + 1 12 ∆ c(q 2 W) 2 , (4.44)<br />

with<br />

∆ c = 1 2<br />

√<br />

(˜q 3 + ˜q 2 1 )2 +2(˜q 2 1 − ˜q 3)q 2 2 +q4 2 . (4.45)<br />

If q 2 /q 1 ≫ 1, we f<strong>in</strong>d the absolute m<strong>in</strong>imum at k x m<strong>in</strong><br />

≈ ± ∆ 24 (24−(q 2W) 2 ),<br />

) (˜q<br />

E m<strong>in</strong>2 = kx 2 2<br />

m<strong>in</strong><br />

−k x m<strong>in</strong><br />

q 1 ˜q 3<br />

2 2<br />

q2<br />

2 +2 −<br />

16k x m<strong>in</strong><br />

q 2<br />

+∆ 2 + ˜q )<br />

3<br />

(˜q<br />

2<br />

1<br />

+1<br />

2 q2<br />

2 (˜q3˜q 1<br />

2 −<br />

12 − ˜q 3 2q 2<br />

+ q4 2<br />

24 − (˜q<br />

2<br />

1<br />

24 + q2 2<br />

12<br />

3072kx 3 − q2 2<br />

m<strong>in</strong><br />

24 (˜q 3 − ˜q 1)<br />

2<br />

)<br />

q 2 k x m<strong>in</strong><br />

− q ((<br />

2 ˜q<br />

2<br />

3<br />

k x m<strong>in</strong><br />

128 + ˜q 3˜q 1<br />

2<br />

192<br />

)<br />

− ˜q 3q2<br />

2 ))<br />

W 2 . (4.46)<br />

96<br />

We can conclude that reduc<strong>in</strong>g wire width W will not cancel the contribution due to cubic<br />

Dresselhaus SOC to the sp<strong>in</strong> relaxation rate.<br />

4.4 Weak Localization<br />

In Ref.[Ket07] and the previous chapter the crossover from WL to WAL due<br />

to change <strong>of</strong> wire width and SOC strength was expla<strong>in</strong>ed <strong>in</strong> the case <strong>of</strong> a (001) system.<br />

Whether WL or WAL is present depends on the suppression <strong>of</strong> the triplet modes <strong>of</strong> the<br />

Cooperon. The suppression <strong>in</strong> turn is dom<strong>in</strong>ated by the absolute m<strong>in</strong>imum <strong>of</strong> the spectrum<br />

<strong>of</strong> the Cooperon Hamiltonian H c . The f<strong>in</strong>d<strong>in</strong>gs presented <strong>in</strong> Sec.4.2.2 therefore po<strong>in</strong>t out<br />

that e.g. the crossover width, at which the system changes from WL to WAL, can shift<br />

with the wire direction θ. Recently experimental results on WL/WAL by J. Nitta et al.,<br />

Ref.[Nit06], seem to show a strong dependence on growth direction. Our presented results


78 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

can also support the method proposed <strong>in</strong> Ref.[SKK + 08] -to determ<strong>in</strong>e the relative strength<br />

<strong>of</strong> Rashba and Dresselhaus SOC from WL/WAL measurements without fitt<strong>in</strong>g parameterswith<br />

<strong>in</strong>clusion <strong>of</strong> cubic Dresselhaus term for wire directions different from [100] and [010]<br />

<strong>in</strong> an analytical manner.<br />

In the (110) system the situation is different: In the 2D case it was shown by Pikus et al.,<br />

Ref.[HPB + 97], that <strong>in</strong> the absence <strong>of</strong> the Rashba terms the negative magnetoconductivity<br />

cannot be observed. In the case <strong>of</strong> a wire geometry we can conclude from Eqs.(4.41-4.46)<br />

that we have no width dependence if Rashba SOC vanishes. A change <strong>of</strong> the quantum<br />

correction to the static conductivity therefore cannot be achieved <strong>in</strong> this wire geometry by<br />

chang<strong>in</strong>g the wire width. The reason is the vector potential <strong>in</strong> the boundary condition,<br />

Eq.(4.40), which only depends on the Rashba SOC.<br />

4.5 Diffusive-Ballistic Crossover<br />

In the follow<strong>in</strong>g we assume a (001) 2D system with both, Rashba and l<strong>in</strong>ear and<br />

cubic Dresselhaus SO coupl<strong>in</strong>g.<br />

Experiments measur<strong>in</strong>g WL <strong>in</strong> diffusive quantum wires with SOC[LSK + 07, SGB + 09] are <strong>in</strong><br />

great agreement with theoretical calculations by S. Kettemann, Ref.[Ket07]. But consider<strong>in</strong>g<br />

e.g. the works Ref.[KKN09, KHG + 10], one realizes that the scope <strong>of</strong> application <strong>of</strong> the<br />

theory has to be extended to describe also the crossover to the ballistic regime, l e > W. In<br />

Fig.4.3 we show an example <strong>of</strong> an experiment done by Kunihashi et al.[KKN09]. The l<strong>in</strong>es<br />

show a fit us<strong>in</strong>g Eq.(10) <strong>in</strong> the paper by S. Kettemann[Ket07] which is equal to the results<br />

presented <strong>in</strong> Sec.4.2.2 if the width dependence <strong>of</strong> the term due to cubic Dresselhaus SOC<br />

is neglected. To get the solid l<strong>in</strong>es the term due to cubic Dresselhaus SOC was neglected,<br />

the dashed <strong>in</strong>clude it. When check<strong>in</strong>g Fig.4.3 (a) the condition l e < W is not fulfilled at all<br />

wire width. It follows that this additional term is suppressed if the condition <strong>of</strong> a diffusive<br />

system does not hold anymore.<br />

We have shown <strong>in</strong> Sec.4.2.2 that the presence <strong>of</strong> cubic Dresselhaus SOC <strong>in</strong> the<br />

sample leads to a f<strong>in</strong>ite sp<strong>in</strong> relaxation even for wire widths Q SO W ≪ 1, regardless <strong>of</strong> the<br />

boundary direction <strong>in</strong> a (001) system. To account for the ballistic case we have to modify<br />

the derivation <strong>of</strong> the Cooperon Hamiltonian, Eq.(4.5). In the case <strong>of</strong> a wire where the mean


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 79<br />

(a)<br />

(b)<br />

Figure 4.3: Example <strong>of</strong> an experiment done by Kunihashi et al.[KKN09]. The width dependence<br />

<strong>of</strong> the sp<strong>in</strong> relaxation length l 1D SO<br />

<strong>of</strong> different carrier density. Solid l<strong>in</strong>es and dashed<br />

l<strong>in</strong>es <strong>in</strong> (b) show the l 1D SO<br />

calculated from the theory presented <strong>in</strong> this work, with neglect<strong>in</strong>g<br />

cubic Dresselhaus term and tak<strong>in</strong>g <strong>in</strong>to account full SOIs, respectively.


80 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

free path l e is comparable to the wire width W, we cannot perform the step from<br />

to<br />

∑<br />

q<br />

E,p + q<br />

E ′ ,p ′ −q<br />

∫ 2π<br />

0<br />

= 1<br />

2πντ<br />

∑<br />

GE,σ R (p+q)G E A ′ ,σ ′(p′ −q) (4.47)<br />

q<br />

dϕ<br />

2π (1+iτve ϕ(Q+2m e a(ϕ).S)) −1 (4.48)<br />

and <strong>in</strong>tegrate <strong>in</strong> Eq.(4.3) over the Fermi surface <strong>in</strong> a cont<strong>in</strong>uous way. Instead, we assume<br />

k F /W to be f<strong>in</strong>ite and sum over the number <strong>of</strong> discrete channels N = [k F W/π], where [...]<br />

is the <strong>in</strong>teger part. Because H γD ∼ E 2 F<br />

this constant term due to cubic Dresselhaus should<br />

reduce if we reduce the number <strong>of</strong> channels. If we expand the Cooperon to second order <strong>in</strong><br />

(Q+2eA+2m e âS) before averag<strong>in</strong>g over the Fermi surface, 〈...〉, and use the Matsubara<br />

trick, we get<br />

C −1 ( ( ) )<br />

= 2f 1 Q y +2α 2 S x +2 α 1 −γ D v 2f 2<br />

3<br />

S y<br />

D e f 1<br />

( )<br />

+2f 2<br />

(Q x −2α 2 S y −2 α 1 −γ D v 2f 2<br />

3<br />

S x<br />

)<br />

f 2<br />

[( ) ( )<br />

+8γDv 2 4 f 4 − f2 3<br />

Sx 2 + f 5 − f2 3<br />

Sy<br />

2 f 2 f 1<br />

]<br />

, (4.49)<br />

with m e = 1 and functions f i (ϕ) (App.E) which depend on the number <strong>of</strong> transverse modes<br />

N. In the diffusive case we can perform the cont<strong>in</strong>uous sum over the angle ϕ <strong>in</strong> Eq.(E.3)-<br />

(E.7), and we receive the old result with f 1 = f 2 = 1/2, f 3 = 1/8 and f 4 = f 5 = 1/16:<br />

H c = (Q y +2α 2 S x +2<br />

(α 1 − 1 )<br />

2 γ DE F S y ) 2 +(Q x −2α 2 S y −2<br />

(α 1 − 1 )<br />

2 γ DE F S x ) 2<br />

+(γ D E F ) 2 (S 2 x +S 2 y). (4.50)<br />

4.5.1 <strong>Sp<strong>in</strong></strong> Relaxation at Q SO W ≪ 1<br />

In the first section we analyzed the lowest sp<strong>in</strong> relaxation <strong>in</strong> wires <strong>of</strong> different<br />

direction <strong>in</strong>a(001) system. We have shown, that forevery direction thereis still af<strong>in</strong>itesp<strong>in</strong><br />

relaxation at wire width which fulfill the condition Q SO W ≪ 1 due to cubic Dresselhaus<br />

SOC. It is clear that this f<strong>in</strong>ite sp<strong>in</strong> relaxation vanishes when the width is equal to the<br />

Fermi wave length λ F . In the follow<strong>in</strong>g we show how this f<strong>in</strong>ite sp<strong>in</strong> relaxation depends<br />

on the number <strong>of</strong> transverse channels N. We show <strong>in</strong> Ref.[WK] that the f<strong>in</strong>d<strong>in</strong>gs are<br />

consistent with calculations go<strong>in</strong>g beyond the perturbative ansatz. This is possible <strong>in</strong> a


Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover 81<br />

similar manner as has been done previously <strong>in</strong> Ref.[KM02] for wires without SOC, which<br />

showed the crossover <strong>of</strong> the magnetic phase shift<strong>in</strong>g rate, which had been known before <strong>in</strong><br />

the diffusive and ballistic limit, only.<br />

T<strong>of</strong><strong>in</strong>dthespectrum<strong>of</strong>theCooperonHamiltonianwithboundaryconditionsas<strong>in</strong>Sec.4.2.2,<br />

we stay <strong>in</strong> the 0-mode approximation <strong>in</strong> the Q space and proceed as before: Accord<strong>in</strong>g to<br />

Eq.(4.49), the non-Abelian gauge transformation for the transversal direction y is given by<br />

( [ ( ) )<br />

U = exp −i 2α 2 S x +2 α 1 −γ D v 2f 3<br />

S y<br />

]y . (4.51)<br />

f 1<br />

To concentrate on the constant width <strong>in</strong>dependent part <strong>of</strong> the spectrum we extract the<br />

absolute m<strong>in</strong>imum at Q = 0, Fig.(4.4) and Fig.(4.5). A clear reduction <strong>of</strong> the absolute<br />

m<strong>in</strong>imum is visible. Due to the factor f 3 /f 1 <strong>in</strong> the transformation U, the decrease <strong>of</strong> the<br />

m<strong>in</strong>imal sp<strong>in</strong> relaxation depends also on the ratio <strong>of</strong> Rashba and l<strong>in</strong>ear Dresselhaus SOC.<br />

E m<strong>in</strong> /ǫ 2 F γ2 D<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.<br />

0.5<br />

1.<br />

1.5<br />

2.<br />

2.5<br />

3.<br />

0.2<br />

0.0<br />

5 10 15 20 25 30<br />

N<br />

Figure 4.4: The lowest eigenvalues <strong>of</strong> the conf<strong>in</strong>ed Cooperon Hamiltonian Eq.(4.49), equivalent<br />

to the lowest sp<strong>in</strong> relaxation rate, are shown for Q = 0 for different number <strong>of</strong> modes<br />

N = k F W/π. Different curves correspond to different values <strong>of</strong> α 2 /q s .<br />

From Eq.(4.49) it is clear, that not only the H γD is affected by the reduction<br />

<strong>of</strong> the number <strong>of</strong> channels N but also the shift <strong>of</strong> the l<strong>in</strong>. Dresselhaus SOC, α 1 , <strong>in</strong> the<br />

orbital part. A model to extract the ratio <strong>of</strong> Rashba and l<strong>in</strong>. Dresselhaus SOC developed<br />

<strong>in</strong> Ref.[SANR09] by Scheid et al. did not show much difference between the strict 1D case<br />

and the non-diffusive case with wire <strong>of</strong> f<strong>in</strong>ite width. The results presented here should allow<br />

for extend<strong>in</strong>g the model to f<strong>in</strong>ite cubic Dresselhaus SOC. Deduc<strong>in</strong>g from our theory, the<br />

direction <strong>of</strong> the SO field should change with the number <strong>of</strong> channels due to the mentioned<br />

N dependent shift.


82 Chapter 4: Direction Dependence <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation and Diffusive-Ballistic Crossover<br />

E m<strong>in</strong> /ǫ 2 F γ2 D<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.<br />

0.5<br />

1.<br />

1.5<br />

2.<br />

2.5<br />

3.<br />

0.2<br />

0.0<br />

5 10 15 20 25 30<br />

N<br />

Figure 4.5: The lowest eigenvalues <strong>of</strong> the conf<strong>in</strong>ed Cooperon Hamiltonian Eq.(4.49), equivalent<br />

to the lowest sp<strong>in</strong> relaxation rate, are shown for Q = 0 for different number <strong>of</strong> modes<br />

N = k F W/π. Different curves correspond to different values <strong>of</strong> α 1 /q s .<br />

4.6 Conclusions<br />

Summariz<strong>in</strong>g the results <strong>of</strong> this chapter, we have characterized the anisotropy and<br />

width dependence <strong>of</strong> sp<strong>in</strong> relaxation <strong>in</strong> a (001) quantum wire. There are special angles<br />

θ which are optimal for sp<strong>in</strong> transport <strong>in</strong> quantum wires <strong>of</strong> f<strong>in</strong>ite width: The [110] and<br />

the [110] direction. At [110] we f<strong>in</strong>d the longest sp<strong>in</strong> dephas<strong>in</strong>g time T 2 . If the absolute<br />

m<strong>in</strong>imum <strong>of</strong> sp<strong>in</strong> relaxation is found at [110] or [110] direction depends on the strength <strong>of</strong><br />

cubicDresselhausandwirewidth. Thef<strong>in</strong>d<strong>in</strong>gsforthesp<strong>in</strong>dephas<strong>in</strong>gtimeare<strong>in</strong>agreement<br />

with numerical results. The analytical expression for T 2 allows to see directly the <strong>in</strong>terplay<br />

between the cubic Dresselhaus SOC and the dimensional reduction, hav<strong>in</strong>g effect on T 2 .<br />

In addition we analyzed the special case <strong>of</strong> a (110) system and found the m<strong>in</strong>imal sp<strong>in</strong><br />

relaxation rates depend<strong>in</strong>g on Rashba and l<strong>in</strong>. and cubic Dresselhaus SOC <strong>in</strong> the presence<br />

<strong>of</strong> boundaries. This results can be used to understand width and direction dependent WL<br />

measurements <strong>in</strong> quantum wires. F<strong>in</strong>ally, we have shown how the reduction <strong>of</strong> channels<br />

<strong>in</strong> the wire reduces the f<strong>in</strong>ite sp<strong>in</strong> relaxation rate which is due to cubic Dresselhaus SOC<br />

and does not reduce if the wire is small, Wq s ≪ 1, and diffusive, W ≫ l e . The change <strong>in</strong><br />

channel number also changes the shift <strong>of</strong> l<strong>in</strong>. Dresselhaus SOC strength, ˜α 1 . This has to<br />

be considered if extract<strong>in</strong>g SOC strength from wires with only few transverse channels.


Chapter 5<br />

<strong>Sp<strong>in</strong></strong> Hall Effect<br />

5.1 Introduction<br />

In order to realize sp<strong>in</strong>tronic devices like the sp<strong>in</strong> field effect transistor, one needs<br />

to <strong>in</strong>duce sp<strong>in</strong> polarized electrons <strong>in</strong> low dimensional electron systems (LDES)[icvacFDS04,<br />

DD90]. <strong>Sp<strong>in</strong></strong> polarized electrons can be generated by <strong>in</strong>ject<strong>in</strong>g a current with ferromagnetic<br />

metallic leads <strong>in</strong>to the LDES. However, it has been found that <strong>in</strong> practice the efficiency<br />

<strong>of</strong> such sp<strong>in</strong> <strong>in</strong>jection is poor because <strong>of</strong> the conductivity mismatch. Therefore, recently<br />

there has been a strong effort to f<strong>in</strong>d new ways for generat<strong>in</strong>g polarized sp<strong>in</strong> currents. One<br />

possibility is to use a T-shaped conductor with SOC as proposed by Yamamoto[YDKO06],<br />

whose efficiency is restricted however strongly by impurity scatter<strong>in</strong>g and is therefore only<br />

applicable to narrow wires with few channels. Another approach is to dope the semiconductor<br />

with magnetic impurities: When doped with several percent one speaks <strong>of</strong> dilute<br />

magnetic semiconductors. With transition metal atoms, like Mn, these materials become<br />

ferromagnetic, such as In 1−x Mn x As which has been discovered by Ohno to have a ferromagnetic<br />

phase with a relatively high critical temperature[OMP + 92]. Due to the still relatively<br />

small concentration <strong>of</strong> impurities this system can still be manipulated <strong>in</strong> a wide range <strong>of</strong><br />

carrier density, impurity concentration and acceptor level energy: As these impurities not<br />

only provide sp<strong>in</strong>, but also dope the system with holes, the density <strong>of</strong> charge carriers, and<br />

the Fermi energy can be changed, by chang<strong>in</strong>g their concentration. By choos<strong>in</strong>g different<br />

elements, also the acceptor level energy can be changed. This together with the disorder<br />

provided by these magnetic impurities does change the magnetic properties substantially,<br />

s<strong>in</strong>ce the effective magnetic coupl<strong>in</strong>g between the magnetic ions does depend itself on the<br />

83


84 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

DOS, their average distance and the disorder strength [SBK + 10].<br />

Another way for efficiently <strong>in</strong>ject<strong>in</strong>g sp<strong>in</strong> currents <strong>in</strong>to semiconductors is to make<br />

use <strong>of</strong> the sp<strong>in</strong> Hall effect (SHE). This effect was first proposed by D’yakonov and Perel<br />

[DP71a] and describes <strong>in</strong> today’s term<strong>in</strong>ology the extr<strong>in</strong>sic SHE which requires sp<strong>in</strong> dependent<br />

impurity scatter<strong>in</strong>g. It was experimentally confirmed by the angle-resolved optical<br />

detection <strong>of</strong> sp<strong>in</strong> polarization at the edges <strong>of</strong> a two-dimensional layer [WKSJ05, KMGA04].<br />

The theory <strong>of</strong> the SHE has been developed <strong>in</strong> the last 10 years, as reviewed <strong>in</strong> Refs.<br />

[Sch06, ERH07].<br />

The sp<strong>in</strong> transport not only occurs due to the sp<strong>in</strong> precession <strong>in</strong> the bulk, but is affected<br />

by the scatter<strong>in</strong>g from nonmagnetic impurities, which can depend on the direction <strong>of</strong> the<br />

sp<strong>in</strong> itself due to the so called skew scatter<strong>in</strong>g and the side jump mechanism[Sch06]. In the<br />

first part <strong>of</strong> the follow<strong>in</strong>g chapter, we will outl<strong>in</strong>e the formalism to calculate analytically<br />

the SHE which arises even <strong>in</strong> the absence <strong>of</strong> impurities, the so called <strong>in</strong>tr<strong>in</strong>sic SHE, which<br />

is due to the bulk SOC. For the clean case we will <strong>in</strong>clude both, Rashba and Dresselhaus<br />

SOC. In the second part we focus on numerical methods: The first attempt is application <strong>of</strong><br />

exact diagonalization which has <strong>of</strong> course strong limitations concern<strong>in</strong>g the system size. To<br />

overcome this limitations, we apply the Kernel Polynomial Method (KPM) <strong>in</strong> the last part<br />

<strong>of</strong> this thesis. It is first applied to treat the metal-<strong>in</strong>sulator transition (MIT) <strong>in</strong> a symplectic<br />

system, f<strong>in</strong>ally we calculate the SHE us<strong>in</strong>g the KPM.<br />

5.1.1 About the Def<strong>in</strong>ition <strong>of</strong> <strong>Sp<strong>in</strong></strong> Current<br />

We have learned <strong>in</strong> Sec.2.3.4 that <strong>in</strong> presence <strong>of</strong> the sp<strong>in</strong>-orbit <strong>in</strong>teraction, the<br />

sp<strong>in</strong> current components are not conserved[CSS + 04, SZXN06] even if we assume no sp<strong>in</strong><br />

relaxation: In the cont<strong>in</strong>uity equation<br />

∂s z<br />

∂t +D e∇J sp<strong>in</strong> = T s − 1<br />

(ˆτ s ) ij<br />

s j (5.1)<br />

= τ〈∇v F (B SO (k)×S) i<br />

〉− 1<br />

(ˆτ s ) ij<br />

s j . (5.2)<br />

anadditionaltorquetermT s appears<strong>in</strong>contrasttoEq.2.12besidesthetermwhichdescribes<br />

the sp<strong>in</strong> relaxation rate 1/τ s . It follows that Noether’s theorem is not applicable to def<strong>in</strong>e<br />

the sp<strong>in</strong> current. This is a reason why the comparison between results done with Kubo<br />

formalism and calculations us<strong>in</strong>g Landauer-Büttiker approach is not trivial, where the sp<strong>in</strong>


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 85<br />

current <strong>in</strong> the latter is def<strong>in</strong>ed by[MM05]<br />

where<br />

I s p,µ = 1<br />

4π<br />

∑<br />

p≠q,ν<br />

Tr[Γ µ pG R Γ ν qG A ](V p −V q ), (5.3)<br />

Γ µ p = i(Σ µ p −(Σ µ p) † ), (5.4)<br />

with the retarded self energy Σ µ p due to coupl<strong>in</strong>g <strong>of</strong> lead p, which has voltage V p , and<br />

sample for sp<strong>in</strong> channel µ and advanced and retarded Green’s function G A/R . However,<br />

<strong>in</strong> an experiment the SHE can be measured. This measurement has to be connected to<br />

the theoretical description us<strong>in</strong>g l<strong>in</strong>ear response to a transverse electrical field with the<br />

frequency ω which yields, Appendix B, Eq.(B.16),<br />

σ µν (ω) = i V<br />

∑<br />

m,n<br />

(f(E m )−f(E n ))〈m|j ν |n〉〈n|j µ |m〉<br />

E n −E m E n −E m +ω +iη . (5.5)<br />

To calculate the sp<strong>in</strong> Hall conductivity (SHC) the correlation function consists, <strong>in</strong> contrast<br />

to charge conductivity, <strong>of</strong> the charge current and a current which conta<strong>in</strong>s a sp<strong>in</strong> operator.<br />

In the simplified model we use the sp<strong>in</strong> current which is given by the anticommutator <strong>of</strong><br />

the sp<strong>in</strong> and the group velocity v = i[H,r],<br />

J z = 4 {σ z,v}. (5.6)<br />

This sp<strong>in</strong> current does not differ from the one def<strong>in</strong>ed <strong>in</strong> Sec.2.3.[ESL05, BNnM04] Note<br />

that this quantity is dissipationless, because the sp<strong>in</strong> current is even under the time-reversal<br />

operation.<br />

5.2 SHE without Impurities: Exact Calculation<br />

We consider theHamiltonian for alattice which provides l<strong>in</strong>ear Rashba, Eq.(2.15),<br />

and l<strong>in</strong>ear and cubic Dresselhaus SOC, Eq.(2.14), as <strong>in</strong>troduced <strong>in</strong> Sec.2.3.3. The conf<strong>in</strong>ement<br />

to generate the 2D electron gas is <strong>in</strong> [001] direction. In order to do calculations<br />

numerically, one needs to def<strong>in</strong>e a tight b<strong>in</strong>d<strong>in</strong>g model on a discrete lattice <strong>of</strong> f<strong>in</strong>ite lattice<br />

spac<strong>in</strong>g a. It has the follow<strong>in</strong>g characterization:<br />

• square lattice,


86 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

• localized site orbitals are <strong>of</strong> s symmetry.<br />

Apply<strong>in</strong>g this assumptions the tight-b<strong>in</strong>d<strong>in</strong>g version <strong>of</strong> the Hamiltonian is given by<br />

