Вы находитесь на странице: 1из 243

UNIVERSITY OF CALIFORNIA

Santa Barbara
Optical control and detection of spin
coherence in semiconductor
nanostructures
A dissertation submitted in partial satisfaction
of the requirements for the degree of
Doctor of Philosophy
in
Physics
by
Jesse A. Berezovsky
Committee in charge:
Professor David D. Awschalom, Chairperson
Professor Andrew Cleland
Professor Leon Balents
December 2007
UMI Number: 3291324
3291324
2008
UMI Microform
Copyright
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.
ProQuest Information and Learning Company
300 North Zeeb Road
P.O. Box 1346
Ann Arbor, MI 48106-1346
by ProQuest Information and Learning Company.
The dissertation of Jesse A. Berezovsky is approved:
Committee Chairperson
December 2007
Optical control and detection of spin coherence in
semiconductor nanostructures
Copyright c _ 2007
by
Jesse A. Berezovsky
iii
Acknowledgments
There are many people who deserve credit for making my experience
at Santa Barbara what it was: educational, productive, exciting, and fun.
Foremost, I couldnt have been more fortunate than to have David as my
advisor. Not only does he provide an unparalleled environment in which
to learn and work, but also takes his role as advisor seriously, shepherding
us through the valley of the shadow of death (or grad school, as it is also
known).
In Davids lab, I have had the fortune to work with a number of tal-
ented collaborators. Over the last few years, I have worked closely with
Maiken Mikkelsen on the single spin work. The collaboration was produc-
tive under normal circumstances, but when the data was really owing,
like the individual Voltron robots joining together to make a giant robot,
we transformed into an unstoppable 24-hour-a-day data-taking machine.
This work also could not have been done without the materials expertise
of Nick Stoltz or the theoretical stylings of Oliver Gywat. Before that, on
the nanocrystal work, Min Ouyang initiated me into the esoteric secrets
of chemistry, and taught me not to fear the eects of pyridine on ones
reproductive organs. David Steuerman and Yong-Qing Li then joined me
on the nanocrystal work, of which there was a large quantity even if it
wasnt always successful. Of course, we wouldnt have been able to un-
derstand the results without the theory eorts of both Oliver and Florian
Meier. And right after I joined the group, I was lucky to jump into work
with Jason Stephens and Roland Kawakami on imaging nuclear spins in
semiconductors.
Along, with Jason and Roland, I am grateful to my elders in the lab,
iv
who showed me the ropes and taught me most of what I know about
doing lab work. In particular, Ryan Epstein (bouncing his latest quantum
computer design o me, or organizing a chair race in the hallway), Yui
Kato (trying to explain Group Theory to me, or calling an impromptu Beer
Meeting), and Martino Poggio and Roberto Myers (spewing vitriol at each
other about the correct sign of the g-factor or something, then suddenly
going o sailing together). Also, a number of postdocs came through the
lab, sharing their unique expertise: Ronald Hanson, Alex Holleitner, and
Sai Ghosh, in addition to those already mentioned.
Other labmates who were my own contemporaries, provided valuable
insight and support (both scientic and moral). Felix Mendoza has been
my comrade-in-arms since prospective student visiting day. He can always
be counted on to solve problems ranging from malfunctioning door locks
to broken snorkels. Hadrian Knotz has provided a calm presence in the
lab, somehow managing to keep all the lab computers running smoothly.
And whenever I felt a need to talk to someone about physics, Nate Stern
was there for me.
I wish good luck to all of the newer additions to the lab, whose antics
in and out of the lab have provided much entertainment.
The various sta members who keep the whole ship aoat deserve many
thanks. Holly Woo and the other CNSI sta have been tireless in their
eorts to counteract my great negligence in nancial and bureaucratic
matters. Mike Deal and the facilities sta have always discharged their
building-maintenance duties admirably. Je Dutter and the rest of the
machine shop guys have been an amazing resource particularly, Mike
Wrocklage who oversees the student machine shop with a Buddha-like
calm. And nally, Id like to acknowledge the cleanroom sta who manage
v
to keep that show going.