H = H 0 +H R +H D,l<strong>in</strong> +H D,cubic , (5.7)<br />

= ∑ ǫ i c † i,σ c i,σ −t ∑<br />

c † i,σ c j,σ<br />

i,σ 〈i,j〉,σ<br />

∑<br />

{c † l,m,σ<br />

(iσ ′ y ) σσ ′c l+1,m,σ −c † l,m,σ<br />

(iσ ′ x ) σσ ′c l,m+1,σ }<br />

+ α 2<br />

2a<br />

+ α 1<br />

2a<br />

σ,σ ′<br />

l,m<br />

∑<br />

{c † l,m,σ<br />

(iσ ′ x ) σσ ′c l+1,m,σ −c † l,m,σ<br />

(iσ ′ y ) σσ ′c l,m+1,σ }<br />

σ,σ ′<br />

l,m<br />

⎧<br />

⎪⎨<br />

+ γ ∑<br />

D<br />

a 3 {c †<br />

⎪<br />

l,m,σ<br />

(−iσ<br />

⎩ ′ x ) σσ ′c l+1,m,σ +c † l,m,σ<br />

(iσ ′ y ) σσ ′c l,m+1,σ }<br />

σ,σ ′<br />

l,m<br />

σ,σ ′<br />

l,m<br />

⎫<br />

+ 1 ∑<br />

⎪⎬<br />

{c †<br />

2<br />

l,m,σ<br />

(i(σ ′ x −σ y )) σσ ′c l+1,m+1,σ +c † l,m,σ<br />

(i(σ ′ x +σ y )) σσ ′c l+1,m−1,σ }<br />

⎪⎭<br />

+h.c.. (5.8)<br />

where c † i,σ is the creation operator at site <strong>in</strong>dex i with sp<strong>in</strong> σ =↑,↓ and c† l,m,σ<br />

the creation<br />

operator at site (<strong>in</strong>dex x ,<strong>in</strong>dex y ) = (l,m). The hopp<strong>in</strong>g coupl<strong>in</strong>g t is given by t = 1/(2m e a)<br />

with the lattice constant a. In the follow<strong>in</strong>g we take the cubic Dresselhaus term only as<br />

a shift <strong>of</strong>, ˜α 1 = α 1 − 2γ D /a 2 , accord<strong>in</strong>g to Eq.(3.44), and assume a clean system, i.e the<br />

on-site energy is set to ǫ i = 0. Apply<strong>in</strong>g a Fourier transformation to Eq.(5.7) and go<strong>in</strong>g to<br />

momentum space we get (we set a ≡ 1)<br />

⎧<br />

H = ∑ ⎪⎨<br />

−2t(cos(k<br />

⎪ x )+cos(k y )) δ σσ ′c †<br />

kx,ky ⎩ } {{ }<br />

k,σ<br />

c ′ k,σ<br />

σ,σ ′ E 0<br />

+ (α 2 s<strong>in</strong>(k y )− ˜α 1 s<strong>in</strong>(k x ))c † k,σ ′ (σ x ) σσ ′c k,σ<br />

+ (˜α 1 s<strong>in</strong>(k y )−α 2 s<strong>in</strong>(k x ))c † k,σ ′ (σ y ) σσ ′c k,σ<br />

}<br />

. (5.9)<br />

The correspond<strong>in</strong>g eigenvalues are<br />

E ± (k) = E 0 (k)±∆(k) (5.10)


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 87<br />

with ∆(k) = √ (α 2 2 + ˜α2 1 )(s<strong>in</strong>(k x) 2 +s<strong>in</strong>(k y ) 2 )−4α 2˜α 1 s<strong>in</strong>(k x )s<strong>in</strong>(k y )<br />

(E − (k) is plotted <strong>in</strong> Fig.5.1). The eigenvectors are given by<br />

⎛<br />

|+/−〉 = √ 1 ⎝<br />

2<br />

∓({˜α 1 s<strong>in</strong>(k x)−α 2 s<strong>in</strong>(k y)}+i{˜α 1 s<strong>in</strong>(k y)−α 2 s<strong>in</strong>(k x)})<br />

∆<br />

1<br />

⎞<br />

⎠ (5.11)<br />

(a)<br />

(b)<br />

Figure 5.1: Energy band E − (k), Eq.(5.10) is plotted for pure Rashba SOC as function<br />

<strong>of</strong> wave vector k. The contour l<strong>in</strong>es <strong>in</strong>dicate the energy at which one f<strong>in</strong>ds a Van Hove<br />

s<strong>in</strong>gularity <strong>in</strong> the DOS (below half-fill<strong>in</strong>g).<br />

To calculate the SHE us<strong>in</strong>g Kubo formula, Eq.(5.5), we have to calculate the<br />

matrix elements <strong>of</strong> sp<strong>in</strong> current operator and velocity operator. In the site basis <strong>of</strong> our<br />

lattice they have the follow<strong>in</strong>g form:<br />

〈n|v|m〉 = 〈0| ∑ ij<br />

αβ<br />

= 〈0| ∑ ij<br />

αβ<br />

ψn(i,α)c ∗ iα vψ m (j,β)c † jβ<br />

|0〉, (5.12)<br />

ψn(i,α)c ∗ 1<br />

iα<br />

i [r,H]ψ m(j,β)c † jβ<br />

|0〉, (5.13)<br />

with r = ∑ kσ r kc † kσ c kσ<br />

= 1 i 〈0|∑ kγ<br />

∑<br />

ψn ∗ (i,α)c iα[r k c † kγ c kγ]Hc † jβ ψ m(j,β)<br />

ij<br />

αβ<br />

−ψ ∗ n(i,α)c iα<br />

1<br />

i H[c† kγ c kγr k ]c † jβ ψ m(j,β)|0〉, (5.14)


88 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

with c iα c † iα c iα|0〉 = c iα |0〉 and c iα c † jβ c jβ = c † jβ c jβc iα<br />

= 1 i<br />

v = 1 i<br />

∑<br />

ij<br />

αβ<br />

ψ ∗ n(i,α)[(r i −r j )H αβ<br />

ij ]ψ m(j,β), (5.15)<br />

∑<br />

(r i −r j )H αβ<br />

ij , (5.16)<br />

ij<br />

αβ<br />

where we used i,j and k as site <strong>in</strong>dices and the other for sp<strong>in</strong>. Us<strong>in</strong>g the def<strong>in</strong>ition <strong>of</strong> the<br />

sp<strong>in</strong> current, Eq.(5.6), we get accord<strong>in</strong>gly<br />

〈n|J z |m〉 = e ∑<br />

(r i −r j )ψ n (i,α){σ z ,H} αβ<br />

ij<br />

4i<br />

ψ n(j,β). (5.17)<br />

ij<br />

αβ<br />

Go<strong>in</strong>g <strong>in</strong>to momentum space we are left with<br />

v y = ∑ k y<br />

2ts<strong>in</strong>(k y )½+(α 2 σ x + ˜α 1 σ y )cos(k y ), (5.18)<br />

and for the sp<strong>in</strong> current<br />

J z x = ∑ k x<br />

ts<strong>in</strong>(k x )σ z . (5.19)<br />

The last Eq. follows from {σ z ,H D } = {σ z ,H R } = 0. To evaluate the rhs <strong>of</strong> Eq.(5.5) we<br />

need the matrix elements <strong>of</strong> both operators <strong>in</strong> the eigenvalue basis, which yields the pure<br />

imag<strong>in</strong>ary result<br />

〈∓|Jx|±〉〈±|v z y |∓〉 = ±it(α 2 2 − ˜α 2 1) s<strong>in</strong>(k x) 2 cos(k y )<br />

. (5.20)<br />

∆(k)<br />

Assum<strong>in</strong>gzerotemperaturetocalculateσ z xy ,Eq.(5.5)f<strong>in</strong>allysimplifiesto[NSSM05,MMF08]<br />

σ z xy ≡ σ SH = − ie<br />

V<br />

= 2 e V<br />

∑<br />

m,n<br />

∑<br />

f(E m )−f(E n )〈m|Jx z |n〉〈n|v y |m〉<br />

E m −E n (E m −E n )+iη , (5.21)<br />

I(〈m|Jx z |n〉〈n|v y |m〉)<br />

(E n −E m ) 2 +η 2 . (5.22)<br />

E m


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 89<br />

1.0<br />

1.0<br />

0.5<br />

0.5<br />

ΣsH e 8Π<br />

0.0<br />

σsH/(e/8π)<br />

0.0<br />

0.5<br />

0.5<br />

1.0<br />

4 2 0 2 4<br />

EFermi<br />

1.0<br />

4 2 0 2 4<br />

E/t<br />

(a)<br />

(b)<br />

Figure 5.2: (a) SHC σ SH as a function <strong>of</strong> Fermi energy E F , <strong>in</strong> a clean system <strong>of</strong> size<br />

L 2 = 170×170 with both Rashba and l<strong>in</strong>ear Dresselhaus SOC with α 2 > α 1 (blue curve)<br />

and α 2 < α 1 (red curve). (b) SHC σ SH as a function <strong>of</strong> Fermi energy E F , <strong>in</strong> a clean system<br />

<strong>of</strong> size V = 150×150 with only Rashba SOC <strong>of</strong> strength α 2 = 0.8t (blue/solid), α 2 = 1.4t<br />

(red/dotted) and α 2 = 2t (yellow/dashed).<br />

where we used Eq.(5.20). The level broaden<strong>in</strong>g η is <strong>in</strong>troduced for regularization <strong>in</strong> f<strong>in</strong>itesize<br />

systems [NSSM05, MMF08]. It is chosen to be <strong>of</strong> the order <strong>of</strong> the level spac<strong>in</strong>g δE,<br />

vanish<strong>in</strong>g <strong>in</strong> the thermodynamic limit. Integration over the Brillou<strong>in</strong> zone we f<strong>in</strong>ally arrive<br />

at the clean solution for the SHC which is plotted <strong>in</strong> Fig.(5.2).<br />

Results and Discussion<br />

TheSHC shows electron-hole symmetry: TheSHCvanishes at half-fill<strong>in</strong>g, E F = 0,<br />

and is an odd function <strong>of</strong> E F . From Eq.(5.18) one can see that the sign-change is due to<br />

the term which is proportional to the SOC strength. It is worth notic<strong>in</strong>g that an evaluation<br />

<strong>of</strong> the commutator [r,H] <strong>in</strong> Eq.(5.13) will lead to<br />

v = − i m e<br />

∂ R½2×2 +α 2 (σ x e y −σ y e x )+α 1 (σ y e y −σ x e x ) (5.24)<br />

(compare for pure Rashba case e.g. with [SCN + 04]). A straight forward tight b<strong>in</strong>d<strong>in</strong>g<br />

formulation <strong>of</strong> the velocity operator <strong>in</strong> this form would yield, <strong>in</strong> contrast to Eq.(5.16), onsite<br />

matrix elements proportional to the sp<strong>in</strong>-orbit <strong>in</strong>teraction which could yield unphysical<br />

results: The contribution <strong>of</strong> the velocity operator to the SHC would give 〈+|v y |−〉 =<br />

iα 2 s<strong>in</strong>(k x )/∆(k). Obviouslythemiss<strong>in</strong>gcos(k y )factor leads toaneven andthereforewrong<br />

SHC function <strong>of</strong> E F . Because the l<strong>in</strong>ear <strong>in</strong> momentum SO <strong>in</strong>teractions can be <strong>in</strong>terpreted


90 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

as an effective magnetic field, which however does not break time reversal symmetry, this<br />

is very similar to the case where a Peierls substitution is applied,[Pei33] i.e. where a Peierls<br />

phase is added to the electron whenever it hops <strong>in</strong> the direction <strong>of</strong> f<strong>in</strong>ite vector field.<br />

The sign changes accord<strong>in</strong>g to the sign <strong>of</strong> α 2 2 − ˜α2 1 and exhibits the value e/(8π) for small<br />

fill<strong>in</strong>g with the condition that both bands, E ± , are filled, <strong>in</strong>dependent <strong>of</strong> the strength <strong>of</strong><br />

SOC. This can bee seen by expand<strong>in</strong>g the spectrum around the Γ po<strong>in</strong>t which yields[She04]<br />

σ SH =<br />

∫<br />

e 2π<br />

16m e π 2<br />

= e<br />

8π<br />

with k x = kcos(ϕ) and k x = ks<strong>in</strong>(ϕ).<br />

0<br />

dϕ (α2 2 − ˜α2 1 )cos2 (ϕ)(k + −k − )<br />

(α 2 2 + ˜α2 1 −2α , (5.25)<br />

2˜α 1 s<strong>in</strong>(2ϕ)) 3 2<br />

α 2 2 − ˜α2 1<br />

|α 2 2 − (5.26)<br />

˜α2 1 |,<br />

Look<strong>in</strong>g at the results from the calculation on a clean lattice, Fig.(5.2) (b), it can be seen<br />

that the value e/(8π) decreases with <strong>in</strong>creas<strong>in</strong>g SOC strength <strong>in</strong> the case <strong>of</strong> pure Rashba<br />

SOC. This can be understood by notic<strong>in</strong>g that the value <strong>of</strong> the SHC[SCN + 04]<br />

σ SH =<br />

e<br />

16m e πα 2<br />

(k F+ −k F− ) (5.27)<br />

is dim<strong>in</strong>ished when we add corrections to the parabolic assumption: On the lattice we have<br />

( )<br />

2<br />

(k F+ −k F− ) = arccos<br />

1+ ( α 2<br />

) 2<br />

−1<br />

(5.28)<br />

2t<br />

= 2m e α 2 − 2 3 (m eα 2 ) 3 +O(m e α 2 ) 4 (5.29)<br />

and therefore the dim<strong>in</strong>ishment is given by<br />

σ SH = e<br />

8π − e<br />

24π (m eα 2 ) 2 . (5.30)<br />

5.3 Numerical Analysis <strong>of</strong> SHE<br />

5.3.1 Exact Diagonalization<br />

For l<strong>in</strong>ear Rashba coupl<strong>in</strong>g, the value σ SH = e 2 /(8π), as presented <strong>in</strong> the previous<br />

section, has been obta<strong>in</strong>ed both by analytical calculations <strong>in</strong> the cont<strong>in</strong>uum model, and<br />

by numerical calculations <strong>of</strong> the tight b<strong>in</strong>d<strong>in</strong>g model[Sch06]. However, <strong>in</strong> the presence<br />

<strong>of</strong> nonmagnetic impurities, the DC sp<strong>in</strong> Hall conductance is dim<strong>in</strong>ished to exactly zero,


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 91<br />

as soon as the system size L exceeds the elastic mean free path l e . In the case <strong>of</strong> δ-<br />

function potential and pure l<strong>in</strong>ear Rashba SOC <strong>in</strong> the thermodynamic limit this result is<br />

even <strong>in</strong>dependent<strong>of</strong> theimpurity strength, as discovered by Schwab andRaimondi[RS05] by<br />

us<strong>in</strong>gselfconsistentBornapproximation: Thevertex correction <strong>in</strong>theladderapproximation<br />

cancel exactly the term com<strong>in</strong>g from the one-loop part. Follow<strong>in</strong>g Rashba, Ref.[Ras04], this<br />

can be understood as follows: Correspond<strong>in</strong>g to Eq.(5.10) and Eq.(5.22) we have, at low<br />

fill<strong>in</strong>g, positive contribution to σ SH due to <strong>in</strong>terbranch transitions between the occupied<br />

(E + ) and unoccupied (E − ) states. Add<strong>in</strong>g a perturbation by apply<strong>in</strong>g an <strong>in</strong>f<strong>in</strong>itesimal<br />

small magnetic field, <strong>in</strong>trabranch transitions at the Fermi energy give rise to a negative<br />

contribution. Surpris<strong>in</strong>gly, this additional contribution cancels the first and we are left<br />

with zero SHC. This counter<strong>in</strong>tuitive f<strong>in</strong>d<strong>in</strong>g led to several controversies because the first<br />

numerical calculations <strong>in</strong> this field showed f<strong>in</strong>ite SHC when extrapolated to <strong>in</strong>f<strong>in</strong>ite large<br />

samples[NSJ + 05]. However, new numerical results, e.g. Ref.[NSJ + 06] as an erratum to<br />

Ref.[NSJ + 05], showed agreement with the analytical predictions. In contrast, tak<strong>in</strong>g <strong>in</strong>to<br />

account also sp<strong>in</strong>-orbit terms which are cubic <strong>in</strong> momentum, as they are present <strong>in</strong> any<br />

system with broken <strong>in</strong>version symmetry (cubic Dresselhaus terms), or <strong>in</strong> quantum wells<br />

with strongly asymmetric conf<strong>in</strong>ement (cubic Rashba terms), the sp<strong>in</strong> Hall conductance<br />

has a quite universal value, <strong>of</strong> σ SH = Ne 2 /(8π), where N is the number <strong>of</strong> times the sp<strong>in</strong>orbit<br />

field B SO (k) w<strong>in</strong>ds around a circle, as the momentum is moved once around the Fermi<br />

surface[Sch06].<br />

Resonant Impurities<br />

The effect <strong>of</strong> resonant impurities and <strong>of</strong> magnetic impurities on the SHC has<br />

not been studied yet <strong>in</strong> that detail[LX06, WLZ07a]. Especially, it is unclear how large<br />

its magnitude is, when the Fermi energy is <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the resonant levels close to<br />

the metal-<strong>in</strong>sulator transition, where it has been observed that the sp<strong>in</strong> relaxation rate is<br />

m<strong>in</strong>imal[DKK + 02], mak<strong>in</strong>g it a potentially attractive regime for sp<strong>in</strong>tronic applications.<br />

In the follow<strong>in</strong>g we analyze the reduction <strong>of</strong> the SHC <strong>in</strong> a f<strong>in</strong>ite system with periodic<br />

boundary conditions <strong>in</strong> presence <strong>of</strong> non-magnetic impurities <strong>of</strong> b<strong>in</strong>ary type, i.e. on-site<br />

potential V i = p i V where p i = 1 for impurity sites and p i = 0 otherwise. The lattice is<br />

assumed to be contam<strong>in</strong>ated with 10% <strong>of</strong> impurities. To have a better understand<strong>in</strong>g <strong>of</strong><br />

how the SHC changes with the strength <strong>of</strong> impurities, V, we first calculate the DOS. The


92 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

result is presented <strong>in</strong> Fig.5.3. For small V one would expect the Van Hove s<strong>in</strong>gularity at<br />

half fill<strong>in</strong>g. Due to f<strong>in</strong>ite Rashba SOC it is split to f<strong>in</strong>ite energies E = ±(2t− √ 4+α 2 2 t),<br />

<strong>in</strong>dicated <strong>in</strong> Fig.5.1 for the energy below half fill<strong>in</strong>g. If the impurity strength is <strong>in</strong>creased,<br />

a preformed impurity band is created. Us<strong>in</strong>g exact diagonalization 1 and apply<strong>in</strong>g the Kubo<br />

formalism, Eq.5.22, we calculate the SHC. Exemplarily we show the result for σ SH (E) at<br />

V = −2.8t <strong>in</strong> Fig.5.4. The SHC is strongly reduced but shows an additional maximum at<br />

energy where the preformed impurity band is located, as can be seen <strong>in</strong> comparison with<br />

the DOS, Fig.5.4(b).<br />

Toanalyze thereduction <strong>of</strong>SHCfor agiven fill<strong>in</strong>gn, wevary theimpuritystrength<br />

up to V = −5t and keep the concentration constant at 10%. Similar to the results <strong>in</strong> the<br />

case <strong>of</strong> block distribution <strong>of</strong> impurity strength, we see a monotone suppression at all fill<strong>in</strong>gs,<br />

even <strong>in</strong> the preformed impurity band, Fig.5.5.<br />

5.3.2 Kernel Polynomial Method<br />

The numerical calculations presented <strong>in</strong> the previous section, which were based<br />

on exact diagonalization us<strong>in</strong>g LAPACK[LAP] rout<strong>in</strong>es, are limited to small system sizes.<br />

This leads to f<strong>in</strong>ite size effects like oscillations <strong>in</strong> the SHC, e.g. Fig.5.4(b) above half fill<strong>in</strong>g.<br />

For further calculations concern<strong>in</strong>g the role <strong>of</strong> the impurity band it is necessary to do a<br />

f<strong>in</strong>ite size scal<strong>in</strong>g analysis and consider system-sizes beyond L = 64 which makes an exact<br />

treatment on current hardware impossible: for a D-dimensional matrix such a calculation<br />

requires memory <strong>of</strong> the order <strong>of</strong> D 2 , and the LAPACK rout<strong>in</strong>e scales as D 3 .<br />

Another problem is the adjustment <strong>of</strong> the cut<strong>of</strong>f η, see Eq.(5.22), which has to be taken<br />

with care as analyzed e.g. by Nomura et al., Ref.[NSSM05].<br />

Toovercome thelimitation onsmall systems, therearedifferent numerical order-Dmethods.<br />

Oneprocedureisthetimeevolution projection methoddevelopedbyTanakaandItoh[TI98],<br />

which was already used to calculate SHC[MMF08, MM07]. However, the algorithm requires<br />

both the choice <strong>of</strong> a sufficient number <strong>of</strong> time steps and an adjustment <strong>of</strong> cut<strong>of</strong>f η. A more<br />

effective method, which uses Chebyshev expansion based on Kernel Polynomial Method,<br />

will be presented <strong>in</strong> the follow<strong>in</strong>g.<br />

1 us<strong>in</strong>g LAPACK[LAP] and OpenMP[OMP] <strong>in</strong> C++


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 93<br />

V⩵0.2<br />

V⩵2.8<br />

0.20<br />

0.15<br />

Ρ<br />

0.15<br />

0.10<br />

0.05<br />

Ρ<br />

0.10<br />

0.05<br />

0.00<br />

10 5 0 5 10<br />

E<br />

0.00<br />

10 5 0 5 10<br />

E<br />

V⩵ 3<br />

V⩵ 3.6<br />

0.15<br />

0.15<br />

Ρ<br />

0.10<br />

Ρ<br />

0.10<br />

0.05<br />

0.05<br />

0.00<br />

10 5 0 5 10<br />

E<br />

0.00<br />

10 5 0 5 10<br />

E<br />

V⩵4<br />

V⩵5<br />

0.15<br />

0.15<br />

0.10<br />

0.10<br />

Ρ<br />

Ρ<br />

0.05<br />

0.05<br />

0.00<br />

10 5 0 5 10<br />

E<br />

0.00<br />

10 5 0 5 10<br />

E<br />

Figure 5.3: DOS, as a function <strong>of</strong> Fermi energy <strong>in</strong> units <strong>of</strong> t, <strong>in</strong> presence <strong>of</strong> impurities <strong>of</strong><br />

b<strong>in</strong>ary type with a concentration <strong>of</strong> 10% calculated us<strong>in</strong>g exact diagonalization. The system<br />

size is L 2 = 32 2 , and the SOC is Rashba type with α 2 = 1.2t with cut<strong>of</strong>f η = 0.02t.


94 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

0.10<br />

V⩵2.8<br />

Σ SH e 8Π<br />

0.05<br />

0.00<br />

0.05<br />

0.10<br />

0.15<br />

0.20<br />

Ρ<br />

0.15<br />

0.10<br />

0.05<br />

2⋆Σ SH e 8Π<br />

4 2 0 2 4<br />

(a)<br />

E<br />

0.00<br />

10 5 0 5 10<br />

(b)<br />

E<br />

Figure 5.4: (a) SHC, as a function <strong>of</strong> Fermi energy <strong>in</strong> units <strong>of</strong> t, <strong>in</strong> presence <strong>of</strong> impurities<br />

<strong>of</strong> b<strong>in</strong>ary type calculated us<strong>in</strong>g exact diagonalization. The impurity strength is V = −2.8t<br />

with a concentration <strong>of</strong> 10%. The system size is L 2 = 32 2 , and the SOC is Rashba type<br />

with α 2 = 1.2t with cut<strong>of</strong>f η = 0.06. (b) Comparison <strong>of</strong> a) with DOS (blue curve).<br />

KPM <strong>in</strong> a Nutshell<br />

The Kernel Polynomial Method (KPM) was first proposed by Silver et al.[SR94]<br />

to calculate DOS <strong>of</strong> large systems. It is a method to expand <strong>in</strong>tegrable functions def<strong>in</strong>ed<br />

on a f<strong>in</strong>ite <strong>in</strong>terval f : [a,b] −→ R <strong>in</strong> terms <strong>of</strong> Chebyshev polynomials <strong>of</strong> the first,<br />

T n (x) = cos(narccos(x)), (5.31)<br />

or second k<strong>in</strong>d<br />

i.e., we can write for <strong>in</strong>stance<br />

f(x) =<br />

U n (x) = s<strong>in</strong>((n+1)arccos(x)) ,<br />

s<strong>in</strong>(arccos(x))<br />

[<br />

1<br />

π √ µ 0 +2<br />

1−x 2<br />

]<br />

∞∑<br />

µ n T n (x) , (5.32)<br />

if we assume that the function f has been rescaled to ˜f : [−1,1] −→ R and can be expanded<br />

us<strong>in</strong>g the polynomials <strong>of</strong> the first k<strong>in</strong>d, which are (as the one <strong>of</strong> the second k<strong>in</strong>d) def<strong>in</strong>ed<br />

n=1<br />

on the <strong>in</strong>terval [−1,1]. If so, the coefficients are given by<br />

µ n =<br />

∫ 1<br />

−1<br />

dx ˜f(x)T n (x). (5.33)


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 95<br />

0.4<br />

0.0<br />

V<br />

Σ sH e 8Π<br />

0.3<br />

0.2<br />

0.5<br />

1.0<br />

1.5<br />

0.1<br />

2.0<br />

0.0<br />

0.02 0.04 0.06 0.08 0.10 0.12<br />

n<br />

2.5<br />

(a)<br />

Σ sH e 8Π<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

3.0<br />

3.5<br />

4.0<br />

4.5<br />

V<br />

0.00<br />

0.02 0.04 0.06 0.08 0.10 0.12<br />

n<br />

5.0<br />

(b)<br />

Figure 5.5: SHC σ SH as function <strong>of</strong> fill<strong>in</strong>g n <strong>in</strong> presence <strong>of</strong> impurities <strong>of</strong> b<strong>in</strong>ary type calculated<br />

us<strong>in</strong>g exact diagonalization. The impurity concentration is set to 10% for all plots and<br />

the average is performed over 200 impurity configurations. The system size is L 2 = 32 2 ,<br />

and the SOC is Rashba type with α 2 = 1.2t with cut<strong>of</strong>f η = 0.06 (a) V = −0.2t... −2.8t<br />

(b) V = −2.8t...−5t.


96 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

Us<strong>in</strong>g the recursion relations <strong>of</strong> the polynomials,<br />

T 0 (x) = 1, T −1 (x) = T 1 (x) = x,<br />

T m+1 (x) = 2xT m (x)−T m−1 (x), (5.34)<br />

and correspond<strong>in</strong>gly for the polynomials <strong>of</strong> the second k<strong>in</strong>d<br />

U 0 (x) = 1, U −1 (x) = 0,<br />

U m+1 (x) = 2xU m (x)−U m−1 (x), (5.35)<br />

one can calculate the expansion coefficients µ n iteratively. Replac<strong>in</strong>g the variable x by the<br />

Hamiltonian one can calculate various spectral quantities. The simplest example is the<br />

calculation <strong>of</strong> the spectral density,<br />

with the coefficients given by<br />

ρ(E) = 1 D<br />

µ n =<br />

∫ 1<br />

−1<br />

D−1<br />

∑<br />

k=0<br />

δ(E −E k ), (5.36)<br />

dxρ(x)T n [x] (5.37)<br />

= 1 D Tr[T n( ˜H)], (5.38)<br />

where ˜H is the rescaled Hamiltonian with all D eigenvalues <strong>in</strong>side the <strong>in</strong>terval [−1,1]. The<br />

efficiency <strong>of</strong> the procedure is not yet evident. This changes if one realizes the follow<strong>in</strong>g<br />

aspects:<br />

• Self averag<strong>in</strong>g properties allow for replac<strong>in</strong>g the trace over the operator by a relatively<br />

small number R ≪ D <strong>of</strong> random vectors<br />

|r〉 =<br />

D−1<br />

∑<br />

i=0<br />

ζ ri |i〉, (5.39)<br />

where the amplitudes ζ ri = e iφ are random phases on site i. This makes the effort for<br />

the calculation <strong>of</strong> M coefficients µ n l<strong>in</strong>ear <strong>in</strong> D.<br />

• The most time consum<strong>in</strong>g operation <strong>in</strong> this procedure is the matrix-vector multiplication<br />

(AppendixF). Due to the fact that the number <strong>of</strong> neighbors which a site has<br />

<strong>in</strong> the presented systems, the full multiplication can be replaced by sparse-matrix


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 97<br />

operations us<strong>in</strong>g a decomposition <strong>of</strong> the operators <strong>in</strong>to a matrix which conta<strong>in</strong>s the<br />

connectivity <strong>in</strong>formation <strong>of</strong> the sites and a matrix where the hopp<strong>in</strong>g matrices accord<strong>in</strong>g<br />

to the k<strong>in</strong>d <strong>of</strong> hopp<strong>in</strong>g (e.g. with next nearest neighbor) are stored, as expla<strong>in</strong>ed<br />

<strong>in</strong> more detail <strong>in</strong> AppendixF.<br />

Adjust<strong>in</strong>g the number <strong>of</strong> Moments <strong>in</strong> the KPM<br />

Us<strong>in</strong>g exact diagonalization to calculate e.g. SHC as presented <strong>in</strong> Sec.5.2 and<br />

Sec.5.3.1, we were forced to adjust the cut<strong>of</strong>f η <strong>in</strong> the Kubo formula, Eq.(5.5) to get the 2D<br />

limit. Otherwise we would be left with results which are highly oscillatory dependent on the<br />

Fermi energy. In theKPM the Hamiltonian is approximated by a f<strong>in</strong>itepolynomial, therefor<br />

the cut<strong>of</strong>f is implicitly set by choos<strong>in</strong>g the number <strong>of</strong> moments M. Such a cut<strong>of</strong>f, which<br />

is <strong>in</strong>evitable <strong>in</strong> numerical calculations, leads i.g. to Gibbs oscillations, especially when the<br />

expandedfunction <strong>in</strong>cludesdiscont<strong>in</strong>uities ors<strong>in</strong>gularities. Thereforetheexpandedfunction<br />

is convoluted with a particular kernel (<strong>in</strong> our case the Jackson kernel) which damps this<br />

oscillations.<br />

Exact Calculations on a f<strong>in</strong>ite lattice result <strong>in</strong> δ peaks <strong>in</strong> the DOS which are broadened<br />

due to the polynomial cut<strong>of</strong>f. The broaden<strong>in</strong>g can be approximated by Gauss curves <strong>of</strong><br />

widths σ. To do f<strong>in</strong>ite-size analysis it is crucial to keep the same number <strong>of</strong> states with<strong>in</strong><br />

the kernel, i.e. with<strong>in</strong> the distance <strong>of</strong> σ.[SSB + 10] Choos<strong>in</strong>g a number <strong>of</strong> moments M <strong>in</strong> the<br />

KPM which is too large will lead to oscillations which are due to f<strong>in</strong>ite size <strong>of</strong> the system<br />

as can be seen <strong>in</strong> Fig.(5.6). On the other hand, a too small number will smear out features<br />

<strong>of</strong> the system which are <strong>in</strong>dependent <strong>of</strong> the size. This consideration leads to the condition<br />

σ ! > ∆, with the averaged level spac<strong>in</strong>g ∆. It is helpful to analyze the relation between<br />

M and the broaden<strong>in</strong>g σ. This can be done by a convolution <strong>of</strong> a δ-distribution with the<br />

mentioned kernel, which leads to[WWAF06]<br />

M = πt<br />

σ , (5.40)<br />

and at the boundaries<br />

M =<br />

( ) πt 2/3<br />

.<br />

σ


98 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

This leads us to the follow<strong>in</strong>g condition:<br />

σ ! > ∆ (5.41)<br />

πt<br />

M > D B<br />

N<br />

with the band width D B , and the number <strong>of</strong> states N.<br />

(5.42)<br />

M < πtN<br />

D B<br />

, (5.43)<br />

To get an impression <strong>of</strong> the relation between both the η cut<strong>of</strong>f <strong>in</strong> exact diagonalization and<br />

the f<strong>in</strong>ite number <strong>of</strong> moments <strong>in</strong> the KPM, we fix the system size, apply Rashba SOC, and<br />

calculate the DOS us<strong>in</strong>g the eigenvalues E i calculated with exact diagonalization,<br />

ρ η (E) = 1 ∑<br />

[ ]<br />

1<br />

I . (5.44)<br />

π E −E λ +iη<br />

λ<br />

Now ρ can be calculated us<strong>in</strong>g KPM, and M is adjusted to fit best to ρ η (E). The relation<br />

between M and η is plotted <strong>in</strong> Fig.(5.6). Over a large <strong>in</strong>terval <strong>of</strong> M we have η ∼ 1/M.<br />

Only when the oscillations are too strong the differences between the Lorentz kernel, i.e.<br />

us<strong>in</strong>g Eq.(5.44), and the Jackson kernel appear.<br />

0.5<br />

0.10<br />

Ρ<br />

0.4<br />

0.3<br />

0.2<br />

Η<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.1<br />

0.0<br />

4 2 0 2 4<br />

E<br />

0.00<br />

0.002 0.003 0.004 0.005 0.006<br />

1<br />

M<br />

(a)<br />

(b)<br />

Figure 5.6: (a) DOS <strong>of</strong> a system <strong>of</strong> size L 2 = 40 2 with Rashba SOC, α 2 = 0.8t calculated<br />

with exact diagonalization with cut<strong>of</strong>f η = 0.0215 (blue) and KPM with M = 500 moments.<br />

(b) Relation between number <strong>of</strong> moments M and cut<strong>of</strong>f η.<br />

First application: Metal-Insulator Transition<br />

To determ<strong>in</strong>e if a 2D system has a metal-<strong>in</strong>sulator transition (MIT) it is important<br />

to analyze its symmetries: It is well known from the scal<strong>in</strong>g theory <strong>of</strong> localization that <strong>in</strong>


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 99<br />

the case where time-reversal and sp<strong>in</strong>-rotational symmetry are preserved, i.e. unitary and<br />

orthogonal universality class are present, the scal<strong>in</strong>g function[AALR79]<br />

β(g) = dln(g)<br />

dln(L) , (5.45)<br />

where g is the dimensionless conductivity, L the side length <strong>of</strong> our system, scales like<br />

β(g) ≈ − 1 g<br />

(5.46)<br />

for large g, which means that all macroscopic systems are <strong>in</strong>sulators. Here we used the<br />

scal<strong>in</strong>g parameter g = E Th /∆ LS , where E Th = D e /L 2 is the Thouless energy and ∆ LS =<br />

1/(ρ EF L 2 ) the typical energy level spac<strong>in</strong>g. However, Hikami et al. could show <strong>in</strong> Ref.<br />

[HLN80] that if the universality class changes from orthogonal to symplectic, a MIT can<br />

appear <strong>in</strong> a 2D system. Because SOC breaks sp<strong>in</strong> rotation symmetry one can show that<br />

SO <strong>in</strong>teraction can enhance the localization length ξ drastically.[AT92, KKA10, SST05]<br />

Recall<strong>in</strong>g results from transfer matrix calculations <strong>of</strong> the Anderson model <strong>in</strong> 2D with SOC,<br />

the critical disorder strength V c for the MIT is for strong Rashba SOC α 2 = 1t given by<br />

V c ≈ 6.3t[SST05, And89] and for weaker SOC α 2 = 0.1t given by V c ≈ 4.6t[SST05]. As a<br />

first application <strong>of</strong> the KPM we use the typical DOS,<br />

ρ typ (E) = exp[〈〈log(ρ i (E))〉〉], (5.47)<br />

<strong>in</strong> comparison to the local DOS<br />

ρ i (E) = 1 D<br />

D−1<br />

∑<br />

k=0<br />

|〈i|k〉| 2 δ(E −E k ), (5.48)<br />

as an <strong>in</strong>dicator for the metallic or <strong>in</strong>sulat<strong>in</strong>g regime. In contrast to the arithmetic mean <strong>of</strong><br />

ρ i , ρ avr (E) = 〈〈ρ i (E)〉〉, the geometric mean is suppressed until it vanished for V > V c : The<br />

impurity is added to the Hamiltonian by add<strong>in</strong>g the term H imp = ∑ iσ ǫ ic † iσ c iσ where the<br />

ǫ i are uniformly distributed between [−V/2,V/2]. In the follow<strong>in</strong>g we consider the Fermi<br />

energy to be at half-fill<strong>in</strong>g. In the metallic regime we have ρ = 1/(L 2 ∆ B ) at all sites,<br />

therefore we expect exp[〈〈log(ρ i (E))〉〉]/〈ρ〉 = 1.<br />

In contrast, be<strong>in</strong>g deeply <strong>in</strong> the localized regime the states decay exponentially on the


100 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

localization length ξ, |ψ(x)| 2 ∼ exp(−x/ξ), which leads to<br />

ρ exp[<br />

1<br />

∫ ( ( )]<br />

L<br />

typ πL<br />

=<br />

2 0 drr∫ 2π<br />

0<br />

dϕ ln exp − r ξ<br />

)ρ 0<br />

ρ<br />

∫ [<br />

(5.49)<br />

avr 1 L<br />

πL 2 0 drr∫ 2π<br />

0<br />

dϕ exp − r ξ<br />

]ρ 0<br />

]<br />

L 2<br />

=−<br />

=−<br />

L<br />

ξ<br />

exp[<br />

1<br />

3<br />

(<br />

2ξ L+ξ −ξexp[<br />

L<br />

ξ<br />

with u = L/ξ. From the last equation we can conclude that<br />

]) (5.50)<br />

exp [ 1<br />

3 u] u 2<br />

2(u+1−exp[u]) , (5.51)<br />

ρ<br />

• this value vanishes <strong>in</strong> the thermodynamic limit, lim typ<br />

L→∞<br />

ρ avr<br />

= 0,<br />

• the fraction is a monotone function <strong>of</strong> the system length L (<strong>in</strong> contrast to e.g. the 1d<br />

case).<br />

The expla<strong>in</strong>ed difference between the averaged and typical DOS is plotted <strong>in</strong> Fig.5.7: In (a)<br />

the averaged DOS is plotted for different impurity potentials V. The band edges are shifted<br />

to larger energies with larger V. In contrast, <strong>in</strong> (b) the typical DOS is plotted for different<br />

system sizes at impurity strength V = 8t. The product <strong>of</strong> local densities leads to a strong<br />

reduction with the size. This reduction is also significant at the band edges. Know<strong>in</strong>g the<br />

0.25<br />

0.20<br />

0.03<br />

0.15<br />

Ρaver<br />

Ρ typ<br />

0.02<br />

0.10<br />

0.05<br />

0.01<br />

0.00<br />

5 0 5<br />

0.00<br />

6 4 2 0 2 4 6<br />

E<br />

E<br />

(a)<br />

(b)<br />

Figure 5.7: (a) Averaged DOS ρ avr calculated with KPM (30 impurity configurations) for<br />

system size <strong>of</strong> L 2 = 280 2 with Rashba SOC α 2 = 0.5t, for different impurity strengths:<br />

V = 1t(black), V = 4t(red), V = 6t(green), V = 8t(blue) (b) Monotoneous reduction <strong>of</strong><br />

typical DOS ρ typ with system size L = 70, 140, 200, 280, for impurity strength V = 8t.<br />

behavior <strong>of</strong> the typical DOS ρ typ , we carried out a f<strong>in</strong>ite size analysis for different impurity


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 101<br />

strengths V <strong>in</strong> a system with weak Rashba SOC strength α 2 = 0.5t to f<strong>in</strong>d the critical value<br />

V c for the MIT. From Landauer-Bütiker calculations[SST05] it follows that at impurity<br />

strength V = 8t we are already <strong>in</strong> the <strong>in</strong>sulat<strong>in</strong>g regime. The typical DOS ρ typ decays<br />

exponentially with L, as plotted <strong>in</strong> Fig.5.8 (a), which confirms this assumption. If V is<br />

reduced the localization length ξ shows a strong <strong>in</strong>crease, as shown <strong>in</strong> Fig.5.8 (b), which<br />

<strong>in</strong> turn slows down the reduction <strong>of</strong> ρ typ , which comes along with <strong>in</strong>creas<strong>in</strong>g system size,<br />

significantly <strong>in</strong> case <strong>of</strong> large localization lengths. Add<strong>in</strong>g the results from the calculation <strong>of</strong><br />

800<br />

0.030<br />

600<br />

Ρ typ<br />

0.020<br />

Ξ a<br />

400<br />

0.015<br />

200<br />

0.010<br />

50 100 150 200 250<br />

(a)<br />

L a<br />

0<br />

0 2 4 6 8 10<br />

(b)<br />

V t<br />

Figure 5.8: (a) Log-plot <strong>of</strong> typical DOS ρ typ at V = 8t for different system sizes L with<br />

Rashba SOC α 2 = 0.5t at half fill<strong>in</strong>g, calculated with KPM. The dashed l<strong>in</strong>e is a l<strong>in</strong>ear fit<br />

to the log-data which yields a localization length <strong>of</strong> ξ ≈ 100a. (b) Localization length ξ at<br />

E F = 0, plotted as a function <strong>of</strong> impurity strength V.<br />

ρ typ /ρ avr , which is plotted <strong>in</strong> Fig.5.9 as function <strong>of</strong> the <strong>in</strong>verse system size 1/L 2 for E F = 0,<br />

we can conclude that for V 5t we are <strong>in</strong> the <strong>in</strong>sulat<strong>in</strong>g regime. For a more precise analysis<br />

we have to go to larger systems due to the large localization lengths.<br />

5.3.3 SHC calculation us<strong>in</strong>g KPM<br />

In this section we are go<strong>in</strong>g to present calculation <strong>of</strong> SHC for much larger systems<br />

than with exact diagonalization analysis, us<strong>in</strong>g KPM. Also here we will use the Kubo<br />

formalism which is why we have to reformulate Eq.(5.5) to be applicable to this iterative<br />

method. In contrast to the calculation <strong>of</strong> the DOS, here we have to deal with a correlation<br />

<strong>of</strong> two operators. We will follow the approach presented <strong>in</strong> the KPM review by Weiße et<br />

al.[WWAF06].<br />

WestartwithaKPMwhereweexpandafunctiononly<strong>in</strong>onedimension. Givenacorrelation


102 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

1.00<br />

0.70<br />

0.50<br />

Ρ typ<br />

Ρaver<br />

0.30<br />

0.20<br />

0.15<br />

0.10<br />

0.00000 0.00005 0.00010 0.00015 0.00020<br />

1<br />

L 2<br />

Figure 5.9: F<strong>in</strong>ite size analysis <strong>of</strong> the typical DOS <strong>in</strong> relation to the averaged one, ρ typ /ρ avr ,<br />

here plotted for E F = 0: The system size has been changed with L = 70, 140, 200, 280.<br />

The impurity strength for the different curves is given by V/t = 1,3,4,5,6,8 (monotone<br />

from top to bottom).<br />

function<br />

〈<br />

∣ 〉<br />

〈A;B〉 ω = 0<br />

∣ A 1 ∣∣∣<br />

ω +iǫ−H B 0 , (5.52)<br />

the imag<strong>in</strong>ary part <strong>of</strong> this function yields<br />

− 1 π I[〈A;B〉] ω =<br />

D−1<br />

∑<br />

k=0<br />

〈0|A|k〉〈k|B|0〉δ(ω −E k ) (5.53)<br />

assum<strong>in</strong>g that 〈0|A|k〉〈k|B|0〉 is real. To apply KPM, we have to rewrite this expression<br />

<strong>in</strong> terms <strong>of</strong> a trace. This can be done similar to a local DOS calculation accord<strong>in</strong>g to<br />

Eq.(5.37), which gives the coefficients<br />

µ n = 1 D<br />

D−1<br />

∑<br />

k=0<br />

∫ 1<br />

|〈i|k〉| 2 T n (Ẽk) (5.54)<br />

D−1<br />

∑<br />

=<br />

−1d˜ω 1 〈i|½|k〉〈k|½|i〉δ(˜ω −Ẽk)T n (˜ω) (5.55)<br />

D<br />

k=0<br />

= 1 〈<br />

∣<br />

i∣T n (<br />

D<br />

˜H)<br />

〉<br />

∣<br />

∣i . (5.56)


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 103<br />

Compar<strong>in</strong>g directly the <strong>in</strong>tegrand <strong>of</strong> Eq.(5.55) with Eq.(5.53) it follows that the moments<br />

<strong>of</strong> the expansion <strong>of</strong> the imag<strong>in</strong>ary part <strong>of</strong> the correlation function are given by<br />

∫ 1 D−1<br />

∑<br />

µ n = − d˜ω 〈0|A|k〉〈k|B|0〉δ(˜ω −Ẽk) T n (˜ω) (5.57)<br />

−1<br />

k=0<br />

} {{ }<br />

=<br />

≡j(˜ω)<br />

〈<br />

∣<br />

0∣AT n (− ˜H)B<br />

〉<br />

∣<br />

∣0 . (5.58)<br />

F<strong>in</strong>ally the reconstruction is done by us<strong>in</strong>g Eq.(5.32).<br />

This scheme has to be adapted to calculate the SHC σ SH <strong>in</strong> a f<strong>in</strong>ite system, as presented<br />

<strong>in</strong> Eq.(5.22),<br />

σ SH (E F ) = 2 e V<br />

∑<br />

E m


104 Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect<br />

(a)<br />

(b)<br />

Figure 5.10: a)+b) Matrix element density function j(x,y) for a clean system with 70×70<br />

sites and pure Rashba SOC <strong>of</strong> strength α 2 = 1t us<strong>in</strong>g the analytical solution Eq.(5.20).<br />

As already mentioned, we can use the great benefit to replace the trace by an average over<br />

a number R ≪ D = 2×L 2 (the two is due to sp<strong>in</strong> degree <strong>of</strong> freedom) <strong>of</strong> random vectors,<br />

Eq.(5.39). Hav<strong>in</strong>g the µ mn calculated, we can reconstruct j(x,y) accord<strong>in</strong>g to Eq.(5.32),<br />

∞∑ µ mn h mn T m (x)T n (y)<br />

j(x,y) =<br />

π 2√ (1−x 2 )(1−y 2 ) , (5.65)<br />

m,n=0<br />

where we added normalization functions<br />

h mn =<br />

4<br />

(1+δ m,0 )(1+δ n,0 ) . (5.66)<br />

On a computer we have to replace the <strong>in</strong>f<strong>in</strong>ite sum with a f<strong>in</strong>ite one where only M momenta<br />

are taken <strong>in</strong>to account. To cure the mentioned appearance <strong>of</strong> Gibbs oscillations we use<br />

a convolution with a Jackson kernel, i.e. the coefficients µ mn are replaced by µ mn →<br />

µ mn g m (M)g n (M), with kernel-damp<strong>in</strong>g factors[WWAF06]<br />

g n (M) =<br />

(M −n+1)cos(<br />

πn<br />

M+1<br />

)<br />

+s<strong>in</strong>(<br />

πn<br />

M+1<br />

M +1<br />

The function j(x,y) for f<strong>in</strong>ite M is therefore given by<br />

j M (x,y) =<br />

M−1<br />

∑<br />

m,n=0<br />

) )<br />

cot(<br />

π<br />

M+1<br />

. (5.67)<br />

µ mn h mn g m (M)g n (M)T m (x)T n (y)<br />

π 2√ . (5.68)<br />

(1−x 2 )(1−y 2 )


Chapter 5: <strong>Sp<strong>in</strong></strong> Hall Effect 105<br />

F<strong>in</strong>ally, wereconstructσ SH (E F )withthecalculatedmatrixelementdensityfunctionj M (x,y)<br />

us<strong>in</strong>g a f<strong>in</strong>ite number <strong>of</strong> moments,<br />

σ SH (ẼF,M) = e V<br />

∫ 1 ∫ 1<br />

−1<br />

−1<br />

dxdy f(x)−f(y)<br />

(y −x) 2 j M (x,y). (5.69)<br />

We set η to zero because now the divergent terms are damped by the fact that we use only<br />

a f<strong>in</strong>ite number <strong>of</strong> expansion terms M, which correspond to η as shown <strong>in</strong> Sec.5.3.2. As a<br />

presentation <strong>of</strong> the KPM we chose the same system as used for Fig.5.10. The Lorentzian<br />

function scale parameter was chosen as η = 0.07 which is why we choose M = 200 moments<br />

for the expansion, accord<strong>in</strong>g to Fig.5.11 (b). The result<strong>in</strong>g SHC σ SH (E F ,M) is plotted <strong>in</strong><br />

Fig.5.11 (red curve) and for comparison the analytical solution (blue curve). The slight<br />

asymmetry is due to an asymmetric choice <strong>of</strong> discrete k-values <strong>in</strong> the arguments <strong>of</strong> the<br />

trigonometric functions. This asymmetry disappears for larger systems.<br />

1.0<br />

0.5<br />

ΣsHe 8Π<br />

0.0<br />

0.5<br />

1.0<br />

4 2 0 2 4<br />

Figure 5.11: SHC σ SH (E F ,M) calculated with KPM for a system with 70×70 sites, pure<br />

Rashba SOC <strong>of</strong> strength α 2 = 1t and M = 200 moments (blue curve). For comparison the<br />

analytical solution is plotted (red curve), correspond<strong>in</strong>g to Fig.5.10.<br />

E F


Chapter 6<br />

Critical Discussion and Future<br />

Perspective<br />

At first we address the topic <strong>of</strong> diffusive-ballistic crossover which was discussed <strong>in</strong><br />

Sec.4.5.1: The ansatz which was used to show a reduction <strong>of</strong> sp<strong>in</strong> relaxation rate appear<strong>in</strong>g<br />

due to cubic Dresselhaus SOC, was a discretization <strong>of</strong> angles when summ<strong>in</strong>g over momenta<br />

<strong>of</strong> the Cooperon. Although the constra<strong>in</strong>t for the angles reduces the Cooperon eigenvalues<br />

significantly, this ansatz is still based on l<strong>in</strong>ear response. One consequence is that contributions<br />

appear<strong>in</strong>g at low channel number and steam<strong>in</strong>g from edge to edge skipp<strong>in</strong>g orbits, as<br />

shown <strong>in</strong> Ref.[BvH88b], are not <strong>in</strong>cluded. Such orbits can lead to flux cancellation effects,<br />

which e.g. can weaken the magnetic field dependence <strong>of</strong> WL correction to the conductivity.<br />

However, <strong>in</strong> further work we will show the reduction <strong>of</strong> sp<strong>in</strong> relaxation rate dependence on<br />

the number <strong>of</strong> transverse channels <strong>in</strong> the framework <strong>of</strong> a nonperturbative theory based on<br />

the paper by S. Kettemann et al., Ref.[KM02]. In latter work the magnetic phase-shift<strong>in</strong>g<br />

rate 1/τ B has been identified with a correlation function <strong>of</strong> the magnetic vector potential.<br />

In turn this correlation function is related to a term <strong>in</strong> the nonl<strong>in</strong>ear σ-model which appears<br />

due to time-reversal symmetry break<strong>in</strong>g. Thus, <strong>in</strong> case <strong>of</strong> an effective magnetic field due to<br />

SOC, which brakes sp<strong>in</strong> rotation symmetry, the respective term <strong>in</strong> the nonl<strong>in</strong>ear σ-model<br />

has to be identified to yield a nonperturbative expression for the sp<strong>in</strong> relaxation rate 1/τ s .<br />

The last chapter <strong>of</strong> this work stands out from the rest by the fact that the focus is more<br />

on numerical calculations. The code was developed as general as possible by decompos<strong>in</strong>g<br />

all matrix operations <strong>in</strong> connectivity and hopp<strong>in</strong>g matrices <strong>in</strong>clud<strong>in</strong>g magnetic field and<br />

106


Chapter 6: Critical Discussion and Future Perspective 107<br />

different k<strong>in</strong>ds <strong>of</strong> SOC. Hav<strong>in</strong>g coded the KPM <strong>in</strong> a general form, one <strong>of</strong> the next steps<br />

will be the analysis <strong>of</strong> diluted magnetic semiconductors us<strong>in</strong>g the V-J pd model developed<br />

by G. Bouzerar et al.[BBZ07]. Already analytical calculations show[WLZ07b] that <strong>in</strong> contrast<br />

to a lattice with only nonmagnetic impurities, the vertex correction correspond<strong>in</strong>g to<br />

ladder diagrams, σ L SH<br />

does not cancel the one-loop part σ 0 SH<br />

which is equal to that derived<br />

by S<strong>in</strong>ova.[SCN + 04] This can be calculated now rigorously for large systems, i.e. also the<br />

cluster<strong>in</strong>g <strong>of</strong> Mn 2+ can be <strong>in</strong>cluded[Bou]. Numerical works like Ref.[LX06] show <strong>in</strong>terest<strong>in</strong>g<br />

change <strong>of</strong> sign <strong>of</strong> the SHC by chang<strong>in</strong>g the exchange <strong>in</strong>teraction or the impurity density<br />

and strength. However the work was limited to smaller sizes, L 2 = 20×20, and focused on<br />

energies away from the impurity band.<br />

Due to the advantaged <strong>of</strong> the KPM concern<strong>in</strong>g the obsolete cut<strong>of</strong>f η adjustment, we are<br />

able to go beyond calculations which extracted localization lengths us<strong>in</strong>g the computation<br />

<strong>of</strong> SHC by application <strong>of</strong> the time evolution projection method (see e.g. Ref.[MM07]).<br />

Last but not least, due to the general structure <strong>of</strong> the KPM code, also the extension to the<br />

evaluation <strong>of</strong> other quantities like the anomalous Hall effect is possible.