Of course, my years in Santa Barbara have not been all work. There
have been a number of people outside the lab that have made the experi-
ence lots of fun. Through numerous parties, various barbecues, countless
Wednesday Wine Nights, and an innite number of happy hours, Felix
Mendoza, Dave Wood, Corrinne Mills, Jean-Luc Fraikin, Chris McKen-
ney, Melvin McLaurin, and various others have kept my whistle wet, my
belly stued with good food, and my mental state sane. Also, the rst
year study group, consisting of Sara Hastings-Simon, Felix Mendoza, Juan
Hodelin and myself, dulled the pain of coursework and provided lots of
great memories the interstices of an evenings work were always packed
with fun. And of course, our rst-year coursework would not have been
nearly as successful if it werent for the assistance Dr. Skipper, who never
failed to slake our thirst for knowledge.
vi
Vit
Education
2002 B.S., Physics, University of Minnesota.
2005 M.A., Physics, University of California, Santa Barbara.
2007 Ph.D., Physics, University of California, Santa Barbara.
Publications
Manipulation of spins and coherence in semiconductors, N. P. Stern,
J. Berezovsky, S. Ghosh and D. D. Awschalom, in Handbook on Magnetism
and Advanced Magnetic Materials, H. Kronm uller and S. Parkin, eds., John
Wiley & Sons (2007).
Spin coherence in semiconductors, J. Berezovsky, W. H. Lau, S.
Ghosh, J. Stephens, N. P. Stern, and D. D. Awschalom, in Manipulat-
ing Quantum Coherence in Solid State Systems, M. E. Flatte and I. T ifrea,
eds., Springer (2007).
Nondestructive optical measurements of a single electron spin in a
quantum dot, J. Berezovsky, M. H. Mikkelsen, O. Gywat, N. G. Stoltz,
L. A. Coldren, and D. D. Awschalom, Science 314, 1916 (2006).
Initialization and read-out of spins in coupled core-shell quantum
dots, J. Berezovsky, O. Gywat, F. Meier, D. Battaglia, X. Peng, and
D. D. Awschalom, Nature Physics 2, 831 (2006).
Cavity enhanced Faraday rotation of semiconductor quantum dots,
vii
Y. Q. Li, D. W. Steuerman, J. Berezovsky, D. S. Seferos, G. C. Bazan, and
D. D. Awschalom. Appl. Phys. Lett. 88, 193126 (2006).
Spin dynamics and level structure of quantum-dot quantum wells, J.
Berezovsky, M. Ouyang, F. Meier, D. D. Awschalom, D. Battaglia, and X.
Peng. Phys. Rev. B 71, 081309(R) (2005).
Spintronics: Semiconductors, molecules, and quantum information,
Y. Kato, J. Berezovsky, and D. D. Awschalom. IEDM Technical Digest,
IEEE International, p.537 (2004).
Optically patterned nuclear doughnuts in GaAs/MnAs heterostruc-
tures, J. Stephens, J. Berezovsky, R. K. Kawakami, A. C. Gossard, and
D. D. Awschalom. Appl. Phys. Lett. 85, 1184 (2004).
Spin accumulation in forward-biased MnAs/GaAs Schottky diodes,
J. Stephens, J. Berezovsky, J. P. McGuire, L. J. Sham, A. C. Gossard, and
D. D. Awschalom. Phys. Rev. Lett. 93, 097602 (2004).
Spatial imaging of magnetically patterned nuclear spins in GaAs, J.
Stephens, R. K. Kawakami, J. Berezovsky, M. Hanson, D. P. Shepherd, A.
C. Gossard, and D. D. Awschalom. Phys. Rev. B 68, 041307 (2003).
Imaging of spin dynamics in closure domain and vortex structures, J.
P. Park, P. Eames, D. M. Engebretson, J. Berezovsky, and P. A. Crowell.
Phys. Rev. B 67, 020403(R) (2003).
Spatially resolved dynamics of localized spin-wave modes in ferromag-
netic wires, J. P. Park, P. Eames, D. M. Engebretson, J. Berezovsky, and
P. A. Crowell. Phys. Rev. Lett. 89, 277201 (2002).
viii
Time-domain ferromagnetic resonance in epitaxial thin lms, D. M.