List <strong>of</strong> Symbols<br />

SOC sp<strong>in</strong>-orbit coupl<strong>in</strong>g.......................................................2<br />

2DES 2D electron system ......................................................2<br />

m e effective electron mass .................................................. 3<br />

λ F Fermi wavelength .......................................................3<br />

WL weak localization ........................................................4<br />

WAL weak antilocalization ....................................................4<br />

m e0 free electron mass .......................................................6<br />

∆ SO Pauli splitt<strong>in</strong>g .......................................................... 7<br />

γ g gyromagnetic ratio ......................................................8<br />

τ elastic scatter<strong>in</strong>g time ...................................................9<br />

l e elastic mean free path ...................................................9<br />

D e diffusion constant .......................................................9<br />

v F Fermi velocity .......................................................... 9<br />

d D diffusion dimension <strong>of</strong> the electron system ...............................9<br />

ν density <strong>of</strong> states per sp<strong>in</strong> at E F .........................................9<br />

DOS density <strong>of</strong> states ........................................................ 9<br />

L D diffusion length ........................................................10<br />

a z thickness <strong>of</strong> the 2D electron system ....................................11<br />

γ D Dresselhaus-sp<strong>in</strong>-orbit coefficient .......................................11<br />

α 1 l<strong>in</strong>ear Dresselhaus parameter ...........................................11<br />

BIA bulk <strong>in</strong>version asymmetry ..............................................11<br />

SIA structural <strong>in</strong>version asymmetryy .......................................13<br />

α 2 l<strong>in</strong>ear Rashba parameter ...............................................13<br />

L SO sp<strong>in</strong> precession length ..................................................14<br />

108


Chapter 6: Critical Discussion and Future Perspective 109<br />

τ s sp<strong>in</strong> relaxation time ....................................................16<br />

L s sp<strong>in</strong> relaxation length ..................................................19<br />

DP D’yakonov-Perel’ .......................................................19<br />

E F Fermi energy .......................................................... 20<br />

W wire width .............................................................26<br />

l ϕ phase coherence length .................................................27<br />

Ĉ Cooperon propagator .................................................. 34<br />

˜α 1 reduced l<strong>in</strong>ear Dresselhaus parameter ..................................36<br />

l B magnetic length ........................................................37<br />

τ B magnetic phase-shift<strong>in</strong>g rate ........................................... 37<br />

Q SO sp<strong>in</strong> precession wave vector equal to 2π/L SO ...........................38<br />

H c Cooperon Hamiltonian .................................................38<br />

H SD sp<strong>in</strong>-diffusion Hamiltonian .............................................39<br />

B ∗ effective external magnetic field ........................................45<br />

H s sp<strong>in</strong> relaxation field ....................................................45<br />

∆ α sp<strong>in</strong> orbit splitt<strong>in</strong>g 2k F α 2 ..............................................53<br />

LDES low dimensional electron systems .......................................83<br />

SHE sp<strong>in</strong> Hall effect .........................................................84<br />

KPM Kernel Polynomial Method .............................................84<br />

SHC sp<strong>in</strong> Hall conductivity ................................................. 85<br />

J z sp<strong>in</strong> current ............................................................85<br />

σxy z ≡ σ SH sp<strong>in</strong> hall conductivity ..................................................88<br />

∆ B band width ............................................................98<br />

MIT metal-<strong>in</strong>sulator transition ..............................................98<br />

g scal<strong>in</strong>g parameter <strong>in</strong> scal<strong>in</strong>g theory <strong>of</strong> localization ......................99<br />

E Th Thouless energy ....................................................... 99<br />

∆ LS typical energy level spac<strong>in</strong>g ............................................ 99


List <strong>of</strong> Figures<br />

1.1 (a) a) Schematic <strong>of</strong> Datta-Das sp<strong>in</strong> modulator device <strong>in</strong> a cross-section. The<br />

2D electron gas (2DEG) has a distance <strong>of</strong> L from the emitter (↑) to the<br />

collector (↓). Normal to this cross-section there is an additional conf<strong>in</strong>ement.<br />

b) The conduction band which conf<strong>in</strong>es the electrons to a 2D system and the<br />

electron distribution (dotted) are shown. Taken from Ref.[NATE97]. (b)<br />

Manipulation <strong>of</strong> sp<strong>in</strong> precession due to SOC by gate voltage. . . . . . . . . 2<br />

2.1 Schematic representation <strong>of</strong> the band structure <strong>of</strong> GaAs around the Γ po<strong>in</strong>t<br />

(extracted from Ref.[Bur08]). . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.2 Thesp<strong>in</strong>-orbitvector fieldsforl<strong>in</strong>earstructure<strong>in</strong>versionasymmetry(Rashba)<br />

coupl<strong>in</strong>g, and for l<strong>in</strong>ear bulk <strong>in</strong>version asymmetry (BIA) sp<strong>in</strong>-orbit coupl<strong>in</strong>g<br />

for quantum wells grown <strong>in</strong> [111], [001] and [110] direction, respectively. . . 12<br />

2.3 Precession <strong>of</strong> a sp<strong>in</strong> <strong>in</strong>jected at x = 0, polarized <strong>in</strong> z-direction, as it moves<br />

by one sp<strong>in</strong> precession length L SO = π/m e α through the wire with l<strong>in</strong>ear<br />

Rashba sp<strong>in</strong>-orbit coupl<strong>in</strong>g α 2 . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.4 The sp<strong>in</strong> density for l<strong>in</strong>ear Rashba coupl<strong>in</strong>g which is a solution <strong>of</strong> the sp<strong>in</strong><br />

diffusion equation with the relaxation rate 7/16τ s . The sp<strong>in</strong> po<strong>in</strong>ts <strong>in</strong>itially<br />

<strong>in</strong> the x−y-plane <strong>in</strong> the direction (1,1,0). . . . . . . . . . . . . . . . . . . 17<br />

2.5 The sp<strong>in</strong> density for l<strong>in</strong>ear Rashba coupl<strong>in</strong>g which is a solution <strong>of</strong> the sp<strong>in</strong><br />

diffusion equation with the relaxation rate 1/τ s = 7/16τ s0 . Note that, compared<br />

to the ballistic sp<strong>in</strong> density, Fig.2.3, the period is slightly enhanced by<br />

a factor 4/ √ 15. Also, the amplitude <strong>of</strong> the sp<strong>in</strong> density changes with the position<br />

x, <strong>in</strong> contrast to the ballistic case. The color is chang<strong>in</strong>g <strong>in</strong> proportion<br />

to the sp<strong>in</strong> density amplitude. . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.6 Elastic scatter<strong>in</strong>g from impurities changes the direction <strong>of</strong> the sp<strong>in</strong>-orbit field<br />

around which the electron sp<strong>in</strong> is precess<strong>in</strong>g. . . . . . . . . . . . . . . . . . 19<br />

3.1 Two different experimental approaches to extract the wire width dependence<br />

<strong>of</strong> sp<strong>in</strong> relaxation rate. (a) Measurement by Kerr rotation (extracted from<br />

[HSM + 06]) and (b) us<strong>in</strong>g magnetoconductivity experiments (extracted from<br />

[LSK + 07]). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3.2 Measured WL corrections on a th<strong>in</strong> PdAu film, Ref.[DO79]. The resistivity<br />

<strong>in</strong>creases logarithmically as the temperature decreases. . . . . . . . . . . . . 29<br />

110


List <strong>of</strong> Figures 111<br />

3.3 Exemplification <strong>of</strong> the second term <strong>in</strong> Eq.(3.4): Interference <strong>of</strong> electrons<br />

travel<strong>in</strong>g <strong>in</strong> the opposite direction along the same path causes an enhanced<br />

back-scatter<strong>in</strong>g, the WL effect. (a) Closed electron paths enclose a magnetic<br />

fluxfromanexternalmagneticfield, <strong>in</strong>dicatedastheredarrow, break<strong>in</strong>gtime<br />

reversal symmetry, break<strong>in</strong>g constructive <strong>in</strong>terference. (b) The entanglement<br />

<strong>of</strong> sp<strong>in</strong> and charge by SO <strong>in</strong>teraction causes the sp<strong>in</strong> to precess <strong>in</strong>between<br />

two scatterers around an axis which changes with the momentum vector <strong>of</strong><br />

the it<strong>in</strong>erant electron. This effective field can cause WAL. . . . . . . . . . . 30<br />

3.4 2D spectrum <strong>of</strong> H c , K i = Q i /Q SO . The physical mean<strong>in</strong>g <strong>of</strong> the gaps <strong>in</strong> the<br />

triplet modes is more comprehensible if H c is related to sp<strong>in</strong> diffusion, where<br />

the gaps appear as sp<strong>in</strong> relaxation rates. . . . . . . . . . . . . . . . . . . . . 38<br />

3.5 Persistent sp<strong>in</strong> helix solution <strong>of</strong> the sp<strong>in</strong>-diffusion equation for equal magnitude<br />

<strong>of</strong> l<strong>in</strong>ear Rashba and l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g, Eq.(3.68). . . . . . 42<br />

3.6 Longpersist<strong>in</strong>gsp<strong>in</strong>helixsolution<strong>of</strong>thesp<strong>in</strong>-diffusionequation<strong>in</strong>aquantum<br />

wirewhosewidthW issmaller thanthesp<strong>in</strong>precessionlengthL SO forvary<strong>in</strong>g<br />

ratio <strong>of</strong> l<strong>in</strong>ear Rashba α 2 = αs<strong>in</strong>ϕ and l<strong>in</strong>ear Dresselhaus coupl<strong>in</strong>g, α 1 =<br />

αcosϕ, Eq.(3.82), for fixed α and L SO = π/m e α. . . . . . . . . . . . . . . . 46<br />

3.7 Dispersion <strong>of</strong> the triplet Cooperon modes for different dimensionless wire<br />

units Q SO W: (a) Q SO W = 2, (b) Q SO W = 8, (c) Q SO W = 12, (d) Q SO W =<br />

30, plotted as function <strong>of</strong> K x = Q x /Q SO . For Q SO W ≫ 3, E {t0,0} and E {t−,0}<br />

evolve <strong>in</strong>to degenerate branches for large K x . (For Q SO W = 30, not all<br />

high-energy branches are shown.) . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.8 Probability density <strong>of</strong> the Cooperon eigenmodes <strong>in</strong> the wire for Q SO W/π =<br />

30. (a) 3D plot, (b) density plot for one <strong>of</strong> the two lowest branches, show<strong>in</strong>g<br />

their edge mode character. (c) 3D plot and (d) density plot <strong>of</strong> the density <strong>of</strong><br />

the third lowest mode, which shows bulk character. . . . . . . . . . . . . . . 49<br />

3.9 Absolute m<strong>in</strong>ima <strong>of</strong> the lowest eigenmodes E {t0,0} , E {t−,0} , and E {t+,0} plotted<br />

as function <strong>of</strong> Q SO W/π = 2W/L SO . We note that the m<strong>in</strong>imum <strong>of</strong> E {t−,0}<br />

is located at ±K x ≠ 0. For comparison, the solution <strong>of</strong> the zero-mode approximation<br />

E t0 is shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.10 Lowest eigenvalues at K x = 0 plotted aga<strong>in</strong>st Q SO W/π. For comparison, the<br />

global m<strong>in</strong>imum <strong>of</strong> the Cooperon spectrum for Q SO W 9 is plotted, F 3 .<br />

Curves F 1 [n] are given by 7/16+ ( (n/(Q SO W/π)) √ 15/4 ) 2<br />

, n ∈ N. F2 shows<br />

the energy m<strong>in</strong>imum <strong>of</strong> the 2D case, F 2 ≡ F 1 [n = 0]. Vertical dotted l<strong>in</strong>es<br />

<strong>in</strong>dicate the widths at which the lowest two branches degenerate at K x = 0.<br />

They aregiven by n/( √ 15/4); consider that the wave vector for them<strong>in</strong>imum<br />

<strong>of</strong> the E T− mode is ( √ 15/4)Q SO . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.11 The quantum conductivity correction <strong>in</strong> units <strong>of</strong> 2e 2 /2π as function <strong>of</strong> magnetic<br />

field B (scaled with bulk relaxation field H s ), and the wire width W<br />

scaled with 1/Q SO for pure Rashba coupl<strong>in</strong>g and cut<strong>of</strong>fs 1/D e Q 2 τ SO ϕ = 0.08,<br />

1/D e Q 2 SOτ = 4: Comparison <strong>of</strong> thezero-mode calculation (grid without shad<strong>in</strong>g)<br />

to the exact diagonalization where the lowest 21 triplet branches and<br />

seven s<strong>in</strong>glet branches were taken <strong>in</strong>to account. . . . . . . . . . . . . . . . . 52


112 List <strong>of</strong> Figures<br />

3.12 The relative magnetoconductivity ∆σ(B) − ∆σ(B = 0) <strong>in</strong> units <strong>of</strong> 2e 2 /2π,<br />

with the same parameters and number <strong>of</strong> modes as <strong>in</strong> Fig.3.11. . . . . . . . 53<br />

3.13 Width <strong>of</strong> wire WQ SO at which there is a crossover from negative to positive<br />

magnetoconductivity as function <strong>of</strong> the lower cut<strong>of</strong>f 1/D e Q 2 τ SO ϕ. . . . . . . 54<br />

3.14 Spectrum <strong>in</strong> the case <strong>of</strong> adiabatic boundaries. Except for states which are<br />

polarized <strong>in</strong> z direction, the relaxation rates for all states diverge at small<br />

wire widths, Q SO W ≪ 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.15 (a) Real and (b) imag<strong>in</strong>ary parts <strong>of</strong> the spectrum <strong>of</strong> the 2D Cooperon with<br />

Zeeman term <strong>of</strong> strength gµ B B/D e Q 2 = 0.4. E SO B,2D,1 (black), E B,2D,2 (red<br />

dashed), E B,2D,3 (green), E B,2D,4 (blue dashed). Dashed vertical l<strong>in</strong>es are<br />

located at K x = ±1/ √ 2, the wave vector where the triplet mode E T− and<br />

the s<strong>in</strong>glet mode E S are cross<strong>in</strong>g each other (without loss <strong>of</strong> generality K y = 0). 59<br />

3.16 (a) Real and (b) imag<strong>in</strong>ary parts <strong>of</strong> the spectrum <strong>of</strong> the 2D Cooperon with<br />

Zeeman term <strong>of</strong> the strength gµ B B/D e Q 2 SO<br />

= 0...2 <strong>in</strong> steps <strong>of</strong> 0.25 (w.l.o.g<br />

K y = 0). The B <strong>in</strong>dependent mode E T0 is not shown.. . . . . . . . . . . . . 61<br />

3.17 (a) Real and (b) imag<strong>in</strong>ary parts <strong>of</strong> the spectrum <strong>of</strong> the Cooperon with<br />

Zeeman term <strong>of</strong> the strength gµ B B/D e Q 2 SO<br />

= 0...0.8 <strong>in</strong> steps <strong>of</strong> 0.1 <strong>in</strong> a<br />

f<strong>in</strong>ite wire <strong>of</strong> the width Q SO W = 0.5. The B <strong>in</strong>dependent mode E t0 is not<br />

shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.18 The magnetoconductance ∆σ(B) <strong>in</strong> a magnetic field perpendicular to the<br />

quantum well, where its coupl<strong>in</strong>g via the Zeeman term is considered by exact<br />

diagonalization while its effect on the orbital motion is considered effectively<br />

by the magnetic phase-shift<strong>in</strong>g rate 1/τ B (W) for wire width Q SO W = 1,<br />

dephas<strong>in</strong>grate 1/τ ϕ = 0.06D e Q 2 SO, andelastic-scatter<strong>in</strong>g rate 1/τ = 4D e Q 2 SO.<br />

Thestrength<strong>of</strong>thecontribution<strong>of</strong>theZeemantermisvariedbythematerialdependent<br />

factor ˜g = gµ B H s /D e Q 2 SO<br />

<strong>in</strong> the range ˜g = 0...1.5 <strong>in</strong> steps <strong>of</strong> 0.5:<br />

The system changes from positive magnetoconductivity (˜g = 0, green) to<br />

negative one (˜g = 0.5...1.5, cont<strong>in</strong>uous decrease <strong>of</strong> the absolute m<strong>in</strong>imum). 64<br />

3.19 (a) Change <strong>of</strong> crossover width W c with g factor: The magnetic field is <strong>in</strong>cluded<br />

as an effective field 1/τ B and <strong>in</strong> the Zeeman term. The strength <strong>of</strong> the<br />

contribution <strong>of</strong> the Zeeman term is varied by the material dependent factor<br />

˜g = gµ B H s /D e Q 2 . (b) Change <strong>of</strong> crossover width W SO c with Zeeman field:<br />

To calculate W c , we fix the Zeeman field to a certa<strong>in</strong> value, horizontal axis,<br />

while we vary the effective field <strong>in</strong>dependently and calculate if negative or<br />

positive magnetoconductivity is present. For different Zeeman fields B Z /H s ,<br />

we f<strong>in</strong>d thereby a different width W c . Here, we set g/8m e D e = 1. In (a)<br />

and (b), the cut<strong>of</strong>f due to dephas<strong>in</strong>g is varied: 1/D e Q 2 SOτ ϕ ∈<br />

= 0.04,0.06,0.08<br />

(lowest first). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

4.1 Dependence <strong>of</strong> the W 2 coefficient <strong>in</strong> Eq.(4.27) on the lateral rotation (θ).<br />

The absolute m<strong>in</strong>imum is found for α x1 (θ = 0) = −q 2 (here: q 1 /q 2 = 1.63)<br />

and for different SO strength we f<strong>in</strong>d the m<strong>in</strong>imum at θ = (1/4+n)π, n<br />

if q 1 < (q s3 / √ 2) (dashed l<strong>in</strong>e: q 1 = (q s3 / √ 2)) and at θ = (3/4+n)π, n<br />

else. Here we set q s3 = 0.9. The scal<strong>in</strong>g is arbitrary. . . . . . . . . . . . . . 73


List <strong>of</strong> Figures 113<br />

4.2 The sp<strong>in</strong> dephas<strong>in</strong>g time T 2 <strong>of</strong> a sp<strong>in</strong> <strong>in</strong>itially oriented along the [001] direction<br />

<strong>in</strong> units <strong>of</strong> (D e q2 2 ) for the special case <strong>of</strong> equal Rashba and l<strong>in</strong>. Dresselhaus<br />

SOC. The different curves show different strength <strong>of</strong> cubic Dresselhaus<br />

<strong>in</strong> units <strong>of</strong> q s3 /q 2 . In the case <strong>of</strong> f<strong>in</strong>ite cubic Dresselhaus SOC we set<br />

W = 0.4/q 2 . If q s3 = 0: T 2 diverges at θ = (1/4 + n)π, n ∈(dashed<br />

vertical l<strong>in</strong>es). The horizontal dashed l<strong>in</strong>e <strong>in</strong>dicated the 2D sp<strong>in</strong> dephas<strong>in</strong>g<br />

time, T 2 = 1/(4q2 2D e). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.3 Example <strong>of</strong> an experiment done by Kunihashi et al.[KKN09]. The width dependence<br />

<strong>of</strong> the sp<strong>in</strong> relaxation length l 1D<br />

SO<br />

<strong>of</strong> different carrier density. Solid<br />

l<strong>in</strong>es and dashed l<strong>in</strong>es <strong>in</strong> (b) show the l 1D<br />

SO<br />

calculated from the theory presented<br />

<strong>in</strong> this work, with neglect<strong>in</strong>g cubic Dresselhaus term and tak<strong>in</strong>g <strong>in</strong>to<br />

account full SOIs, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

4.4 The lowest eigenvalues <strong>of</strong> the conf<strong>in</strong>ed Cooperon Hamiltonian Eq.(4.49),<br />

equivalent to the lowest sp<strong>in</strong> relaxation rate, are shown for Q = 0 for different<br />

number <strong>of</strong> modes N = k F W/π. Different curves correspond to different<br />

values <strong>of</strong> α 2 /q s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.5 The lowest eigenvalues <strong>of</strong> the conf<strong>in</strong>ed Cooperon Hamiltonian Eq.(4.49),<br />

equivalent to the lowest sp<strong>in</strong> relaxation rate, are shown for Q = 0 for different<br />

number <strong>of</strong> modes N = k F W/π. Different curves correspond to different<br />

values <strong>of</strong> α 1 /q s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.1 Energy band E − (k), Eq.(5.10) is plotted for pure Rashba SOC as function<br />

<strong>of</strong> wave vector k. The contour l<strong>in</strong>es <strong>in</strong>dicate the energy at which one f<strong>in</strong>ds a<br />

Van Hove s<strong>in</strong>gularity <strong>in</strong> the DOS (below half-fill<strong>in</strong>g). . . . . . . . . . . . . . 87<br />

5.2 (a) SHC σ SH as a function <strong>of</strong> Fermi energy E F , <strong>in</strong> a clean system <strong>of</strong> size<br />

L 2 = 170×170 with both Rashba and l<strong>in</strong>ear Dresselhaus SOC with α 2 > α 1<br />

(blue curve) and α 2 < α 1 (red curve). (b) SHC σ SH as a function <strong>of</strong> Fermi<br />

energy E F , <strong>in</strong> a clean system <strong>of</strong> size V = 150×150 with only Rashba SOC<br />

<strong>of</strong> strength α 2 = 0.8t (blue/solid), α 2 = 1.4t (red/dotted) and α 2 = 2t<br />

(yellow/dashed). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

5.3 DOS, as a function <strong>of</strong> Fermi energy <strong>in</strong> units <strong>of</strong> t, <strong>in</strong> presence <strong>of</strong> impurities<br />

<strong>of</strong> b<strong>in</strong>ary type with a concentration <strong>of</strong> 10% calculated us<strong>in</strong>g exact diagonalization.<br />

The system size is L 2 = 32 2 , and the SOC is Rashba type with<br />

α 2 = 1.2t with cut<strong>of</strong>f η = 0.02t. . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

5.4 (a) SHC, as a function <strong>of</strong> Fermi energy <strong>in</strong> units <strong>of</strong> t, <strong>in</strong> presence <strong>of</strong> impurities<br />

<strong>of</strong> b<strong>in</strong>ary type calculated us<strong>in</strong>g exact diagonalization. The impurity strength<br />

is V = −2.8t with a concentration <strong>of</strong> 10%. The system size is L 2 = 32 2 , and<br />

the SOC is Rashba type with α 2 = 1.2t with cut<strong>of</strong>f η = 0.06. (b) Comparison<br />

<strong>of</strong> a) with DOS (blue curve). . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

5.5 SHC σ SH as function <strong>of</strong> fill<strong>in</strong>g n <strong>in</strong> presence <strong>of</strong> impurities <strong>of</strong> b<strong>in</strong>ary type calculated<br />

us<strong>in</strong>g exact diagonalization. The impurity concentration is set to 10%<br />

for all plots and the average is performed over 200 impurity configurations.<br />

The system size is L 2 = 32 2 , and the SOC is Rashba type with α 2 = 1.2t<br />

with cut<strong>of</strong>f η = 0.06 (a) V = −0.2t...−2.8t (b) V = −2.8t...−5t. . . . . 95


114 List <strong>of</strong> Figures<br />

5.6 (a) DOS <strong>of</strong> a system <strong>of</strong> size L 2 = 40 2 with Rashba SOC, α 2 = 0.8t calculated<br />

with exact diagonalization with cut<strong>of</strong>f η = 0.0215 (blue) and KPM with<br />

M = 500 moments. (b) Relation between number <strong>of</strong> moments M and cut<strong>of</strong>f η. 98<br />

5.7 (a) Averaged DOS ρ avr calculated with KPM (30 impurity configurations)<br />

for system size <strong>of</strong> L 2 = 280 2 with Rashba SOC α 2 = 0.5t, for different<br />

impurity strengths: V = 1t(black), V = 4t(red), V = 6t(green), V =<br />

8t(blue) (b) Monotoneous reduction <strong>of</strong> typical DOS ρ typ with system size<br />

L = 70, 140, 200, 280, for impurity strength V = 8t. . . . . . . . . . . . . . 100<br />

5.8 (a) Log-plot <strong>of</strong> typical DOS ρ typ at V = 8t for different system sizes L with<br />

Rashba SOC α 2 = 0.5t at half fill<strong>in</strong>g, calculated with KPM. The dashed l<strong>in</strong>e<br />

isal<strong>in</strong>earfittothelog-data whichyieldsalocalization length<strong>of</strong>ξ ≈ 100a. (b)<br />

Localization length ξ at E F = 0, plotted as a function <strong>of</strong> impurity strength V.101<br />

5.9 F<strong>in</strong>ite size analysis <strong>of</strong> the typical DOS <strong>in</strong> relation to the averaged one,<br />

ρ typ /ρ avr , here plotted for E F = 0: The system size has been changed with<br />

L = 70, 140, 200, 280. The impurity strength for the different curves is given<br />

by V/t = 1,3,4,5,6,8 (monotone from top to bottom). . . . . . . . . . . . . 102<br />

5.10 a)+b) Matrix element density function j(x,y) for a clean system with 70×70<br />

sites and pure Rashba SOC <strong>of</strong> strength α 2 = 1t us<strong>in</strong>g the analytical solution<br />

Eq.(5.20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

5.11 SHC σ SH (E F ,M) calculated with KPM for a system with 70×70 sites, pure<br />

Rashba SOC <strong>of</strong> strength α 2 = 1t and M = 200 moments (blue curve). For<br />

comparison the analytical solution is plotted (red curve), correspond<strong>in</strong>g to<br />

Fig.5.10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

C.1 Weak localization correction <strong>in</strong> 2D <strong>in</strong> units <strong>of</strong> (2e 2 /2π). The parameters are<br />

c 1 = 1/D e Q 2 SOτ ϕ and c 2 = 1/D e Q 2 SOτ. Thick l<strong>in</strong>e <strong>in</strong>dicates ∆σ = 0. . . . . . 143


List <strong>of</strong> Tables<br />

3.1 S<strong>in</strong>glet and triplet states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

A.1 Values <strong>of</strong> Rashba parameter α 2 measured <strong>in</strong> experiments [QW=quantum<br />

wire]. (List extracted from[FMAE + 07].) . . . . . . . . . . . . . . . . . . . . 130<br />

A.2 Experimental values <strong>of</strong> m e [VMRM01] . . . . . . . . . . . . . . . . . . . . . 130<br />

A.3 Values <strong>of</strong> Dresselhaus parameter γ D measured <strong>in</strong> experiments. (List extracted<br />

from[FMAE + 07].) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

A.4 Measured α 2 /α 1 ratios with sp<strong>in</strong>-galvanic effect (SGE) and Circular photogalvaniceffect(CPGE)withtheaccord<strong>in</strong>gparametersat4.2K<br />

[QW=quantum<br />

wire].[GGB + 07] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

115


Bibliography<br />

[AAKL82]<br />

[AALR79]<br />

[AB59]<br />

B.L.Altshuler, A.G. Aronov, D. E.Khmelnitskii, andA.I.Lark<strong>in</strong>. Quantum<br />

Theory <strong>of</strong> Solids. Mir, Moscow, 1982.<br />

E. Abrahams, P. W. Anderson, D. C. Licciardello, and T. V. Ramakrishnan.<br />

Scal<strong>in</strong>g theory <strong>of</strong> localization: Absence <strong>of</strong> quantum diffusion <strong>in</strong> two<br />

dimensions. Phys. Rev. Lett., 42(10):673–676, Mar 1979.<br />

Y. Aharonov and D. Bohm. Significance <strong>of</strong> electromagnetic potentials <strong>in</strong> the<br />

quantum theory. Phys. Rev., 115(3):485–491, Aug 1959.<br />

[ADK + 08] G. V. Astakhov, R. I. Dzhioev, K. V. Kavok<strong>in</strong>, V. L. Korenev, M. V.<br />

Lazarev, M. N. Tkachuk, Yu. G. Kusrayev, T. Kiessl<strong>in</strong>g, W. Ossau, and<br />

L. W. Molenkamp. Suppression <strong>of</strong> Electron <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Mn-Doped<br />

GaAs. Phys. Rev. Lett., 101(7):076602, 2008.<br />

[AF01]<br />

[AF07]<br />

I. L. Ale<strong>in</strong>er and Vladimir I. Fal’ko. <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g Effects on Quantum<br />

Transport <strong>in</strong> Lateral Semiconductor Dots. Phys. Rev. Lett., 87(25):256801,<br />

Nov 2001.<br />

David D Awschalom and Michael E Flatte. Challenges for semiconductor<br />

sp<strong>in</strong>tronics. Nat Phys, 3(3):153–159, Mar 2007.<br />

[Ale06] I. L. Ale<strong>in</strong>er. privtate communication. 2006.<br />

[And89]<br />

T. Ando. Numerical study <strong>of</strong> symmetry effects on localization <strong>in</strong> two dimensions.<br />

Phys. Rev. B, 40(8):5325–5339, Sep 1989.<br />

[AOMO01] T. Adachi, Y. Ohno, F. Matsukura, and H. Ohno. <strong>Sp<strong>in</strong></strong> relaxation <strong>in</strong> n-<br />

modulationdopedGaAs/AlGaAs (110) quantumwells. Physica E: Low-dim.<br />

Sys. and Nanostr., 10(1-3):36–39, 2001.<br />

[AT92]<br />

[BAP76]<br />

T. Ando and H. Tamura. Conductance fluctuations <strong>in</strong> quantum wires with<br />

sp<strong>in</strong>-orbit and boundary-roughness scatter<strong>in</strong>g. Phys. Rev. B, 46(4):2332–<br />

2338, Jul 1992.<br />

G. L. Bir, A. G. Aronov, and G. E. Pikus. <strong>Sp<strong>in</strong></strong> Relaxation <strong>of</strong> Electrons due<br />

to Scatter<strong>in</strong>g by Holes. Sov. Phys.-JETP, 42:705, 1976.<br />

116


Bibliography 117<br />

[BB04]<br />

[BBF + 88]<br />

[BBZ07]<br />

[Ber84]<br />

[BEW + 99]<br />

[BGSZ89]<br />

[BNnM04]<br />

[Bog80]<br />

[Bou]<br />

[BOZ06]<br />

[BPP48]<br />

[BR84]<br />

[Bur08]<br />

[BvH88a]<br />

A. A. Burkov and Leon Balents. <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> a Two-Dimensional<br />

ElectronGas <strong>in</strong>aPerpendicularMagnetic Field. Phys. Rev. B,69(24):245312,<br />

Jun 2004.<br />

M. Baibich, J. Broto, A. Fert, F. Van Dau, F. Petr<strong>of</strong>f, P. Etienne, G. Creuzet,<br />

A. Friederich, and J. Chazelas. Giant Magnetoresistance <strong>of</strong> (001)Fe/(001)Cr<br />

Magnetic Superlattices. Phys. Rev. Lett., 61(21):2472–2475, Nov 1988.<br />

R Bouzerar, G Bouzerar, and T Ziman. Non-perturbative V-Jpd model and<br />

ferromagnetism <strong>in</strong> dilute magnets. EPL (Europhysics Letters), 78(6):67003,<br />

2007.<br />

Gerd Bergmann. Weak Localization <strong>in</strong> Th<strong>in</strong> Films : a Time-<strong>of</strong>-Flight Experiment<br />

with Conduction Electrons. Physics Reports, 107(1):1–58, 1984.<br />

S. Brosig, K. Enssl<strong>in</strong>, R. J. Warburton, C. Nguyen, B. Brar, M. Thomas, and<br />

H. Kroemer. Zero-field<strong>Sp<strong>in</strong></strong>Splitt<strong>in</strong>g<strong>in</strong> InAs-AlSbQuantumWells Revisited.<br />

Phys. Rev. B, 60(20):R13989–R13992, Nov 1999.<br />

G. B<strong>in</strong>asch, P. Grünberg, F. Saurenbach, and W. Z<strong>in</strong>n. Enhanced magnetoresistance<br />

<strong>in</strong> layered magnetic structures with antiferromagnetic <strong>in</strong>terlayer<br />

exchange. Phys. Rev. B, 39(7):4828–4830, Mar 1989.<br />

A.A. Burkov, AlvaroS.Núñez, andA. H.MacDonald. Theory<strong>of</strong> sp<strong>in</strong>-chargecoupled<br />

transport <strong>in</strong> a two-dimensional electron gas with rashba sp<strong>in</strong>-orbit<br />

<strong>in</strong>teractions. Phys. Rev. B, 70(15):155308, Oct 2004.<br />

P. Boguslawski. Electron-Electron <strong>Sp<strong>in</strong></strong>-Flip Scatter<strong>in</strong>g and <strong>Sp<strong>in</strong></strong> Relaxation<br />

<strong>in</strong> III-V and II-VI Semiconductors. Solid State Commun., 33:389, 1980.<br />

G. Bouzerar. priv. communication, <strong>in</strong> preparation.<br />

B. Andrei Bernevig, J. Orenste<strong>in</strong>, and Shou-ChengZhang. Exact SU(2) Symmetry<br />

and Persistent <strong>Sp<strong>in</strong></strong> Helix <strong>in</strong> a <strong>Sp<strong>in</strong></strong>-Orbit Coupled System. Phys. Rev.<br />

Lett., 97(23):236601, 2006.<br />

N. Bloembergen, E. M. Purcell, and R. V. Pound. Relaxation Effects <strong>in</strong><br />

Nuclear Magnetic Resonance Absorption. Phys. Rev., 73(7):679–712, Apr<br />

1948.<br />

Ya Bychkov and Ei Rashba. Properties <strong>of</strong> A 2d Electron-Gas With Lifted<br />

Spectral Degeneracy. JETP Lett., 39(2):78–81, 1984.<br />

Guido Burkard. Quantum <strong>in</strong>formation: positively sp<strong>in</strong> coherent. Nature<br />

materials, 7(2):100–1, Feb 2008.<br />

C. W. J. Beenakker and H. van Houten. Flux-Cancellation Effect on Narrow-<br />

Channel Magnetoresistance Fluctuations. Phys. Rev. B, 37(11):6544–6546,<br />

Apr 1988.


118 Bibliography<br />

[BvH88b]<br />

[Cha75]<br />

[CS86]<br />

[CSS + 04]<br />

[CWd07]<br />

[DD90]<br />

[DHR + 04]<br />

[DK84]<br />

[DKK + 02]<br />

C. W. J. Beenakker and H. van Houten. Flux-cancellation effect on narrowchannel<br />

magnetoresistance fluctuations. Phys. Rev. B, 37(11):6544–6546, Apr<br />

1988.<br />

J. N. Chazalviel. <strong>Sp<strong>in</strong></strong> Relaxation <strong>of</strong> Conduction Electrons <strong>in</strong> n-Type Indium<br />

Antimonide at Low Temperature. Phys. Rev. B, 11(4):1555–1562, Feb 1975.<br />

Sudip Chakravarty and Albert Schmid. Weak Localization: The Quasiclassical<br />

Theory <strong>of</strong> Electrons <strong>in</strong> a Random Potential. Physics Reports, 140(4):193–<br />

236, 1986.<br />

Dimitrie Culcer, Jairo S<strong>in</strong>ova, N. A. S<strong>in</strong>itsyn, T. Jungwirth, A. H. MacDonald,<br />

and Q. Niu. Semiclassical sp<strong>in</strong> transport <strong>in</strong> sp<strong>in</strong>-orbit-coupled bands.<br />

Phys. Rev. Lett., 93(4):046602, Jul 2004.<br />

J. Cheng, M. Wu, and I. da Cunha Lima. Anisotropic sp<strong>in</strong> transport <strong>in</strong><br />

GaAs quantum wells <strong>in</strong> the presence <strong>of</strong> compet<strong>in</strong>g Dresselhaus and Rashba<br />

sp<strong>in</strong>-orbit coupl<strong>in</strong>g. Phys. Rev. B, 75(20):205328, May 2007.<br />

Supriyo Datta and Biswajit Das. Electronic Analog <strong>of</strong> The Electro-Optic<br />

Modulator. Appl. Phys. Lett., 56(7):665–667, 1990.<br />

S.Döhrmann, D. Hägele, J.Rudolph, M.Bichler, D. Schuh, andM. Oestreich.<br />

Anomalous <strong>Sp<strong>in</strong></strong> Dephas<strong>in</strong>g <strong>in</strong> (110) GaAs Quantum Wells: Anisotropy and<br />

Intersubband Effects. Phys. Rev. Lett., 93(14), September 2004.<br />

V.K.DugaevandD.E.Khmel’nitskii. Magnetoresistance <strong>of</strong>MetalFilmswith<br />

Low Impurity Concentrations <strong>in</strong> a Parallel Magnetic Field. Soviet Physics -<br />

JETP, 59(5):1038–1041, 1984.<br />

R. I. Dzhioev, K. V. Kavok<strong>in</strong>, V. L. Korenev, M. V. Lazarev, B. Ya. Meltser,<br />

M. N. Stepanova, B. P. Zakharchenya, D. Gammon, and D. S. Katzer. Low-<br />

Temperature <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> n-Type GaAs. Phys. Rev. B, 66(24):245204,<br />

Dec 2002.<br />

[DLS + 05] R. D<strong>in</strong>ter, S. Löhr, S. Schulz, Ch. Heyn, and W. Hansen. unpublished, 2005.<br />

[DMD + 89]<br />

[DO79]<br />

[DP71a]<br />

B Das, D C Miller, S Datta, R Reifenberger, Hong W P., P K<br />

Bhattacharya, J S<strong>in</strong>gh, and M Jaffe. Evidence for <strong>Sp<strong>in</strong></strong> Splitt<strong>in</strong>g <strong>in</strong><br />

In x Ga 1−x As/In 0.52 Al 0.48 As Heterostructures as B → 0. Phys. Rev. B,<br />

39(2):1411–1414, January 1989.<br />

G. Dolan and D. Osher<strong>of</strong>f. Nonmetallic Conduction <strong>in</strong> Th<strong>in</strong> Metal Films at<br />

Low Temperatures. Phys. Rev. Lett., 43(10):721–724, Sep 1979.<br />

M. I. D’yakonov and V. I. Perel’. Possibility <strong>of</strong> Orient<strong>in</strong>g Electron <strong>Sp<strong>in</strong></strong>s with<br />

Current . JETP Lett., 13(11):467, 1971.


Bibliography 119<br />

[DP71b]<br />

[DP71c]<br />

[DP72]<br />

[DPWS92]<br />

[DR93]<br />

[DR04]<br />

[Dre55]<br />

M I D’yakonov and V I Perel’. <strong>Sp<strong>in</strong></strong> Orientation <strong>of</strong> Electrons Associated with<br />

Interband Absorption <strong>of</strong> Light <strong>in</strong> Semiconductors. Soviet Physics Jetp-Ussr,<br />

33(5):1053–&, 1971. [Zh. Eksp. Teor. Fiz., 60:1954, 1971].<br />

M. I. D’yakonov and V. I. Perel’. <strong>Sp<strong>in</strong></strong> Relaxation <strong>of</strong> Conduction Electrons <strong>in</strong><br />

Noncentrosymmetric Semiconductors. Fiz. Tverd. Tela, 13:3581, 1971. [Sov.<br />

Phys. Solid State 13, 3023-3026 (1971)].<br />

M I D’yakonov and V I Perel’. <strong>Sp<strong>in</strong></strong> Relaxation <strong>of</strong> Conduction Electrons <strong>in</strong><br />

Noncentrosymmetric Semiconductors. Sov. Phys. Solid State, 13:3023–3026,<br />

1972.<br />

P. D. Dresselhaus, C. M. A. Papavassiliou, R. G. Wheeler, and R. N. Sacks.<br />

Observation <strong>of</strong> <strong>Sp<strong>in</strong></strong> Precession <strong>in</strong> GaAs Inversion Layers Us<strong>in</strong>g Antilocalization.<br />

Phys. Rev. Lett., 68(1):106–109, Jan 1992.<br />

T. Darnh<strong>of</strong>er and U. Rössler. Effects <strong>of</strong> Band Structureand <strong>Sp<strong>in</strong></strong> <strong>in</strong> Quantum<br />

Dots. Phys. Rev. B, 47(23):16020–16023, Jun 1993.<br />

A. Dyson and B. K. Ridley. <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Cubic III-V Semiconductors<br />

via Interaction with Polar Optical Phonons. Phys. Rev. B, 69(12):125211,<br />

Mar 2004.<br />

G. Dresselhaus. <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g Effects <strong>in</strong> Z<strong>in</strong>c Blende <strong>Structures</strong>. Phys.<br />

Rev., 100(2):580–586, Oct 1955.<br />

[Dya04] M. I. Dyakonov. <strong>Sp<strong>in</strong></strong>tronics? cond-mat/0401369, Jan 2004.<br />

[Ell54]<br />

[ELSL97]<br />

[ERH07]<br />

[ESL05]<br />

R. J. Elliott. Theory <strong>of</strong> the Effect <strong>of</strong> <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g on Magnetic Resonance<br />

<strong>in</strong> Some Semiconductors. Phys. Rev., 96(2):266–279, Oct 1954.<br />

G. Engels, J. Lange, Th. Schäpers, and H. Lüth. Experimental and Theoretical<br />

Approach to <strong>Sp<strong>in</strong></strong> Splitt<strong>in</strong>g <strong>in</strong> Modulation-Doped In x Ga 1−x As/InP<br />

Quantum Wells for B → 0 . Phys. Rev. B, 55(4):R1958–R1961, Jan 1997.<br />

Hans-Andreas Engel, Emmanuel I. Rashba, and Bertrand I. Halper<strong>in</strong>. MaterialsTheory<br />

<strong>of</strong> <strong>Sp<strong>in</strong></strong> Hall Effects <strong>in</strong> Semiconductors. John Wiley & Sons, Ltd,<br />

Cambridge, MA, USA, December 2007.<br />

Sigurdur I. Erl<strong>in</strong>gsson, John Schliemann, and Daniel Loss. <strong>Sp<strong>in</strong></strong> susceptibilities,<br />

sp<strong>in</strong> densities, and their connection to sp<strong>in</strong> currents. Phys. Rev. B,<br />

71(3):035319, Jan 2005.<br />

[Fal03] V. L. Fal’ko. private communication. 2003.<br />

[FMAE + 07] J.Fabian, A.Matos-Abiague, C.Ertler, P.Stano, andI.Zutic. Semiconductor<br />

<strong>Sp<strong>in</strong></strong>tronics. Acta Physica Slovaca, 57:565, 2007.


120 Bibliography<br />

[Fro01]<br />

[GF97]<br />

[GGB + 07]<br />

[GI02]<br />

[GI04]<br />

[GKD04]<br />

[Gol05]<br />

[Gru00]<br />

[HLN80]<br />

[HNA + 99]<br />

[HPB + 97]<br />

V. A. Froltsov. Diffusion <strong>of</strong> <strong>in</strong>homogeneous sp<strong>in</strong> distribution <strong>in</strong> a magnetic<br />

field parallel to <strong>in</strong>terfaces <strong>of</strong> a III-V semiconductor quantum well. Phys. Rev.<br />

B, 64(4):045311, Jun 2001.<br />

Claudio Grimaldi and Peter Fulde. Theory <strong>of</strong> screen<strong>in</strong>g <strong>of</strong> the phononmodulated<br />

sp<strong>in</strong>-orbit <strong>in</strong>teraction <strong>in</strong> metals . Phys. Rev. B, 55(23):15523–<br />

15530, Jun 1997.<br />

S. Giglberger, L. E. Golub, V. V. Bel’kov, S. N. Danilov, D. Schuh, C. Gerl,<br />

F. Rohlf<strong>in</strong>g, J. Stahl, W. Wegscheider, D. Weiss, W. Prettl, and S. D.<br />

Ganichev. Rashba and dresselhaus sp<strong>in</strong> splitt<strong>in</strong>gs <strong>in</strong> semiconductor quantum<br />

wells measured by sp<strong>in</strong> photocurrents. Phys. Rev. B, 75(3):035327, Jan<br />

2007.<br />

M. M. Glazov and E. L. Ivchenko. Precession sp<strong>in</strong> relaxation mechanism<br />

caused by frequent electron-electron collisions. Jetp Lett., 75:403, 2002.<br />

M. M. Glazov and E. L. Ivchenko. Effect <strong>of</strong> electron-electron <strong>in</strong>teraction on<br />

sp<strong>in</strong> relaxation <strong>of</strong> charge carriers <strong>in</strong> semiconductors. J. Exp. and Theoret.<br />

Phys., 99:1279, 2004.<br />

Alexander O. Govorov, Alexander V. Kalameitsev, and John P. Dulka. <strong>Sp<strong>in</strong></strong>dependent<br />

transport <strong>of</strong> electrons <strong>in</strong> the presence <strong>of</strong> a smooth lateral potential<br />

and sp<strong>in</strong>-orbit <strong>in</strong>teraction. Phys. Rev. B, 70(24):245310, Dec 2004.<br />

L. E. Golub. Weak Antilocalization <strong>in</strong> High-Mobility Two-Dimensional Systems.<br />

Phys. Rev. B, 71(23):235310, 2005.<br />

Dirk Grundler. Large Rashba Splitt<strong>in</strong>g <strong>in</strong> InAs Quantum Wells due to Electron<br />

Wave Function Penetration <strong>in</strong>to the Barrier Layers. Phys. Rev. Lett.,<br />

84(26):6074–6077, Jun 2000.<br />

Sh<strong>in</strong>obuHikami, Anatoly I. Lark<strong>in</strong>, and Yosuke Nagaoka. <strong>Sp<strong>in</strong></strong>-Orbit Interaction<br />

and Magnetoresistance <strong>in</strong> the Two Dimensional Random System. Prog.<br />

Theor. Phys., 63(2):707–710, 1980.<br />

Can-M<strong>in</strong>g Hu, Junsaku Nitta, Tatsushi Akazaki, Hideaki Takayanagi, Jiro<br />

Osaka, P. Pfeffer, and W. Zawadzki. Zero-field sp<strong>in</strong> splitt<strong>in</strong>g <strong>in</strong> an <strong>in</strong>verted<br />

In 0.53 Ga 0.47 As/In 0.52 Al 0.48 As heterostructure: Band nonparabolicity <strong>in</strong>fluence<br />

and the subband dependence. Phys. Rev. B, 60(11):7736–7739, Sep<br />

1999.<br />

T Hassenkam, S Pedersen, K Baklanov, A Kristensen, C B Sorensen, P E<br />

L<strong>in</strong>del<strong>of</strong>, F G Pikus, and G E Pikus. <strong>Sp<strong>in</strong></strong> splitt<strong>in</strong>g and weak localization<br />

<strong>in</strong> (110) GaAs/AlxGa1-xAs quantum wells. Phys. Rev. B, 55(15):9298–9301,<br />

1997.


Bibliography 121<br />

[HSM + 06]<br />

[HSM + 07]<br />

A. W. Holleitner, V. Sih, R. C. Myers, A. C. Gossard, and D. D. Awschalom.<br />

Suppression<strong>of</strong><strong>Sp<strong>in</strong></strong>Relaxation <strong>in</strong>SubmicronInGaAsWires. Phys. Rev. Lett.,<br />

97(3):036805, 2006.<br />

A W Holleitner, V Sih, R C Myers, A C Gossard, and D D Awschalom. Dimensionallyconstra<strong>in</strong>edD’yakonov&ndash;Perel’<br />

sp<strong>in</strong>relaxation <strong>in</strong>n-InGaAs<br />

channels: transition from 2D to 1D. New Journal <strong>of</strong> Physics, 9(9):342, 2007.<br />

[HvWK + 98] J. P. Heida, B. J. van Wees, J. J. Kuipers, T. M. Klapwijk, and G. Borghs.<br />

<strong>Sp<strong>in</strong></strong>-orbit<strong>in</strong>teraction <strong>in</strong> a two-dimensional electron gas <strong>in</strong> a InAs/AlSb quantum<br />

well with gate-controlled electron density. Phys. Rev. B, 57(19):11911–<br />

11914, May 1998.<br />

[icvacFDS04] Igor Žutić, Jaroslav Fabian, and S. Das Sarma. <strong>Sp<strong>in</strong></strong>tronics: Fundamentals<br />

and applications. Rev. Mod. Phys., 76(2):323–410, Apr 2004.<br />

[Ivc73]<br />

[JLS + 08]<br />

[JRA + 95]<br />

[KCCC92]<br />

[Ket07]<br />

[KH07]<br />

[KHG + 10]<br />

[KK00]<br />

[KKA10]<br />

E. L. Ivchenko. <strong>Sp<strong>in</strong></strong> Relaxation <strong>of</strong> Free Carriers <strong>in</strong> a Noncentrosymmetric<br />

Semiconductor <strong>in</strong> a Longitud<strong>in</strong>al Magnetic Field. Sov. Phys. Solid State,<br />

15:1048, 1973.<br />

M<strong>in</strong>kyung Jung, Joon Sung Lee, Woon Song, Young Heon Kim, Sang Don<br />

Lee, Nam Kim, Jeunghee Park, Mahn-Soo Choi, Sh<strong>in</strong>go Katsumoto, Hyoyoung<br />

Lee, and J<strong>in</strong>hee Kim. Quantum Interference <strong>in</strong> Radial Heterostructure<br />

Nanowires. Nano Letters, 8(10):3189–3193, Sep 2008.<br />

Bernard Jusserand, David Richards, Guy Allan, Cather<strong>in</strong>e Priester, and<br />

Bernard Etienne. <strong>Sp<strong>in</strong></strong> orientation at semiconductor hetero<strong>in</strong>terfaces. Phys.<br />

Rev. B, 51(7):4707–4710, Feb 1995.<br />

C Kurdak, A. M. Chang, A. Ch<strong>in</strong>, and T. Y. Chang. Quantum Interference<br />

Effects and <strong>Sp<strong>in</strong></strong>-Orbit Interaction <strong>in</strong> Quasi-One-Dimensional Wires and<br />

R<strong>in</strong>gs. Phys. Rev. B, 46(11):6846–6856, Sep 1992.<br />

S. Kettemann. Dimensional Control <strong>of</strong> Antilocalization and <strong>Sp<strong>in</strong></strong> Relaxation<br />

<strong>in</strong> Quantum Wires. Phys. Rev. Lett., 98(17):176808–4, April 2007.<br />

Jacob J. Krich and Bertrand I. Halper<strong>in</strong>. Cubic Dresselhaus <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g<br />

<strong>in</strong> 2D Electron Quantum Dots. Phys. Rev. Lett., 98(22):226802, 2007.<br />

R L Kallaher, J J Heremans, N Goel, S J Chung, and M B Santos. <strong>Sp<strong>in</strong></strong> and<br />

phase coherence lengths <strong>in</strong> n-InSb quasi-one-dimensional wires. Phys. Rev.<br />

B, 81(3):35335, January 2010.<br />

A. A. Kiselev and K. W. Kim. Progressive Suppression <strong>of</strong> <strong>Sp<strong>in</strong></strong> Relaxation<br />

<strong>in</strong> Two-Dimensional Channels <strong>of</strong> F<strong>in</strong>ite Width. Phys. Rev. B, 61(19):13115–<br />

13120, May 2000.<br />

Tomoaki Kaneko, Mikito Kosh<strong>in</strong>o, and Tsuneya Ando. Symmetry crossover<br />

<strong>in</strong> quantum wires with sp<strong>in</strong>-orbit <strong>in</strong>teraction. Phys. Rev. B, 81(15):155310,<br />

Apr 2010.