Engebretson, J. Berezovsky, J. P. Park, L. C. Chen, C. J. Palmstrm, and
P. A. Crowell. J. Appl. Phys. 91, 8040 (2002).
Control of magnetic anisotropy in Fe
1x
Co
x
lms on vicinal GaAs and
Sc
1y
Er
y
As surfaces, A. F. Isakovic, J. Berezovsky, P. A. Crowell, L. C.
Chen, D. M. Carr, B. D. Schultz, and C. J. Palmstrm. J. Appl. Phys.
89, 6674 (2001).
Epitaxial ferromagnetic metal/GaAs(100) heterostructures, L. C. Chen,
J. W. Dong, B. D. Schultz, C. J. Palmstrm, J. Berezovsky, A. Isakovic,
P. A. Crowell, and N. Tabat. J. Vac. Sci. Technol. B 18, 2057 (2000).
Fields of study
Major eld: Physics
Optical control and detection of spin coherence in semiconductor
nanostructures
Professor David D. Awschalom
ix
Abstract
Optical control and detection of spin coherence in
semiconductor nanostructures
by
Jesse A. Berezovsky
Understanding the coherent dynamics of electron spins in quantum dots
(QDs) is important for potential applications in solid-state, spin-based elec-
tronics and quantum information processing. Here, results are presented
focusing on optical initialization, manipulation, and readout of spin co-
herence in various semiconductor nanostructures. Layered semiconduc-
tor nanocrystals are fabricated containing a spherical quantum shell in
which electrons and holes are conned. As in a planar quantum well, the
quantized energy levels and g-factors are found to depend on the shell
thickness. Taking this idea a step further, nanocrystals with a concentric,
tunnel-coupled core and shell are investigated. Based on the energy and
g-factor dependences in these structures, spins can be selectively initial-
ized into, and read out from, states in the core and shell. In contrast to
these two ensemble measurements, we next turn to measurements of sin-
gle electron spins in single QDs. First, we demonstrate the detection of a
single electron spin in a QD using a nondestructive, continuously averaged
magneto-optical Kerr rotation (KR) measurement. This continuous sin-
gle QD KR technique is then extended into the time domain using pulsed
x
pump and probe lasers, allowing the observation of the coherent evolution
of an electron spin state with nanosecond temporal resolution. By sweep-
ing the delay between the pump and probe, the dynamics of the spin in
the QD are mapped out in time, providing a direct measurement of the
electron g-factor and spin lifetime. Finally, this time-resolved single spin
measurement is used to observe ultrafast coherent manipulation of the spin
in the QD using an o-resonant optical pulse. Via the optical Stark eect,
this optical pulse coherently rotates the spin state through angles up to
radians, on picosecond timescales.
xi
Contents
Chapter 1 Introduction 1
1.1 Spin: A brief history . . . . . . . . . . . . . . . . . . . . . 1
1.2 Spintronics . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Quantum information processing . . . . . . . . . . . . . . 8
Chapter 2 Background 11
2.1 Spins in semiconductors . . . . . . . . . . . . . . . . . . . 11
2.2 Quantum dots . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Optical properties of spins in semiconductors . . . . . . . . 29
2.4 Faraday rotation: theory . . . . . . . . . . . . . . . . . . . 36
2.5 Faraday rotation: experiment . . . . . . . . . . . . . . . . 41
Chapter 3 Quantum Shells in Semiconductor Nanocrystals 45
3.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2 Sample structure and characterization . . . . . . . . . . . 49
3.3 Experimental setup . . . . . . . . . . . . . . . . . . . . . . 51
3.4 Spin dynamics in quantum shells . . . . . . . . . . . . . . 53
3.5 Energy levels in quantum shells . . . . . . . . . . . . . . . 56
3.6 Theoretical description . . . . . . . . . . . . . . . . . . . . 59
xii
Chapter 4 Coupled Shells in Layered Colloidal Nanocrystals 68
4.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.2 Sample structure and characterization . . . . . . . . . . . 69
4.3 Experimental setup . . . . . . . . . . . . . . . . . . . . . . 73
4.4 Theoretical description . . . . . . . . . . . . . . . . . . . . 76
4.5 Transient absorption and luminescence measurements . . . 80
4.6 Time-resolved Faraday rotation spectroscopy . . . . . . . . 84
4.7 Analysis of spin dynamics and core-shell coupling . . . . . 89
4.8 Nanocrystal QDs in an optical cavity . . . . . . . . . . . . 94
Chapter 5 Non-destructive Measurement of a Single Electron
Spin 101
5.1 Motivation and Background . . . . . . . . . . . . . . . . . 101
5.2 Sample structure . . . . . . . . . . . . . . . . . . . . . . . 107
5.3 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5.4 Experimental setup . . . . . . . . . . . . . . . . . . . . . . 113
5.5 Characterization: PL and Hanle measurements . . . . . . . 117
5.6 Single spin Kerr rotation . . . . . . . . . . . . . . . . . . . 124
Chapter 6 Coherent Dynamics of a Single Spin 139
6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
6.2 Experimental scheme . . . . . . . . . . . . . . . . . . . . . 141
6.3 Theoretical model . . . . . . . . . . . . . . . . . . . . . . . 145
6.4 Time-resolved single spin measurements . . . . . . . . . . 148
Chapter 7 Ultrafast Manipulation of Single Spin Coherence156
7.1 Motivation and Background . . . . . . . . . . . . . . . . . 156
7.2 Experimental scheme . . . . . . . . . . . . . . . . . . . . . 159
xiii
7.3 Theoretical model . . . . . . . . . . . . . . . . . . . . . . . 167
7.4 Ultrafast optical spin manipulation . . . . . . . . . . . . . 171
7.5 Further exploration and control measurements . . . . . . . 174
Chapter 8 Conclusion 183
Appendix A Sample structure and processing 187
Appendix B Details of single spin detection and manipula-
tion 191
B.1 Optical path . . . . . . . . . . . . . . . . . . . . . . . . . . 193
B.2 Measurement control scheme . . . . . . . . . . . . . . . . . 196
B.3 Odds and ends . . . . . . . . . . . . . . . . . . . . . . . . 197
Appendix C Other theoretical views of Faraday rotation 200
xiv
List of gures
1.1 The Datta-Das spin transistor. . . . . . . . . . . . . . . . . 6
2.1 The Bloch sphere. . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Typical direct-gap band diagram. . . . . . . . . . . . . . . 14
2.3 Spin decoherence: a random walk. . . . . . . . . . . . . . . 20
2.4 Schematic of conning potential, energy levels, and wave-
functions in a quantum dot. . . . . . . . . . . . . . . . . . 26
2.5 Types of quantum dots. . . . . . . . . . . . . . . . . . . . 28
2.6 Selection rules for interband optical transitions. . . . . . . 32
2.7 Schematic of a Hanle measurement setup. . . . . . . . . . 34
2.8 Illustration of the Hanle eect. . . . . . . . . . . . . . . . . 35
2.9 Diagram of the Faraday eect. . . . . . . . . . . . . . . . . 37
2.10 Theoretical Faraday rotation spectra. . . . . . . . . . . . . 40
2.11 Typical time-resolved Faraday rotation setup. . . . . . . . 42
2.12 Sensitive polarization detection with a balanced photodiode
bridge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.1 Cutaway schematic of layered nanocrystals. . . . . . . . . . 46
3.2 Photoluminescence of quantum shell samples. . . . . . . . 50
3.3 Setup for quantum shell TRFR. . . . . . . . . . . . . . . . 51
xv
3.4 Spin precession in quantum shells. . . . . . . . . . . . . . . 53
3.5 Temperature dependence of the quantum shell spin lifetime. 55
3.6 Precession frequency and g-factors as a function of shell
thickness. . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.7 Quantum shell Faraday rotation spectra. . . . . . . . . . . 58
3.8 Calculated electron and hole energy levels in quantum shells. 60
3.9 Calculated electron and hole wavefunctions in a quantum
shell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.10 Calculated FR spectra of quantum shells. . . . . . . . . . . 64
4.1 Diagram of a coupled core-shell nanocrystal. . . . . . . . . 70
4.2 Photoluminescence of coupled core-shell nanocrystals. . . . 71
4.3 Coupled core-shell PL as a function of core and shell thickness. 73
4.4 Setup for TRFR and TA measurements on coupled core-shell
nanocrystals. . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.5 Calculated core and shell wavefunctions. . . . . . . . . . . 78
4.6 Measured and calculated energy levels as a function of core
and shell thickness. . . . . . . . . . . . . . . . . . . . . . . 79
4.7 Transient absorption measurements on coupled core-shell
nanocrystals. . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.8 Time-resolved photoluminescence of coupled core-shell nanocrys-
tals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.9 Time-resolved Faraday rotation on coupled core-shell nanocrys-
tals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.10 Three precession frequencies in coupled-core shell nanocrys-