122 Bibliography<br />

[KKN09]<br />

[KM02]<br />

Yoji Kunihashi, Makoto Kohda, and Junsaku Nitta. Enhancement <strong>of</strong> sp<strong>in</strong><br />

lifetime <strong>in</strong> gate-fitted InGaAs narrow wires. Phys. Rev. Lett., 102(22):226601,<br />

2009.<br />

Stefan Kettemann and Riccardo Mazzarello. Magnetolocalization <strong>in</strong> Disordered<br />

Quantum Wires. Phys. Rev. B, 65(8):085318, Feb 2002.<br />

[KMGA04] Y K Kato, R C Myers, A C Gossard, and D D Awschalom. Observation<br />

<strong>of</strong> the sp<strong>in</strong> Hall effect <strong>in</strong> semiconductors. Science (New York, N.Y.),<br />

306(5703):1910–3, December 2004.<br />

[KRW03]<br />

[KSZ + 96]<br />

J. Ka<strong>in</strong>z, U. Rössler, and R. W<strong>in</strong>kler. Anisotropic sp<strong>in</strong>-splitt<strong>in</strong>g and sp<strong>in</strong>relaxation<br />

<strong>in</strong> asymmetric z<strong>in</strong>c blende semiconductor quantum structures.<br />

Phys. Rev. B, 68(7), August 2003.<br />

W. Knap, C. Skierbiszewski, A. Zduniak, E. Litw<strong>in</strong>-Staszewska, D. Bertho,<br />

F. Kobbi, J. L. Robert, G. E. Pikus, F. G. Pikus, S. V. Iordanskii, V. Mosser,<br />

K. Zekentes, and Yu. B. Lyanda-Geller. Weak antilocalization and sp<strong>in</strong> precession<br />

<strong>in</strong> quantum wells. Phys. Rev. B, 53(7):3912–3924, Feb 1996.<br />

[KWO + 09] J D Koralek, C P Weber, J Orenste<strong>in</strong>, B A Bernevig, Shou-Cheng Zhang,<br />

S Mack, and D D Awschalom. Emergence <strong>of</strong> the persistent sp<strong>in</strong> helix <strong>in</strong><br />

semiconductor quantum wells. Nature, 458(7238):610–613, Apr 2009.<br />

[LAP]<br />

[LCCC06]<br />

[LG98]<br />

[LLK + 10]<br />

[LMFS88]<br />

[LSK + 07]<br />

[LX06]<br />

Lapack l<strong>in</strong>ear algebra package. Available at: http://www.netlib.org/lapack/.<br />

M<strong>in</strong>g-Hao Liu, Kuo-Wei Chen, Son-Hsien Chen, and Ch<strong>in</strong>g-Ray Chang. Persistent<br />

sp<strong>in</strong> helix <strong>in</strong> rashba-dresselhaus two-dimensional electron systems.<br />

Phys. Rev. B, 74(23):235322, 2006.<br />

Yuli Lyanda-Geller. Quantum Interference and Electron-Electron Interactions<br />

at Strong <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g <strong>in</strong> Disordered Systems. Phys. Rev. Lett.,<br />

80(19):4273–4276, May 1998.<br />

J. Liu, T. Last, E. Koop, S. Denega, B. van Wees, and C. van der Wal.<br />

<strong>Sp<strong>in</strong></strong>-dephas<strong>in</strong>g anisotropy for electrons <strong>in</strong> a diffusive quasi-1d gaas wire. J.<br />

Supercond. Nov. Magn., 23(1):11–15, 2010.<br />

J. Luo, H. Munekata, F. F. Fang, and P. J. Stiles. Observation <strong>of</strong> the zer<strong>of</strong>ield<br />

sp<strong>in</strong> splitt<strong>in</strong>g <strong>of</strong> the ground electron subband <strong>in</strong> gasb-<strong>in</strong>as-gasb quantum<br />

wells. Phys. Rev. B, 38(14):10142—-10145, Nov 1988.<br />

P. Lehnen, T. Schapers, N. Kaluza, N. Thillosen, and H. Hardtdegen. Enhanced<br />

sp<strong>in</strong>-orbit scatter<strong>in</strong>g length <strong>in</strong> narrow AlxGa1-xN/GaN wires. Phys.<br />

Rev. B, 76(20):205307, 2007.<br />

P. Liu and S.-J. Xiong. <strong>Sp<strong>in</strong></strong>-Hall transport <strong>in</strong> a two-dimensional electron system<br />

with magnetic impurities. The European Physical Journal B, 51(4):513–<br />

516, June 2006.


Bibliography 123<br />

[MACR06]<br />

T. Micklitz, A. Altland, T. A. Costi, and A. Rosch. Universal Dephas<strong>in</strong>g Rate<br />

due to Diluted Kondo Impurities. Phys. Rev. Lett., 96(22):226601, 2006.<br />

[Mah00] G. Mahan. Many-Particle Physics, 3rd ed. Plenum, New York, 2000.<br />

[MC00]<br />

[MCW97]<br />

A. G. Mal’shukov and K. A. Chao. Waveguide Diffusion Modes and Slowdown<br />

<strong>of</strong> D’yakonov-Perel’ <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Narrow Two-Dimensional Semiconductor<br />

Channels. Phys. Rev. B, 61(4):R2413–R2416, Jan 2000.<br />

A. G. Mal’shukov, K. A. Chao, and M. Willander. Magnetoresistance <strong>of</strong> a<br />

Weakly Disordered III-V Semiconductor Quantum Well <strong>in</strong> a Magnetic Field<br />

Parallel to Interfaces. Phys. Rev. B, 56(11):6436–6439, Sep 1997.<br />

[Mei05] F. E. Meijer. private communication. 2005.<br />

[MFA02]<br />

[MHEZ71]<br />

[MJM + 04]<br />

J S Meyer, V I Fal’ko, and B L Altshuler. No Title, volume 72 <strong>of</strong> Nato Science<br />

Series II, page 117. Kluwer Academic Publishers, Dordrecht, 2002.<br />

Müller-Hartmann, E., and J. Zittartz. Kondo Effect <strong>in</strong> Superconductors.<br />

Phys. Rev. Lett., 26(8):428–432, Feb 1971.<br />

H. Malissa, W. Jantsch, M. Mühlberger, F. Schäffler, Z. Wilamowski,<br />

M. Draxler, and P. Bauer. Anisotropy <strong>of</strong> g-factor and electron sp<strong>in</strong> resonance<br />

l<strong>in</strong>ewidth <strong>in</strong> modulation doped SiGe quantum wells. Appl. Phys. Lett.,<br />

85(10):1739–1741, 2004.<br />

[MKMM00] T. Matsuyama, R. Kürsten, C. Meißner, and U. Merkt. Rashba sp<strong>in</strong> splitt<strong>in</strong>g<br />

<strong>in</strong> <strong>in</strong>version layers on p-type bulk InAs. Phys. Rev. B, 61(23):15588–15591,<br />

Jun 2000.<br />

[MM05]<br />

[MM07]<br />

[MMF08]<br />

[MST83]<br />

[MZM + 03]<br />

Cătăl<strong>in</strong> Pa şcu Moca and D. C. Mar<strong>in</strong>escu. Longitud<strong>in</strong>al and sp<strong>in</strong>-hall conductance<br />

<strong>of</strong> a two-dimensional rashba system with arbitrary disorder. Phys.<br />

Rev. B, 72(16):165335, Oct 2005.<br />

C. P. Moca and D. C. Mar<strong>in</strong>escu. F<strong>in</strong>ite-size effects <strong>in</strong> a two-dimensional<br />

electron gas with rashba sp<strong>in</strong>-orbit <strong>in</strong>teraction. Phys. Rev. B, 75(3):035325,<br />

Jan 2007.<br />

C. Moca, D. Mar<strong>in</strong>escu, and S. Filip. <strong>Sp<strong>in</strong></strong>Hall effect <strong>in</strong> asymmetricquantum<br />

well by a random Rashba field. Phys. Rev. B, 77(19), May 2008.<br />

V. A. Marushchak, M. N. Stepanova, and A. N. Titkov. <strong>Sp<strong>in</strong></strong> Relaxation <strong>of</strong><br />

Conduction Electrons <strong>in</strong> Moderately Doped Gallium arsenide crystals. Soviet<br />

Physics - Solid State, 25(12):2035–2038, 1983.<br />

J. B. Miller, D. M. Zumbühl, C. M. Marcus, Y. B. Lyanda-Geller,<br />

D. Goldhaber-Gordon, K. Campman, and A. C. Gossard. Gate-Controlled<br />

<strong>Sp<strong>in</strong></strong>-Orbit Quantum Interference Effects <strong>in</strong> Lateral Transport. Phys. Rev.<br />

Lett., 90(7):076807, Feb 2003.


124 Bibliography<br />

[NATE97]<br />

[NATE98]<br />

J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki. Gate control <strong>of</strong> sp<strong>in</strong>orbit<br />

<strong>in</strong>teraction <strong>in</strong> an <strong>in</strong>verted In(0.53)Ga(0.47)AS/In(0.52)Al(0.48)AS heterostructure.<br />

Phys. Rev. Lett., 78(7):1335–1338, 1997.<br />

Junsaku Nitta, Tatsushi Akazaki, Hideaki Takayanagi, and Takatomo<br />

Enoki. Gate control <strong>of</strong> sp<strong>in</strong>-orbit <strong>in</strong>teraction <strong>in</strong> an InAs-<strong>in</strong>serted<br />

In 0.53 Ga 0.47 As/In 0.52 Al 0.48 As heterostructure. Physica E: Low-dimensional<br />

Systems and Nanostructures, 2(1-4):527–531, 1998.<br />

[Nit06] J. Nitta. privtate communication <strong>of</strong> unpublished results. 2006.<br />

[NSJ + 05]<br />

[NSJ + 06]<br />

[NSSM05]<br />

[OMP]<br />

[OMP + 92]<br />

[OTA + 99]<br />

[OW53]<br />

K. Nomura, Jairo S<strong>in</strong>ova, T. Jungwirth, Q. Niu, and A. H. MacDonald. Nonvanish<strong>in</strong>g<br />

sp<strong>in</strong> hall currents <strong>in</strong> disordered sp<strong>in</strong>-orbit coupl<strong>in</strong>g systems. Phys.<br />

Rev. B, 71(4):041304, Jan 2005.<br />

K. Nomura, Jairo S<strong>in</strong>ova, T. Jungwirth, Q. Niu, and A. H. MacDonald. Erratum:<br />

Nonvanish<strong>in</strong>g sp<strong>in</strong> hall currents <strong>in</strong> disordered sp<strong>in</strong>-orbit coupl<strong>in</strong>g systems<br />

[phys. rev. b 71, 041304(r) (2005)]. Phys. Rev. B, 73(19):199901, May<br />

2006.<br />

K. Nomura, Jairo S<strong>in</strong>ova, N. S<strong>in</strong>itsyn, and A. MacDonald. Dependence <strong>of</strong><br />

the <strong>in</strong>tr<strong>in</strong>sic sp<strong>in</strong>-Hall effect on sp<strong>in</strong>-orbit <strong>in</strong>teraction character. Phys. Rev.<br />

B, 72(16), October 2005.<br />

Openmp.org. Available at: http://openmp.org/wp/.<br />

H. Ohno, H. Munekata, T. Penney, S. von Molnár, and L. Chang. Magnetotransport<br />

properties <strong>of</strong> p-type (In,Mn)As diluted magnetic III-V semiconductors.<br />

Phys. Rev. Lett., 68(17):2664–2667, Apr 1992.<br />

Y. Ohno, R. Terauchi, T. Adachi, F. Matsukura, and H. Ohno. <strong>Sp<strong>in</strong></strong> Relaxation<br />

<strong>in</strong> GaAs(110) Quantum Wells. Phys. Rev. Lett., 83(20):4196–4199, Nov<br />

1999.<br />

Overhauser and Albert W. Paramagnetic Relaxation <strong>in</strong> Metals. Phys. Rev.,<br />

89(4):689–700, Feb 1953.<br />

[Pei33] R. Peierls. Zur Theorie des Diamagnetismus von Leitungselektronen.<br />

Zeitschrift für Physik, 80(11-12):763–791, November 1933.<br />

[PF06]<br />

[PHC + 09]<br />

Alexander Punnoose and Alexander M. F<strong>in</strong>kel’ste<strong>in</strong>. <strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> the<br />

Presence <strong>of</strong> Electron-Electron Interactions. Phys. Rev. Lett., 96(5):057202–4,<br />

February 2006.<br />

G. Petersen, S. Estévez Hernández, R. Calarco, N. Demar<strong>in</strong>a, and Th.<br />

Schäpers. <strong>Sp<strong>in</strong></strong>-orbitcoupl<strong>in</strong>gandphase-coherent transport<strong>in</strong>InNnanowires.<br />

Phys. Rev. B, 80(12):125321, 2009.


Bibliography 125<br />

[PMT88]<br />

[PP95]<br />

[PT84]<br />

[Ram82]<br />

[Ras60]<br />

[Ras04]<br />

[RJA + 96]<br />

[RS05]<br />

[Sak67]<br />

[SANR09]<br />

[SAYI99]<br />

[SBK + 10]<br />

G. E. Pikus, V. A. Marushchak, and A. N. Titkov. <strong>Sp<strong>in</strong></strong> splitt<strong>in</strong>g <strong>of</strong> energy<br />

bands and sp<strong>in</strong> relaxation <strong>of</strong> carriers <strong>in</strong> cubic III-V crystals (Review). Soviet<br />

Physics - Semiconductors, 22(2):115–124, 1988.<br />

F. G. Pikus and G. E. Pikus. Conduction-band sp<strong>in</strong> splitt<strong>in</strong>g and negative<br />

magnetoresistance <strong>in</strong> a 3 b 5 heterostructures. Phys. Rev. B, 51(23):16928–<br />

16935, Jun 1995.<br />

G. E. Pikus and A. N. Titkov. <strong>Sp<strong>in</strong></strong> Relaxation under Optical Orientation<br />

<strong>in</strong> Semiconductors. In F. Meier and B. P. Zakharchenya, editors, Optical<br />

Orientation, volume 8 <strong>of</strong> Modern Problems <strong>in</strong> Condensed Matter Sciences,<br />

chapter 3, pages 73–131. North-Holland, Amsterdam, 1984.<br />

J. Rammer. Quantum Transport Theory. Frontiers <strong>in</strong> Physics. Perseus Books,<br />

1982.<br />

EI Rashba. Properties <strong>of</strong> Semiconductors with an Extremum Loop .1. Cyclotron<br />

and Comb<strong>in</strong>ational Resonance <strong>in</strong> a Magnetic Field Perpendicular to<br />

the Plane <strong>of</strong> the Loop. Sov. Phys.-Solid State, 2(6):1109–1122, 1960.<br />

Emmanuel Rashba. Sum rules for sp<strong>in</strong> Hall conductivity cancellation. Phys.<br />

Rev. B, 70(20), November 2004.<br />

D. Richards, B. Jusserand, G. Allan, C. Priester, and B. Etienne. Electron<br />

sp<strong>in</strong>-flip Raman scatter<strong>in</strong>g <strong>in</strong> asymmetric quantum wells: <strong>Sp<strong>in</strong></strong> orientation.<br />

Solid-State Electronics, 40(1-8):127–131, 1996. Proceed<strong>in</strong>gs <strong>of</strong> the Seventh<br />

International Conference on Modulated Semiconductor <strong>Structures</strong>.<br />

Roberto Raimondi and Peter Schwab. <strong>Sp<strong>in</strong></strong>-hall effect <strong>in</strong> a disordered twodimensional<br />

electron system. Phys. Rev. B, 71(3):033311, Jan 2005.<br />

J. J. Sakurai. Advanced Quantum Mechanic. Addison-Wesley Pub Co Inc,<br />

1967.<br />

Matthias Scheid, Inanc Adagideli, J Nitta, and Klaus Richter. Anisotropic<br />

universal conductance fluctuations <strong>in</strong> disordered quantum wires with Rashba<br />

and Dresselhaus sp<strong>in</strong>-orbit <strong>in</strong>teraction and applied <strong>in</strong>-plane magnetic field.<br />

Semicond. Sci. and Technol., 24(6):64005, 2009.<br />

S. Sasa, K. Anjiki, T. Yamaguchi, and M. Inoue. Electron transport <strong>in</strong> a large<br />

sp<strong>in</strong>-splitt<strong>in</strong>g2DEG <strong>in</strong>InAs/AlGaSb heterostructures. Physica B: Condensed<br />

Matter, 272(1-4):149–152, 1999.<br />

K. Sato, L. Bergqvist, J. Kudrnovský, P. H. Dederichs, O. Eriksson, I. Turek,<br />

B. Sanyal, G. Bouzerar, H. Katayama-Yoshida, V. A. D<strong>in</strong>h, T. Fukushima,<br />

H. Kizaki, and R. Zeller. First-pr<strong>in</strong>ciples theory <strong>of</strong> dilute magnetic semiconductors.<br />

Rev. Mod. Phys., 82(2):1633–1690, May 2010.<br />

[Sch06] John Schliemann. <strong>Sp<strong>in</strong></strong> Hall Effect. IJMPB, 20(9), 2006.


126 Bibliography<br />

[SCN + 04]<br />

[SDGR06]<br />

[SEL03]<br />

[SGB + 09]<br />

[SGP + 06]<br />

[She04]<br />

[SKK + 08]<br />

[SR94]<br />

[SSB + 10]<br />

[SST05]<br />

[SZXN06]<br />

[TI98]<br />

J S<strong>in</strong>ova, D Culcer, Q Niu, N A S<strong>in</strong>itsyn, T Jungwirth, and A H MacDonald.<br />

Universal <strong>in</strong>tr<strong>in</strong>sic sp<strong>in</strong> Hall effect. Phys. Rev. Lett., 92(12), 2004.<br />

Peter Schwab, Michael Dzierzawa, Cosimo Gor<strong>in</strong>i, and Roberto Raimondi.<br />

<strong>Sp<strong>in</strong></strong> Relaxation <strong>in</strong> Narrow Wires <strong>of</strong> a Two-Dimensional Electron Gas. Phys.<br />

Rev. B, 74(15):155316, October 2006.<br />

John Schliemann, J. Egues, and Daniel Loss. Nonballistic <strong>Sp<strong>in</strong></strong>-Field-Effect<br />

Transistor. Phys. Rev. Lett., 90(14):146801, April 2003.<br />

Th Schäpers, V A Guzenko, A Br<strong>in</strong>ger, M Akabori, M Hagedorn, and<br />

H Hardtdegen. <strong>Sp<strong>in</strong></strong>-orbit coupl<strong>in</strong>g <strong>in</strong> Ga x In 1−x As/InP two-dimensional<br />

electron gases and quantum wire structures. Semicond. Sci. and Technol.,<br />

24(6):64001, 2009.<br />

Th. Schäpers, V. A. Guzenko, M. G. Pala, U. Zülicke, M. Governale,<br />

J. Knobbe, and H. Hardtdegen. Suppression <strong>of</strong> Weak Antilocalization <strong>in</strong><br />

Ga x In 1−x As/InP Narrow Quantum Wires. Phys. Rev. B, 74(8):081301(R),<br />

2006.<br />

Shun-Q<strong>in</strong>gShen. <strong>Sp<strong>in</strong></strong> hall effect and berry phase <strong>in</strong> two-dimensional electron<br />

gas. Phys. Rev. B, 70(8):081311, Aug 2004.<br />

Matthias Scheid, Makoto Kohda, Yoji Kunihashi, Klaus Richter, and Junsaku<br />

Nitta. All-Electrical Detection <strong>of</strong> the Relative Strength <strong>of</strong> Rashba and<br />

Dresselhaus <strong>Sp<strong>in</strong></strong>-Orbit Interaction <strong>in</strong> Quantum Wires. Phys. Rev. Lett.,<br />

101(26):266401 – 266401–4, December 2008.<br />

R.N. Silver and H. Röder. Densities <strong>of</strong> States <strong>of</strong> Mega-Dimensional Hamiltonian<br />

Matrices. IJMPC, 5(4):735–753, 1994.<br />

Gerald Schubert, Jens Schleede, Krzyszt<strong>of</strong> Byczuk, Holger Fehske, and Dieter<br />

Vollhardt. Distribution<strong>of</strong>thelocal density<strong>of</strong>states asacriterionforanderson<br />

localization: Numerically exact results for various lattices <strong>in</strong> two and three<br />

dimensions. Phys. Rev. B, 81(15):155106, Apr 2010.<br />

L. Sheng, D. N. Sheng, and C. S. T<strong>in</strong>g. <strong>Sp<strong>in</strong></strong>-hall effect <strong>in</strong> two-dimensional<br />

electron systems with rashba sp<strong>in</strong>-orbit coupl<strong>in</strong>g and disorder. Phys. Rev.<br />

Lett., 94(1):016602, Jan 2005.<br />

Junren Shi, P<strong>in</strong>g Zhang, Di Xiao, and Qian Niu. Proper def<strong>in</strong>ition <strong>of</strong> sp<strong>in</strong><br />

current <strong>in</strong> sp<strong>in</strong>-orbit coupled systems. Phys. Rev. Lett., 96(7):076604, Feb<br />

2006.<br />

HiroshiTanakaandMasakiItoh. Ab<strong>in</strong>itiocalculation<strong>of</strong>thehallconductivity:<br />

Positive hall coefficient <strong>of</strong> liquid fe. Phys. Rev. Lett., 81(17):3727–3730, Oct<br />

1998.


Bibliography 127<br />

[Tor56] H. C. Torrey. Bloch Equations with Diffusion Terms. Phys. Rev., 104(3):563–<br />

565, Nov 1956.<br />

[VMRM01] I. Vurgaftman, J. R. Meyer, and L. R. Ram-Mohan. Band parameters for<br />

III–V compoundsemiconductors and their alloys. Journal <strong>of</strong> Applied Physics,<br />

89(11):5815–5875, 2001.<br />

[Wen07]<br />

[WGZ + 06]<br />

[W<strong>in</strong>03]<br />

[W<strong>in</strong>04a]<br />

[W<strong>in</strong>04b]<br />

[WJW10]<br />

[WK]<br />

[WKSJ05]<br />

[WLZ07a]<br />

[WLZ07b]<br />

P. Wenk. <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g <strong>in</strong> Quantum Wires. Master’s thesis, Universität<br />

Hamburg (Germany), 2007.<br />

A. Wirthmann, Y.S. Gui, C. Zehnder, D. Heitmann, C.-M. Hu, and S. Kettemann.<br />

Weak Antilocalization <strong>in</strong> InAs Quantum Wires. Physica E: Lowdimensional<br />

Systems and Nanostructures, 34(1-2):493–496, 2006. Proceed<strong>in</strong>gs<br />

<strong>of</strong> the 16th International Conference on Electronic Properties <strong>of</strong> Two-<br />

Dimensional Systems (EP2DS-16).<br />

R. W<strong>in</strong>kler. <strong>Sp<strong>in</strong></strong>-Orbit Coupl<strong>in</strong>g Effects <strong>in</strong> Two-Dimensional Electron and<br />

Hole Systems, volume 191 <strong>of</strong> Spr<strong>in</strong>ger Tracts <strong>in</strong> Modern Physics. Spr<strong>in</strong>ger-<br />

Verlag, Berl<strong>in</strong>, 2003.<br />

R. W<strong>in</strong>kler. Rashba sp<strong>in</strong> splitt<strong>in</strong>g and Ehrenfest’s theorem. Physica E: Lowdimensional<br />

Systems and Nanostructures, 22(1-3):450–454, 2004.<br />

R. W<strong>in</strong>kler. <strong>Sp<strong>in</strong></strong> orientation and sp<strong>in</strong> precession <strong>in</strong> <strong>in</strong>version-asymmetric<br />

quasi-two-dimensional electron systems. Phys. Rev. B, 69(4):045317, Jan<br />

2004.<br />

M. W. Wu, J. H. Jiang, and M. Q. Weng. <strong>Sp<strong>in</strong></strong> dynamics <strong>in</strong> semiconductors.<br />

Physics Reports, 493(2-4):61–236, August 2010.<br />

P. Wenk and S. Kettemann. <strong>in</strong> preparation.<br />

J. Wunderlich, B. Kaestner, J. S<strong>in</strong>ova, and T. Jungwirth. Experimental observation<br />

<strong>of</strong> the sp<strong>in</strong>-hall effect <strong>in</strong> a two-dimensional sp<strong>in</strong>-orbit coupled semiconductor<br />

system. Phys. Rev. Lett., 94(4):047204, Feb 2005.<br />

Pei Wang, You-Quan Li, and Xuean Zhao. Nonvanish<strong>in</strong>g sp<strong>in</strong> Hall currents<br />

<strong>in</strong> the presence <strong>of</strong> magnetic impurities. Phys. Rev. B, 75(7), 2007.<br />

Pei Wang, You-Quan Li, and Xuean Zhao. Nonvanish<strong>in</strong>g sp<strong>in</strong> hall currents <strong>in</strong><br />

the presence <strong>of</strong> magnetic impurities. Phys. Rev. B, 75(7):075326, Feb 2007.<br />

[WOB + 07] C. P. Weber, J. Orenste<strong>in</strong>, B. Andrei Bernevig, Shou-Cheng Zhang, Jason<br />

Stephens, and D. D. Awschalom. Nondiffusive <strong>Sp<strong>in</strong></strong> <strong>Dynamics</strong> <strong>in</strong> a Two-<br />

Dimensional Electron Gas. Phys. Rev. Lett., 98(7):076604, 2007.<br />

[WWAF06] Alexander Weiße, Gerhard Welle<strong>in</strong>, Andreas Alvermann, and Holger Fehske.<br />

The kernel polynomial method. Rev. Mod. Phys., 78(1):275–306, Mar 2006.