tals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.11 Faraday rotation spectra of coupled core-shell nanocrystals. 89
xvi
4.12 Coupled wavefunctions determined using the Hubbard model 92
4.13 Diagram of nanocrystals embedded in an optical cavity . . 97
4.14 Optical characterization of cavity structures. . . . . . . . . 98
4.15 Cavity-enhanced time-resolved Faraday rotation. . . . . . . 99
5.1 Previously demonstrated schemes for single spin detection. 102
5.2 Single versus ensemble Kerr rotation spectrum. . . . . . . 106
5.3 Schematic of the sample for single spin measurements. . . 108
5.4 STM image and cartoon of interface uctuation quantum
dots. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.5 Characterization of the single spin sample. . . . . . . . . . 110
5.6 Setup for single spin detection. . . . . . . . . . . . . . . . . 114
5.7 Photoluminescence of a single QD vs. bias voltage, and
polarization thereof. . . . . . . . . . . . . . . . . . . . . . 118
5.8 Pump power dependence of various single QD PL lines. . . 119
5.9 Lowest energy optical transitions in a quantum dot. . . . . 120
5.10 Mechanisms for single spin initialization. . . . . . . . . . . 121
5.11 Single quantum dot Hanle measurements. . . . . . . . . . . 123
5.12 Single quantum dot Kerr rotation. . . . . . . . . . . . . . . 125
5.13 Single spin KR feature as a function of bias voltage. . . . . 127
5.14 Single spin detection in other quantum dots. . . . . . . . . 128
5.15 Energy of single spin Kerr rotation feature compared to tran-
sition energies. . . . . . . . . . . . . . . . . . . . . . . . . 129
5.16 Single spin Kerr rotation Hanle measurements. . . . . . . . 132
5.17 Analysis of single spin measurements as a function of bias
voltage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
5.18 A proposal to use Faraday rotation to couple single spins. . 137
xvii
6.1 Setup for time-resolved single spin measurements. . . . . . 142
6.2 Temporal prole of the pump and probe pulses. . . . . . . 143
6.3 Illustration of spin misalignment leading to dynamic nuclear
polarization. . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.4 Coherent precession of a single spin. . . . . . . . . . . . . . 149
6.5 Single spin precession as a function of magnetic eld. . . . 151
6.6 Single spin dynamics at zero magnetic eld. . . . . . . . . 153
6.7 The eects of nuclear polarization on single spin precession. 154
7.1 Illustration of the optical Stark eect. . . . . . . . . . . . . 157
7.2 Setup for Stark tipping measurements. . . . . . . . . . . . 160
7.3 Energy scales and relevant optical spectra for Stark tipping
measurements. . . . . . . . . . . . . . . . . . . . . . . . . . 161
7.4 Single spin Kerr rotation vs. delay and probe energy with
and without the tipping pulse. . . . . . . . . . . . . . . . . 165
7.5 Sequence of rotations in the Stark tipping model. . . . . . 167
7.6 Coherent rotation of a single electron spin. . . . . . . . . . 172
7.7 Dependence of Stark tipping on tipping pulse intensity and
detuning. . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
7.8 Strength of the Stark eect as a function of detuning. . . . 176
7.9 Comparison of the Stark eect with tipping pulses of oppo-
site helicity. . . . . . . . . . . . . . . . . . . . . . . . . . . 178
7.10 Comparison of measurement with circularly and linearly po-
larized pump. . . . . . . . . . . . . . . . . . . . . . . . . . 180
B.1 Setup for single spin measurement and control. . . . . . . . 192
C.1 Classical view of Faraday rotation. . . . . . . . . . . . . . 201
xviii
Chapter 1
Introduction
1.1 Spin: A brief history
Spin, the intrinsic angular momentum of a particle, was rst described the-
oretically by George Uhlenbeck and Samuel Goudsmit in 1925, and formal-
ized by Wolfgang Pauli in 1926. Experimentally, however, spin phenomena
have been observed and put to practical use for much longer. The earli-
est known spin-based device is most likely the magnetic compass. Here,
a freely rotating needle is constructed out of a material in which electron
spins align with each other under their mutual exchange interaction. This
leads to a macroscopic spin polarization in the needle (ferromagnetism),
causing the needle to align with the Earths magnetic eld due to the Zee-
man energy of a spin in a magnetic eld (see Eq. 2.3). Written records
describing these phenomena exist dating back to the 4th century B.C.