128 Bibliography<br />

[Yaf63]<br />

[YDKO06]<br />

Y. Yafet. g-Factors and <strong>Sp<strong>in</strong></strong>-Lattice Relaxation <strong>of</strong> Conduction Electrons. In<br />

F. Seitz and D. Turnbull, editors, Solid State Physics, volume 14. Academic,<br />

New York, 1963.<br />

M Yamamoto, K Dittmer, B Kramer, and T Ohtsuki. <strong>Sp<strong>in</strong></strong>-polarization <strong>in</strong>duced<br />

by Rashba sp<strong>in</strong>–orbit coupl<strong>in</strong>g <strong>in</strong> three-term<strong>in</strong>al devices. Physica E:<br />

Low-dimensional Systems and Nanostructures, 32(1-2):462–465, Mai 2006.<br />

[ZBvDA04] Gergely Zaránd, László Borda, Jan von Delft, and Natan Andrei. Theory<br />

<strong>of</strong> Inelastic Scatter<strong>in</strong>g from Magnetic Impurities. Phys. Rev. Lett.,<br />

93(10):107204, Sep 2004.<br />

[ZFD04]<br />

Igor Zutic, Jaroslav Fabian, and Sarma S. Das. <strong>Sp<strong>in</strong></strong>tronics: Fundamentals<br />

and Applications. Rev. Mod. Phys., 76(2):323, 2004.


Appendix A<br />

SOC Strength <strong>in</strong> the Experiment<br />

129


130 Appendix A: SOC Strength <strong>in</strong> the Experiment<br />

System <strong>Sp<strong>in</strong></strong> splitt<strong>in</strong>g at E F α 2 Reference<br />

(meV) (meV Å)<br />

AlSb/InAs/AlSb 3.2 - 4.5 60 [HvWK + 98]<br />

AlSb/InAs/AlSb 0 0 [BEW + 99]<br />

AlSb/InAs/AlSb 0 0 [SAYI99]<br />

AlGaAs/GaAs/AlGaAs - 6.9±0.4 [JRA + 95]<br />

2DEG GaAs/AlGaAs - 5±1 [MZM + 03]<br />

AlGaSb/InAs/AlSb 5.6 - 13 120 - 280 [SAYI99]<br />

InAlAs/InGaAs/InAlAs 1.5 40 [DMD + 89]<br />

InAlAs/InGaAs/InAlAs 4.9 - 5.9 63 - 93 [NATE97]<br />

InAlAs/InGaAs/InAlAs - 50 - 100 [HNA + 99]<br />

InGaAs/InAs/InGaAs 5.1 - 6.8 60 - 110 [NATE98]<br />

InGaAs/InAs/InGaAs 9 - 15 200 - 400 [Gru00]<br />

InGaAs/InP/InGaAs - 63 - 153 [ELSL97]<br />

GaSb/InAs/GaSb 3.7 90 [LMFS88]<br />

Si/SiGe QW - 0.03 - 0.12 [MJM + 04]<br />

SiO2/InAs/ 5.5 - 23 100 - 300 [MKMM00]<br />

n-In 0.2 Ga 0.8 As/GaAs QW 50-70 [HSM + 06]<br />

Table A.1: Values <strong>of</strong> Rashba parameter α 2 measured <strong>in</strong> experiments [QW=quantum wire].<br />

(List extracted from[FMAE + 07].)<br />

Parameter AlAs AlP AlSb GaAs GaP GaSb InAs InP InSb<br />

m e 0.15 0.22 0.14 0.067 0.13 0.039 0.026 0.0795 0.0135<br />

Table A.2: Experimental values <strong>of</strong> m e [VMRM01]<br />

System γ D (eV Å 3 ) Reference<br />

GaAs 24.5 [MST83]<br />

GaAs 17.4 - 26 [PMT88]<br />

GaAs 26.1±0.9 [DPWS92]<br />

GaAs 16.5±3 [JRA + 95]<br />

GaAs 11 [RJA + 96]<br />

GaAs 9 [KH07]<br />

GaAs 28±4 [MZM + 03]<br />

InGaAs 24 [KSZ + 96]<br />

Table A.3: Values <strong>of</strong> Dresselhaus parameter γ D measured <strong>in</strong> experiments. (List extracted<br />

from[FMAE + 07].)


Appendix A: SOC Strength <strong>in</strong> the Experiment 131<br />

material QW width spacer 1 spacer 2 mobility density α 2 /α 1 α ′ 2 /α′ 1<br />

Å Å Å (cm 2 /Vs) cm −2 SGE CPGE<br />

InAs/AlGaSb 150 3.0×10 5 8×10 11 2.1 2.3<br />

InAs/AlGaSb 150 2.0×10 5 1.4×10 12 1.8<br />

InAs/InAlAs 60 75 1.1×10 5 7.7×10 11 1.6<br />

GaAs/AlGaAs ∞ 700 3.5×10 6 1.1×10 11 7.6 7.6<br />

GaAs/AlGaAs 82 50 50 2.6×10 6 9.3×10 11 -4.5 -4.2<br />

GaAs/AlGaAs 150 600 300 1.0×10 5 6.6×10 11 -3.8<br />

GaAs/AlGaAs 150 400 500 2.6×10 5 5.3×10 11 -2.4<br />

GaAs/AlGaAs 300 700 3.2×10 6 1.3×10 11 2.8<br />

GaAs/AlGaAs 300 700 1000 3.4×10 6 1.8×10 11 1.5<br />

Table A.4: Measured α 2 /α 1 ratios with sp<strong>in</strong>-galvanic effect (SGE) and Circular photogalvaniceffect(CPGE)withtheaccord<strong>in</strong>gparametersat4.2K<br />

[QW=quantumwire].[GGB + 07]


Appendix B<br />

L<strong>in</strong>ear Response<br />

Apply<strong>in</strong>g l<strong>in</strong>ear response to first order with a perturbation W(t) we get for the<br />

time dependent expectation value <strong>of</strong> an operator Â<br />

[ ] 〈Â〉t = Tr ρ 0 Â −i<br />

∫ ∞<br />

−∞<br />

dt ′ θ(t−t ′ )<br />

〈 〉<br />

[ÂD(t),ŴD(t ′ )] , (B.1)<br />

0<br />

where the <strong>in</strong>dex D <strong>in</strong>dicates the Dirac picture. The source <strong>of</strong> the perturbation is an electric<br />

field E(t) = E 0 e i(ω+iη)t with η be<strong>in</strong>g an <strong>in</strong>f<strong>in</strong>itesimally small positive value. The perturbation<br />

is then given by W(t) = −ˆP · E(t), with the dipole operator ˆP = ∑ i q iˆr i and the<br />

charge q i at the positions r i .<br />

The response to the applied electric field is a current J µ = σ µν E ν . For the next steps we<br />

keep the def<strong>in</strong>ition <strong>of</strong> the current general. Later on we can relate it to the sp<strong>in</strong> Hall current<br />

J z x = −σz xy E y, with the sp<strong>in</strong> Hall conductivity σ z xy ≡ σ SH.<br />

〈J µ 〉 t<br />

= i ∑ ν<br />

∫ ∞<br />

−∞<br />

dtθ(t−t ′ ) 〈 [J µ D (0),P νD(t ′ −t)] 〉 e i(ω+iη)(t′ −t) E 0ν e −i(ω+iη)t<br />

} {{ }<br />

E ν(t)<br />

(B.2)<br />

the sp<strong>in</strong> Hall conductivity (<strong>in</strong> the follow<strong>in</strong>g the subscript D will be left out)<br />

σ µν (ω) = i<br />

∫ ∞<br />

−∞<br />

dtθ(−t)〈[J µ (0),P νD (t)]〉e −i(ω+iη)t<br />

(B.3)<br />

In the next step we want to replace ˆP by its time derivative. This can be accomplished by<br />

apply<strong>in</strong>g the Kubo identity<br />

[A(t),ρ 0 ] = −iρ 0<br />

∫ β<br />

0<br />

dx ˙ A(t−ix)<br />

(B.4)<br />

132


Appendix B: L<strong>in</strong>ear Response 133<br />

<strong>in</strong> comb<strong>in</strong>ation with 〈[J,P]〉 = Tr([P,ρ 0 ]J). We get<br />

σ µν (ω) = V<br />

∫ β<br />

0<br />

dx<br />

∫ ∞<br />

0<br />

dtTr(ρ 0 J ν (0)J µ (t+ix))e i(ω+iη)t ,<br />

us<strong>in</strong>g (1/V)Ṗ = J, with the system volume V, β = 1/k BT and the density matrix<br />

ρ 0 = e−βH 0<br />

Tr(e −βH 0 )<br />

.<br />

(B.5)<br />

(B.6)<br />

Extract<strong>in</strong>g the time dependency <strong>of</strong> the current operator J µ D (t+ix) = ei(t+ix)H J µ e −i(t+ix)H<br />

and us<strong>in</strong>g the eigenvector basis {|i〉} we can write the conductivity tensor <strong>in</strong> the follow<strong>in</strong>g<br />

way<br />

σ µν (ω) = V<br />

∫ β<br />

0<br />

dx<br />

∫ ∞<br />

0<br />

dt ∑ 〈<br />

∣<br />

∣<br />

∣∣m<br />

〈m|J ν (0)|n〉 n∣e i(t+ix)En J µ e<br />

〉e −i(t+ix)Em i(ω+iη)t·<br />

m,n<br />

(B.7)<br />

·Tr(ρ 0 a † m a na † p a q).<br />

(B.8)<br />

We used H = ∑ m E ma † ma m . For the next step we need the identity<br />

Tr(ρ 0 a † ma n a † pa q ) = δ mq δ np f(E m )(1−f(E n )),<br />

(B.9)<br />

where f(E) is the Fermi distribution function.<br />

Pro<strong>of</strong>. The factor δ mq δ np is due to momentum conservation. The second part is yield by<br />

commutator relation<br />

Tr(ρ 0 a † ma n a † pa q ) = Tr(ρ 0 a † ma n (δ nm −a m a † n))<br />

(B.10)<br />

= Tr(ρ 0 (a † m a nδ nm −a † m a na m a † n )) (B.11)<br />

= Tr(ρ 0 (a † ma n δ nm −a † ma m a n a † n)) (B.12)<br />

= Tr(ρ 0 (a † m a nδ nm −a † m a m(1−a † n a n)) (B.13)<br />

= Tr(ρ 0 (a † ma n δ nm −n m (1−n n )). (B.14)<br />

Apply<strong>in</strong>g Eq.(B.9) to Eq.(B.7) and evaluat<strong>in</strong>g the <strong>in</strong>tegral over x we get<br />

σ µν (ω) = V<br />

∫ ∞<br />

0<br />

dt ∑ f(E m )(1−f(E n )) 1−e−β(En−Em)<br />

〈m|J ν |n〉〈n|J µ |m〉<br />

(E<br />

m,n<br />

n −E m )<br />

} {{ }<br />

f(Em)−f(En)<br />

En−Em<br />

·e it(ω+iη+En−Em) . (B.15)


134 Appendix B: L<strong>in</strong>ear Response<br />

F<strong>in</strong>ally we perform the t-<strong>in</strong>tegration and rewrite the formula for currents j:<br />

= i V<br />

∑<br />

m,n<br />

(f(E m )−f(E n ))〈m|j ν |n〉〈n|j µ |m〉<br />

E n −E m E n −E m +ω +iη .<br />

(B.16)<br />

B.1 Kubo Formula for Weak Disorder<br />

In the follow<strong>in</strong>g we are <strong>in</strong>terested <strong>in</strong> the real part <strong>of</strong> Eq.(B.16) for the longitud<strong>in</strong>al<br />

conductivity at zero frequency. We rewrite the real part to a form which is suitable for the<br />

diagrammatic perturbation. The first step is<br />

2π<br />

σ x,x (ω = 0) = lim<br />

ω→0 V<br />

= 2π<br />

V<br />

e 2<br />

m 2 e m,n<br />

e 2 ∫ ∞<br />

m 2 e 0<br />

∑ f(E m )−f(E n )<br />

〈m|p x |n〉〈n|p x |m〉δ(E m −E n +ω) (B.17)<br />

E n −E m<br />

(<br />

dE − ∂f(E) ) ∑ f(E m )−f(E n )<br />

·<br />

∂E E<br />

m,n n −E m<br />

·〈m|p x |n〉〈n|p x |m〉δ(E −E m )δ(E −E n ).<br />

(B.18)<br />

Includ<strong>in</strong>g impurities we average over all configurations, writ<strong>in</strong>g the sum as a trace<br />

σ impx,x (0) = 2π<br />

V<br />

e 2 ∫ ∞<br />

m 2 e 0<br />

dE<br />

(<br />

− ∂f(E) )<br />

〈Tr[δ(E −H 0 )p x δ(E −H 0 )p x ]〉<br />

∂E<br />

imp<br />

.<br />

Now we chose the basis <strong>in</strong> momentum space {|k〉}. After apply<strong>in</strong>g orthogonality relation<br />

we get<br />

σ impx,x (0) = 1 e 2 ∫ ∞<br />

2πV m 2 dE ∑ 〈<br />

〉<br />

k x k x<br />

′ k<br />

1<br />

∣<br />

e 0 E −H 0 −iη − 1<br />

E −H 0 +iη∣ k′ ·<br />

k,k<br />

〈 ∣ ′ 〉<br />

∣∣∣<br />

· k ′ 1<br />

E −H 0 −iη − 1<br />

E −H 0 +iη∣ k<br />

= 1<br />

2πV<br />

e 2 ∫ ∞<br />

m 2 e 0<br />

dE ∑ k,k ′ k x k ′ x·<br />

(B.19)<br />

·〈2G R E(k,k ′ )G A E(k ′ ,k)−G R E(k,k ′ )G R E(k ′ k)−G A E(k,k ′ )G A E(k ′ ,k) 〉 imp<br />

(B.20)<br />

with the def<strong>in</strong>ition<br />

G R/A<br />

E (k′ ,k) =<br />

〈 ∣ ∣∣∣<br />

k ′ 1<br />

E −H 0 ∓iη<br />

〉<br />

∣ k . (B.21)<br />

This can be simplified by notic<strong>in</strong>g that the averages 〈 G R G R〉 and 〈 G A G A〉 are small<br />

imp imp<br />

compared with the other terms


Appendix B: L<strong>in</strong>ear Response 135<br />

Pro<strong>of</strong>. We start by writ<strong>in</strong>g down the Green’s function with f<strong>in</strong>ite complex self-energy Σ<br />

1<br />

G R/A (E) =<br />

E −(H 0 −µ+Σ R/A )<br />

= x∓iI<br />

x 2 +I 2<br />

(B.22)<br />

(B.23)<br />

with x = E − (H 0 − µ + RΣ) and ∓I = IΣ R/A , where µ is the chemical potential. To<br />

calculate G R (E)G R (E), and accord<strong>in</strong>gly for the pair <strong>of</strong> advanced Green’s functions, we<br />

write the expression <strong>in</strong> terms <strong>of</strong> x, I and the spectral function<br />

This yields<br />

S = 1 π<br />

I<br />

x 2 +I 2.<br />

(B.24)<br />

G R G R = (x−iI)2<br />

(x 2 +I 2 ) 2 (B.25)<br />

=<br />

1<br />

(x 2 +I 2 ) − 2iIx<br />

(x 2 +I 2 ) 2 − 2I 2<br />

(x 2 +I 2 ) 2 (B.26)<br />

= π S I −2π2 i S2<br />

I x−2π2 S 2 .<br />

(B.27)<br />

Assum<strong>in</strong>g the weak disorder limit, i.e. the impurity density n imp → 0, it follows that<br />

due to τ ∝ n −1<br />

imp, with τ −1 ≡ −2IΣ R , the spectral function becomes a delta distribution.<br />

Us<strong>in</strong>g[Mah00]<br />

we end up with<br />

In the last step we applied aga<strong>in</strong> Eq.(B.28).<br />

( ) S<br />

lim<br />

I→0 I −2πS2 = 0 (B.28)<br />

G R G R = lim<br />

I→0<br />

−2π 2 i S2<br />

I x<br />

= −iπ 1<br />

I 2xδ(x)<br />

(B.29)<br />

(B.30)<br />

= 0. (B.31)<br />

In contrast to this result we get for the retarded-advanced-pair <strong>of</strong> Green’s functions <strong>in</strong> the<br />

weak disorder limit<br />

which is divergent and therefore significant.<br />

G R G A = lim<br />

I→0<br />

1<br />

x 2 +I 2<br />

= lim<br />

I→0<br />

π S I ,<br />

(B.32)<br />

(B.33)


136 Appendix B: L<strong>in</strong>ear Response<br />

Furthermore we assume low temperature which is why the derivative <strong>of</strong> the Fermi<br />

function fixes the energy <strong>of</strong> the Green’s functions to the Fermi energy,<br />

− ∂f T=0<br />

∂E = δ(E −E F).<br />

(B.34)<br />

We end up with<br />

σ impx,x (0) = 1<br />

πV<br />

e 2<br />

m 2 e<br />

∑<br />

k x k ′ 〈<br />

x G<br />

R<br />

E (k,k ′ )G A E (k′ ,k) 〉 . (B.35)<br />

imp<br />

k,k ′


Appendix C<br />

Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation<br />

C.1 Sum Formula for the Cooperon<br />

Writ<strong>in</strong>g the Cooperon (Eq.3.43)<br />

Ĉ(Q) =<br />

1<br />

D e (Q+2eA+2eA S ) 2 +H γD<br />

.<br />

(C.1)<br />

<strong>in</strong> s<strong>in</strong>glet, |S = 0;m = 0〉 = (|↑↓〉 − |↓↑〉)/ √ 2 ≡ |⇄〉 and triplet |S = 1;m = 0〉 = (|↑↓〉 +<br />

|↓↑〉)/ √ 2 ≡ |⇉〉,|S = 1;m = 1〉 ≡ |⇈〉,|S = 1;m = −1〉 ≡ |〉 representation, without<br />

magnetic field the s<strong>in</strong>glet sector is decoupled from the triplet one. To sum over C αββα <strong>in</strong><br />

the case <strong>of</strong> a f<strong>in</strong>ite magnetic field and hav<strong>in</strong>g calculated the eigenvectors |i〉 and eigenvalue<br />

λ i for the Cooperon Hamiltonian, we can use the follow<strong>in</strong>g simplification:<br />

∑<br />

C αββα = ∑ ∑<br />

〈αβ|mS〉 〈 m ′ S ′ |βα 〉〈 mS|C|m ′ S ′〉 .<br />

αβ αβ mS<br />

m ′ S ′<br />

Only several <strong>of</strong> the prefactors 〈αβ|mS〉〈m ′ S ′ |βα〉 are non-zero:<br />

(C.2)<br />

For αβ =⇈<br />

〈⇈ |⇈〉〈⇈ |⇈〉 = 1, (C.3)<br />

For αβ =<br />

〈 |〉〈 |〉 = 1, (C.4)<br />

137


138 Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation<br />

For αβ =↑↓<br />

〈↑↓ |⇄〉〈⇄ |↓↑〉 = − 1 2 ,<br />

〈↑↓ |⇄〉〈⇉ |↓↑〉 = + 1 2 ,<br />

〈↑↓ |⇉〉〈⇄ |↓↑〉 = − 1 2 ,<br />

〈↑↓ |⇉〉〈⇉ |↓↑〉 = + 1 2 ,<br />

(C.5)<br />

(C.6)<br />

(C.7)<br />

(C.8)<br />

For αβ =↓↑<br />

〈↓↑ |⇄〉〈⇄ |↑↓〉 = − 1 2 ,<br />

〈↓↑ |⇄〉〈⇉ |↑↓〉 = − 1 2 ,<br />

〈↓↑ |⇉〉〈⇄ |↑↓〉 = + 1 2 ,<br />

〈↓↑ |⇉〉〈⇉ |↑↓〉 = + 1 2 .<br />

(C.9)<br />

(C.10)<br />

(C.11)<br />

(C.12)<br />

Insert<strong>in</strong>g the eigenvectors,<br />

∑<br />

C αββα = ∑<br />

αβ αβ<br />

we end up with<br />

∑<br />

〈αβ|mS〉 〈 m ′ S ′ |βα 〉 〈mS|i〉 〈 i|C|m ′ S ′〉 ,<br />

∑<br />

mS i<br />

m ′ S ′<br />

(C.13)<br />

= ∑ i<br />

(〈⇈ |i〉〈i|⇈〉+〈⇉ |i〉〈i|⇉〉+〈 |i〉〈i|〉−〈⇄ |i〉〈i|⇄〉)λ −1<br />

i<br />

. (C.14)<br />

Writ<strong>in</strong>g<br />

∑<br />

C αββα = C ↑↑,↑↑ +C ↑↓,↓↑ +C ↓↑,↑↓ +C ↓↓,↓↓<br />

αβ<br />

⎡⎛<br />

⎞ ⎤<br />

1 0 0 0<br />

0 0 1 0<br />

= Tr<br />

C<br />

⎢⎜<br />

⎣⎝<br />

0 1 0 0 ⎟ ⎥<br />

⎠ ⎦<br />

0 0 0 1<br />

≡ Tr[ΛC]<br />

(C.15)<br />

(C.16)<br />

(C.17)


Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation 139<br />

<strong>in</strong> s<strong>in</strong>glet-triplet representation by us<strong>in</strong>g the transformation U ST ,<br />

⎛ ⎞ ⎛ ⎞ ⎛ ⎞<br />

|⇈〉 |⇄〉 0 √2 1<br />

−√ 1 2<br />

0<br />

|↑↓〉<br />

|⇈〉<br />

1 0 0 0<br />

U ST ≡<br />

⊗<br />

=<br />

, (C.18)<br />

⎜<br />

⎝ |↓↑〉 ⎟ ⎜<br />

⎠ ⎝ |⇉〉 ⎟ ⎜ 1<br />

⎠ ⎝ 0 √2 √2 1<br />

0 ⎟<br />

⎠<br />

|〉 |〉 0 0 0 1<br />

it can be seen that the s<strong>in</strong>glet term has positive contribution to the conductivity <strong>in</strong> contrast<br />

to the triplet terms which have a different sign: Transform<strong>in</strong>g Λ us<strong>in</strong>g U ST we get<br />

⎛ ⎞<br />

−1 0 0 0<br />

U ST ΛU −1<br />

ST = 0 1 0 0<br />

, (C.19)<br />

⎜<br />

⎝ 0 0 1 0 ⎟<br />

⎠<br />

0 0 0 1<br />

where we can immediately extract the signs.<br />

C.2 <strong>Sp<strong>in</strong></strong>-Conserv<strong>in</strong>g Boundary<br />

In the follow<strong>in</strong>g we set γ g = 1. In order to generate a f<strong>in</strong>ite system, we need to<br />

specify the boundary conditions. These can be different for the sp<strong>in</strong> and charge current.<br />

Here we derive the sp<strong>in</strong>-conserv<strong>in</strong>g boundary conditions. Let us first recall the diffusion<br />

current density j at position r as derived from a classical picture <strong>in</strong> Sec.2.3.4, which is given<br />

by<br />

j si (r,t) = 〈vs k i (r,t)〉−D e ∇s i (r,t),<br />

(C.20)<br />

where s k i is the part <strong>of</strong> the sp<strong>in</strong>-density which evolved from the sp<strong>in</strong>-density at r − ∆x<br />

mov<strong>in</strong>g with velocity v and momentum k. Us<strong>in</strong>g the Bloch equation<br />

∂ŝ<br />

∂t = ŝ×B SO(k)− 1ˆτ s<br />

ŝ,<br />

we rewrite the first term <strong>in</strong> Eq.(C.20) yield<strong>in</strong>g the total sp<strong>in</strong>-diffusion current as<br />

(C.21)<br />

j si = −τ〈v F [B SO (k)×S] i<br />

〉−D e ∇s i .<br />

(C.22)<br />

In Sec.3.4.1 we consider specular scatter<strong>in</strong>g from the boundary with the condition that the<br />

sp<strong>in</strong> is conserved, so that the sp<strong>in</strong> current density normal to the boundary must vanish<br />

n·j si | ±W/2 = 0,<br />

(C.23)


140 Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation<br />

where n is the vector normal to the boundary. Not<strong>in</strong>g the relation between the sp<strong>in</strong>diffusion<br />

equation <strong>in</strong> the s i representation and the triplet components <strong>of</strong> the Cooperon<br />

density ˜s i ({|⇈〉,|⇉〉,|〉}), Eq.(3.60),<br />

U CD (ǫ ijk B SO,j ) i=1..3,k=1..3 U † CD = −i(〈˜s i |B SO ·S|˜s k 〉) i=1..3,k=1..3 ,<br />

(C.24)<br />

where the matrix U CD is given by Eq.(3.61), we can thereby transform the boundary condition<br />

for the sp<strong>in</strong>-diffusion current, Eq.(C.23), to the triplet components <strong>of</strong> the Cooperon<br />

density ˜s i ,<br />

0 = n·j˜si | y=±W/2 . (C.25)<br />

Requir<strong>in</strong>g also that the charge density is vanish<strong>in</strong>g normal to the transverse boundaries,<br />

which transforms <strong>in</strong>to the condition −i∂ n˜ρ| Surface = 0 for the s<strong>in</strong>glet component <strong>of</strong> the<br />

Cooperon density ˜ρ, we f<strong>in</strong>ally get the boundary conditions for the Cooperon without external<br />

magnetic field, Eq.(3.70),<br />

(− τ D e<br />

n·〈v F [B SO (k)·S]〉−i∂ n<br />

)<br />

C| Surface = 0. (C.26)<br />

The last expression can be rewritten us<strong>in</strong>g the effective vector potential A S , Eq.(3.43),<br />

(n·2eA S −i∂ n )C| Surface = 0.<br />

(C.27)<br />

In the case <strong>of</strong> Rashba and l<strong>in</strong>ear and cubic Dresselhaus SO coupl<strong>in</strong>g <strong>in</strong> (001) systems, we<br />

get<br />

D e<br />

τ 2eA S = −〈v F (B SO (k)·S)〉<br />

= v 2 Fm e<br />

⎛<br />

⎝ −(α 1 − γ D(m ev F ) 2<br />

4<br />

) −α 2<br />

α 2 α 1 − γ D(m ev F ) 2<br />

4<br />

C.3 Relaxation Tensor<br />

⎞<br />

⎠.S.<br />

(C.28)<br />

To connect the effective vector potential A S with the sp<strong>in</strong> relaxation tensor, we<br />

notice that ˆτ can be rewritten <strong>in</strong> the follow<strong>in</strong>g way:<br />

1<br />

ˆτ s<br />

= τ(〈B SO (k) 2 〉δ ij −〈B SO (k) i B SO (k) j 〉) i=1..3,j=1..3<br />

(C.29)


Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation 141<br />

us<strong>in</strong>g U CD , Eq.(3.60),<br />

= τU † CD{〈B SO (k) x 〉S 2 x +〈B SO(k) y 〉S 2 y<br />

+〈B SO (k) x B SO (k) y 〉(S x S y +S y S x )}U CD<br />

(C.30)<br />

= τU † CD〈(B SO (k).S) 2 〉U CD (C.31)<br />

= τ<br />

vF<br />

2 U † CD〈(v F B SO (k).S) 2 〉U CD .<br />

(C.32)<br />

Because<br />

τ<br />

vF<br />

2 〈(v F [B SO (k).S]) 2 〉 = 2τ<br />

vF<br />

2 〈(v F [B SO (k).S])〉 2 (C.33)<br />

is true for l<strong>in</strong>ear Rashba and l<strong>in</strong>ear Dresselhaus SO coupl<strong>in</strong>g but, <strong>in</strong> general, false if cubic<strong>in</strong>-k<br />

terms are <strong>in</strong>cluded <strong>in</strong> the SO field, we have to write<br />

τ〈(B SO (k).S) 2 〉 = 2τ<br />

vF<br />

2 〈(v F B SO (k).S)〉 2 +ct<br />

(C.34)<br />

so that we conclude<br />

1<br />

ˆτ s<br />

= U † CD(D e (2eA S ) 2 +ct)U CD<br />

(C.35)<br />

with the separated cubic part ct = D e m 2 eE 2 F γ2 D (S2 x+S 2 y). This reflects noth<strong>in</strong>g but the fact<br />

that the effective SO Zeeman term <strong>in</strong> Eq.(3.32) can only be rewritten as a vector potential<br />

A S when the SO coupl<strong>in</strong>g is l<strong>in</strong>ear <strong>in</strong> momentum.<br />

As an example we assume a very general Cooperon that means the growth direction <strong>of</strong> the<br />

material and the SO coupl<strong>in</strong>g should be very general. We set<br />

(2eA S ) 2 +H γ = (α 11 S x +α 21 S y +α 31 S z ) 2 +(α 12 S x +α 22 S y +α 32 S z ) 2 +α 4 S 2 z<br />

(C.36)<br />

After the transformation we get<br />

⎛<br />

⎞<br />

α 2 21<br />

1<br />

ˆτ = 22 +α2 31 +α2 32 +α 4 −α 11 α 21 −α 12 α 22 −α 11 α 31 −α 12 α 32<br />

⎜<br />

⎝ −α 11 α 21 −α 12 α 22 α 2 11 +α2 12 +α2 31 +α2 32 +α 4 −α 21 α 31 −α 22 α 32<br />

⎟<br />

⎠ .<br />

−α 11 α 31 −α 12 α 32 −α 21 α 31 −α 22 α 32 α 2 11 +α2 12 +α2 21 +α2 22<br />

(C.37)<br />

C.4 Weak Localization Correction <strong>in</strong> 2D<br />

In contrast to the case where we have a wire with a f<strong>in</strong>ite width, we can calculate<br />

the weak localization correction to the conductivity analytically <strong>in</strong> the 2D case. The cut<strong>of</strong>fs


142 Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation<br />

duetodephas<strong>in</strong>gc 1 = 1/D e Q 2 SO τ ϕ andelasticscatter<strong>in</strong>gc 2 = 1/D e Q 2 SO τ determ<strong>in</strong>ewhether<br />

we have a positive or negative correction. Integrat<strong>in</strong>g over all possible wave vectors K =<br />

k/Q SO <strong>in</strong> the case without boundaries yields<br />

∆σ = − 2e2<br />

2π<br />

∫ √<br />

1 c2<br />

(<br />

(2π) 2 dK(2πK)<br />

0<br />

1<br />

+<br />

E T+ (Q SO K)/Q 2 +<br />

SO<br />

+c 1<br />

1<br />

1<br />

−<br />

E S (Q SO K)/Q 2 +<br />

SO<br />

+c 1 E T0 (Q SO K)/Q 2 SO<br />

+c 1<br />

)<br />

1<br />

E T− (Q SO K)/Q 2 SO<br />

+c 1<br />

(C.38)<br />

= − 2e2<br />

2π<br />

⎧<br />

⎪⎨<br />

arctan<br />

+<br />

⎪⎩<br />

⎧<br />

⎪⎨<br />

arctan<br />

+<br />

⎪⎩<br />

(− 1 2 ln (<br />

1+ c 2<br />

(<br />

(<br />

c 1<br />

)<br />

5 √ 1<br />

4 7<br />

16 +c 1<br />

3 √ 1<br />

4 7<br />

16 +c 1<br />

)<br />

√<br />

)<br />

√<br />

+ 1 2 ln (<br />

1+ c 2<br />

−arctan<br />

7<br />

16 +c 1<br />

+arctan<br />

7<br />

16 +c 1<br />

1+c 1<br />

)<br />

( √<br />

1<br />

16 +c 2+1<br />

√<br />

7<br />

16 +c 1<br />

( √ 1<br />

16 +c 2−1<br />

√<br />

7<br />

16 +c 1<br />

)<br />

)<br />

⎛<br />

− 1 2 ln ⎝<br />

⎛<br />

− 1 2 ln ⎝<br />

2+c 1<br />

√<br />

3<br />

2 +c 1 +c 2 +2<br />

1+c 1<br />

√<br />

3<br />

2 +c 1 +c 2 −2<br />

1<br />

16 +c 2<br />

1<br />

16 +c 2<br />

⎫<br />

⎞<br />

⎪⎬<br />

⎠<br />

⎪⎭<br />

⎫<br />

⎞<br />

⎪⎬<br />

⎠<br />

). (C.39)<br />

⎪⎭<br />

As an example, we choose parameters which have been used <strong>in</strong> the case <strong>of</strong> boundaries,<br />

1/D e Q 2 τ SO ϕ = 0.08,1/D e Q 2 τ = 4: SO ∆σ/(2e2 /2π) = −0.29. The exact calculation <strong>of</strong> wide<br />

wires (Q SO W > 1) approaches this limit as can be seen <strong>in</strong> Fig.3.11. The weak localization<br />

correction <strong>in</strong> 2D as function <strong>of</strong> these<br />

C.5 Exact Diagonalization<br />

We write the <strong>in</strong>verse Cooperon propagator, the Hamiltonian ˜H c , <strong>in</strong> the representation<br />

<strong>of</strong> the longitud<strong>in</strong>al momentum Q x , the quantized transverse momentum with quantum<br />

numbern ∈ N, and <strong>in</strong> the representation <strong>of</strong> s<strong>in</strong>glet and triplet states with quantum numbers<br />

S,m, where we note that ˜H c is diagonal <strong>in</strong> Q x ,<br />

〈Q x ,n,S,m | ˜H c | Q x ,n ′ ,S ′ ,m ′ 〉. (C.40)


Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation 143<br />

0.00<br />

0.05<br />

c 1<br />

c 2<br />

0.10 0<br />

1<br />

2<br />

3<br />

0.4<br />

0.2<br />

4 0.2<br />

0.0<br />

“ ∆σ ”<br />

2e 2<br />

2π<br />

Figure C.1: Weak localization correction <strong>in</strong> 2D <strong>in</strong> units <strong>of</strong> (2e 2 /2π). The parameters are<br />

c 1 = 1/D e Q 2 SOτ ϕ and c 2 = 1/D e Q 2 SOτ. Thick l<strong>in</strong>e <strong>in</strong>dicates ∆σ = 0.<br />

The sp<strong>in</strong> subspace is thus represented by 4×4 matrices, which we order start<strong>in</strong>g with the<br />

s<strong>in</strong>glet S = 0 and then S = 1,m = 1, m = 0, and m = −1. Thus, we get<br />

⎛<br />

⎞<br />

A n 0 0 0<br />

〈Q x ,n | ˜H c | Q x ,n〉 = Q 2 0 B n iF n D n<br />

SO<br />

. (C.41)<br />

⎜<br />

⎝ 0 −iF n C n iF n<br />

⎟<br />

⎠<br />

0 D n −iF n B n<br />

The calculation <strong>of</strong> the matrix elements yields (we set P = Q SO W/π)<br />

and for n > 0:<br />

A 0 = Kx 2 ,<br />

(C.42)<br />

B 0 = 3 4 +K2 x − 1 s<strong>in</strong>(Pπ)<br />

,<br />

4 Pπ<br />

(C.43)<br />

C 0 = 1 2 +K2 x + 1 s<strong>in</strong>(Pπ)<br />

,<br />

2 Pπ<br />

(C.44)<br />

D 0 = − 1 − 1 s<strong>in</strong>(Pπ)<br />

,<br />

4 4 Pπ<br />

(C.45)<br />

F 0 =<br />

√ s<strong>in</strong>( Pπ<br />

2Kx , (C.46)<br />

2 )<br />

Pπ<br />

2<br />

( n<br />

) 2,<br />

A n = Kx 2 +<br />

(C.47)<br />

P<br />

B n = 3 4 +K2 x + ( n<br />

P<br />

) 2<br />

+<br />

2P 2 −n 2<br />

4(n+P)(n−P)<br />

s<strong>in</strong>(Pπ)<br />

,<br />

Pπ<br />

(C.48)


144 Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation<br />

C n = 1 ( n<br />

) 2<br />

2 +K2 x + 2P 2 −n 2 s<strong>in</strong>(Pπ)<br />

− ,<br />

P 2(n+P)(n−P) Pπ<br />

(C.49)<br />

D n = − 1 4 − n 2 −2P 2 s<strong>in</strong>(Pπ)<br />

, (C.50)<br />

4(n−P)(n+P) Pπ<br />

√<br />

2(2n 2 −P 2 )<br />

F n =<br />

2 ( n− P 2)(<br />

n+<br />

P<br />

) s<strong>in</strong>(Pπ 2 ) . (C.51)<br />

Pπ<br />

2 2<br />

For n ≠ n ′ , the sp<strong>in</strong> matrices have the form<br />

⎛<br />

〈Q x ,n | ˜H c | Q x ,n ′ 〉 = Q2 SO<br />

π P ⎜<br />

⎝<br />

0 0 0 0<br />

0 a ig d<br />

0 −ig b if<br />

0 d −if c<br />

⎞<br />

. (C.52)<br />

⎟<br />

⎠<br />

Calculat<strong>in</strong>g the matrix elements for n = 0,n ′ > 0, we get<br />

⎛(<br />

a = √ 1 1+(−1) n′) (<br />

s<strong>in</strong>(Pπ) 4 −1+(−1) n′) K x cos ( ) ⎞<br />

Pπ<br />

2<br />

⎝<br />

2 (n ′ −2P)(n ′ −<br />

⎠,<br />

+2P) (n ′ −P)(n ′ (C.53)<br />

+P)<br />

√ (<br />

2 1+(−1) n′) s<strong>in</strong>(Pπ)<br />

b = −<br />

(n ′ −2P)(n ′ , (C.54)<br />

+2P)<br />

⎛ (<br />

c = √ 1 ⎝ 4 −1+(−1) n′) K x cos ( ) (<br />

Pπ<br />

2<br />

1+(−1) n′) ⎞<br />

s<strong>in</strong>(Pπ)<br />

2 (n ′ −P)(n ′ + ⎠,<br />

+P) (n ′ −2P)(n ′ (C.55)<br />

+2P)<br />

d =<br />

(<br />

1+(−1) n′) s<strong>in</strong>(Pπ)<br />

√<br />

2(n ′ −2P)(n ′ +2P) ,<br />

⎛(<br />

−1+(−1) n′) (<br />

cos(Pπ)<br />

f = 2⎝<br />

2(n ′ −2P)(n ′ −<br />

+2P)<br />

⎛(<br />

−1+(−1) n′) (<br />

cos(Pπ)<br />

g = −2⎝<br />

2(n ′ −2P)(n ′ +<br />

+2P)<br />

1+(−1) n′) K x s<strong>in</strong> ( Pπ<br />

2<br />

(n ′ −P)(n ′ +P)<br />

1+(−1) n′) K x s<strong>in</strong> ( Pπ<br />

2<br />

(n ′ −P)(n ′ +P)<br />

) ⎞<br />

(C.56)<br />

⎠, (C.57)<br />

) ⎞<br />

⎠. (C.58)


Appendix C: Cooperon and <strong>Sp<strong>in</strong></strong> Relaxation 145<br />

And for n > 0,n ′ > 0, we get<br />

( ) Pπ<br />

a = R {2,+} s<strong>in</strong>(Pπ)+4K x R {1,−} cos , (C.59)<br />

2<br />

b = −2R {2,+} s<strong>in</strong>(Pπ),<br />

(C.60)<br />

( ) Pπ<br />

c = R {2,+} s<strong>in</strong>(Pπ)−4K x R {1,−} cos , (C.61)<br />

2<br />

d = R {2,+} s<strong>in</strong>(Pπ),<br />

(C.62)<br />

f = − √ (<br />

( )) Pπ<br />

2 R {2,−} cos(Pπ)+2K x R {1,+} s<strong>in</strong> , (C.63)<br />

g =<br />

√<br />

2<br />

(<br />

R {2,−} cos(Pπ)−2K x R {1,+} s<strong>in</strong><br />

2<br />

( Pπ<br />

2<br />

))<br />

, (C.64)<br />

with the functions<br />

R {1,±} =<br />

R {2,±} =<br />

(<br />

1±(−1) n+n′) (n 2 +n ′2 −P 2 )<br />

((n−n ′ )−P)((n+n ′ )−P)((n−n ′ )+P)((n+n ′ )+P) ,<br />

(<br />

1±(−1) n+n′) (n 2 +n ′2 −(2P) 2 )<br />

((n−n ′ )−2P)((n+n ′ )−2P)((n−n ′ )+2P)((n+n ′ )+2P) . (C.66)<br />

(C.65)


Appendix D<br />

Hamiltonian <strong>in</strong> [110] growth<br />

direction<br />

with<br />

The Cooperon Hamiltonian <strong>in</strong> the 0-mode approximation is given as follows<br />

⎛ ⎞<br />

A B C<br />

H c,0 = ⎜<br />

⎝ B ∗ D E ⎟<br />

⎠ +M q3,<br />

(D.1)<br />

C ∗ E ∗ F<br />

A = 1<br />

4q 2 W (q (<br />

2 4k<br />

2<br />

x +3 (˜q 1 2 +q2<br />

2 ))<br />

W<br />

( )<br />

q2 W<br />

−16k x˜q 1 s<strong>in</strong> + (˜q 1 2 −q 2<br />

2<br />

2)<br />

s<strong>in</strong>(q2 W)), (D.2)<br />

( ( ) )<br />

i 4k x s<strong>in</strong><br />

q2 W<br />

2<br />

− ˜q 1 s<strong>in</strong>(q 2 W)<br />

B = √ , (D.3)<br />

2W<br />

C = − q 2<br />

(˜q<br />

2<br />

1 +q2<br />

2 ) (<br />

W + q<br />

2<br />

2 − ˜q 1<br />

2 )<br />

s<strong>in</strong>(q2 W)<br />

, (D.4)<br />

4q 2 W<br />

D = q (<br />

2 2k<br />

2<br />

x + ˜q 1 2 ) ( +q2 2 W + q<br />

2<br />

2 − ˜q 1<br />

2 )<br />

s<strong>in</strong>(q2 W)<br />

, (D.5)<br />

2q 2 W<br />

( ( ) )<br />

i 4k x s<strong>in</strong><br />

q2 W<br />

2<br />

+ ˜q 1 s<strong>in</strong>(q 2 W)<br />

E = √ , (D.6)<br />

2W<br />

F = 1<br />

4q 2 W (q (<br />

2 4k<br />

2<br />

x +3 (˜q 1 2 +q2<br />

2 ))<br />

W<br />

( )<br />

q2 W<br />

+16k x˜q 1 s<strong>in</strong> + (˜q 1 2 −q 2<br />

2<br />

2)<br />

s<strong>in</strong>(q2 W)) (D.7)<br />

146


Appendix D: Hamiltonian <strong>in</strong> [110] growth direction 147<br />

and the term due to cubic Dresselhaus SOC<br />

M q3 =<br />

⎛<br />

q 3<br />

⎜<br />

⎝<br />

1<br />

4 s<strong>in</strong>c(q 2W)+ 3 1<br />

4<br />

0<br />

4 s<strong>in</strong>c(q 2W)− 1 4<br />

1<br />

0<br />

2 − 1 2 s<strong>in</strong>c(q 2W) 0<br />

1<br />

4 s<strong>in</strong>c(q 2W)− 1 1<br />

4<br />

0<br />

4 s<strong>in</strong>c(q 2W)+ 3 4<br />

⎞<br />

⎟<br />

⎠ .<br />

(D.8)


Appendix E<br />

Summation over the Fermi Surface<br />

The Cooperon Hamiltonian <strong>in</strong> the 2D case is given by<br />

H c = τv 2 {〈cos 2 (ϕ)〉(Q+2m e a.S) 2 x<br />

+〈s<strong>in</strong> 2 (ϕ)〉(Q+2m e a.S) 2 y<br />

+4m 2 e γ Dv 2 〈cos 2 (ϕ)s<strong>in</strong> 2 (ϕ)〉(Q+2m e a.S) x .S x<br />

−4m 2 eγ D v 2 〈s<strong>in</strong> 2 (ϕ)cos 2 (ϕ)〉(Q+2m e a.S) y .S y<br />

+(2m 3 e γ Dv 2 ) 2 (〈cos 2 (ϕ)s<strong>in</strong> 4 (ϕ)〉Sx<br />

2<br />

+〈s<strong>in</strong> 2 (ϕ)cos 4 (ϕ)〉S 2 y)},<br />

(E.1)<br />

with wave vector Q. We set<br />

m e ≡ 1,<br />

f 1 := 〈s<strong>in</strong> 2 (ϕ)〉,<br />

f 2 := 〈cos 2 (ϕ)〉,<br />

f 3 := 〈s<strong>in</strong> 2 (ϕ)cos 2 (ϕ)〉,<br />

f 4 := 〈s<strong>in</strong> 4 (ϕ)cos 2 (ϕ)〉,<br />

f 5 := 〈s<strong>in</strong> 2 (ϕ)cos 4 (ϕ)〉.<br />

(E.2)<br />

(E.3)<br />

(E.4)<br />

(E.5)<br />

(E.6)<br />

(E.7)<br />

Us<strong>in</strong>g the Matsubara trick we write<br />

∫ 2π<br />

0<br />

dϕ<br />

2π = 2<br />

πN<br />

N∑ 1<br />

√<br />

1− ( )<br />

. (E.8)<br />

s 2<br />

N<br />

s=1<br />

148


Appendix E: Summation over the Fermi Surface 149<br />

This gives us<br />

f 1 = 2 N−1<br />

∑<br />

πN<br />

s=1<br />

f 2 = 2<br />

πN<br />

f 3 = 2<br />

πN<br />

f 4 = 2<br />

πN<br />

f 5 = 2<br />

πN<br />

s 2<br />

N<br />

√1− ( ) , (E.9)<br />

2 s 2<br />

N<br />

√ N∑ ( s<br />

) 2,<br />

1−<br />

(E.10)<br />

N<br />

s=1<br />

N∑ ( s<br />

)<br />

√<br />

2 ( s<br />

) 2,<br />

1−<br />

(E.11)<br />

N N<br />

s=1<br />

N∑ ( s<br />

)<br />

√<br />

4 ( s<br />

) 2,<br />

1−<br />

(E.12)<br />

N N<br />

s=1<br />

N∑<br />

s=1<br />

Writ<strong>in</strong>g Eq.(E.1) <strong>in</strong> a compact way gives us Eq.(4.49).<br />

( ( )3<br />

s 2 ( s 2 2<br />

1− . (E.13)<br />

N)<br />

N)


Appendix F<br />

KPM<br />

The recursion relations <strong>of</strong> the polynomials <strong>of</strong> first and second k<strong>in</strong>d, Eq.(5.34)<br />

and Eq.(5.35), allow for a simple iteration procedure: The core <strong>of</strong> KPM is the iterative<br />

construction <strong>of</strong> the states |α n 〉 = T n ( ˜H)|α〉, described by the follow<strong>in</strong>g steps[WWAF06],<br />

|α 0 〉 = |α〉, (F.1)<br />

|α 1 〉 = ˜H|α 0 〉, (F.2)<br />

|α n+1 〉 = 2 ˜H |α n 〉−|α n−1 〉, (F.3)<br />

where |α〉 is the start<strong>in</strong>g vector. For <strong>in</strong>stance, to calculate the local DOS at site i, our<br />

start<strong>in</strong>g vector would be the site-occupation vector |i〉. Then the coefficients for the T n (E)<br />

polynomial are given by µ n = 〈i|i n 〉 and ρ i (E) is reconstructed by us<strong>in</strong>g Eq.(5.32). Consequently,<br />

the most time consum<strong>in</strong>g part is the matrix-vector multiplication H|α n 〉. Because<br />

H is sparse <strong>in</strong> our case, a very efficient multiplication is done by decompos<strong>in</strong>g the operator<br />

• <strong>in</strong> a matrix A which conta<strong>in</strong>s the <strong>in</strong>formation about connected sites, which comprises<br />

their site number and the type <strong>of</strong> hopp<strong>in</strong>g between them<br />

• and a matrix SOH which conta<strong>in</strong>s all hopp<strong>in</strong>g matrices which are two-dimensional<br />

due to SOC.<br />

The type <strong>of</strong> hopp<strong>in</strong>g is coded <strong>in</strong> a number h ij , e.g. h ij = 0 can be def<strong>in</strong>ed as “hopp<strong>in</strong>g from<br />

i to j ⇐⇒ hopp<strong>in</strong>g <strong>in</strong> the positive x-direction”. To give an example: The site i is connected<br />

with site j = A[i][2 ∗k] 1 and the type <strong>of</strong> hopp<strong>in</strong>g is h ij = A[i][2∗k +1], with k be<strong>in</strong>g an<br />

1 we use C-type <strong>of</strong> writ<strong>in</strong>g matrices<br />

150


Appendix F: KPM 151<br />

<strong>in</strong>teger runn<strong>in</strong>g from 0 up to the number <strong>of</strong> neighbors. The hopp<strong>in</strong>g matrix t(i,j), which<br />

<strong>in</strong>cludes the SOC, is then given by<br />

⎛<br />

t(i,j) =<br />

⎝ SOH[h ij][0][0] SOH[h ij ][0][1]<br />

SOH[h ij ][1][0] SOH[h ij ][1][1]<br />

This leads f<strong>in</strong>ally to the follow<strong>in</strong>g algorithm for the operation H.v = v out :<br />

⎞<br />

⎠. (F.4)<br />

List<strong>in</strong>g F.1: Optimized matrix-vector multiplication<br />

for ( m=0; m

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!