1
However, the earliest known use of a magnetic compass far predates the
appearance of humans on earth, and can be found not in written records,
but in the fossil record.
The oldest known magnetic compass is found in the Gunint Iron For-
mation in Minnesota and Ontario, a region rich in fossils dating back
approximately 2 billion years. Included among this ancient zoo of mi-
croscopic organisms are the remains of creatures known as magnetotactic
bacteria
1
[1, 2]. The distinguishing feature of these bacteria are one or
more microscopic, ferromagnetic needles used to orient the creature with
respect to the earths magnetic eld, aiding in the search for the optimal
oxygen concentration. By following magnetic eld lines in their aquatic
environment, this search is reduced from three dimensions to one.
Some 2 billion years after these bacteria were immortalized in stone,
the now-ubiquitous hairless chimps began to catch on. There is some ev-
idence that the Olmec people of Mesoamerica may have invented a type
of compass earlier than 1000 B.C. However, the rst written description of
ferromagnetism is found in a 4th century B.C. Chinese text entitled, Book
of the Devil Valley Master [3, 4]: Thus, one should at the beginning of
1
It should be noted that the age of 2 billion years is merely one theory. Young Earth
creationists hold that magnetotactic bacteria were created less than 10,000 years ago.
2
cognition rst perceive oneself and then other people. Then the mutual
cognition becomes as clear as the vision of two-eyed sh and as the light
following the shadows. Then one will miss nothing upon examination of
the word, like the magnet draws the needle and the tongue the fried bone
to itself.
2
Over the next twenty-three hundred years or so, the knowledge of mag-
netism spread across the globe. Clever minds devised new uses for the
phenomenon, and rened old ones, ranging from the electric motor, to the
dynamo, to the posting of notes on a refrigerator. Despite its bountiful
technological applications, at the beginning of the twentieth century, the
physical origins of magnetism were still unclear.
As quantum mechanics was being developed in the 1920s, great strides
were made in understanding atomic spectra by quantizing the orbital mo-
mentum of electrons around the atomic nucleus. However, results such as
the Stern-Gerlach experiment, and unexplained splittings in atomic spec-
tra (the anomalous Zeeman eect, and hyperne splitting) indicated that
there were extra quantum degrees of freedom not being taken into account.
2
Translated from the Chinese into German by Kimm (1927) [3], and from the German
into English by F. Mendoza (2007). There is some debate about the dating of this
manuscript (see [4]), though other sources mentioning magnetism from the same period
also exist.
3
A natural candidate for this unknown quantity was the angular mo-
mentum of a particle. The idea was at rst considered to be impossible.
Given the known upper bound on the radius of the electron, the angular
velocity of the electron would need to be impossibly high to provide the
observed splittings. Nevertheless, Uhlenbeck and Goudsmit published the
idea in 1925. Despite its apparent impossibility, the idea of spin nicely
explained the observations. Originally a skeptic, Pauli warmed to the idea
and ran with it, redening spin not as an actual rotation of a particle, but
as an angular momentum intrinsic to the particle, just as charge or mass
are intrinsic properties. He then went on to develop a formalism for dealing
with spin in (non-relativistic) quantum mechanics (see, for example, [5]).
Once this theoretical framework was in place, the experimental study
of spin physics could now proceed hand-in-hand with theory, instead of
the pure phenomenology of the previous millennia. Throughout the rest
of the twentieth century numerous advances were made, such as a detailed
understanding of magnetic materials, nuclear spin physics, and spin reso-
nance phenomena. These discoveries led to revolutionary technologies such
as magnetic resonance imaging (MRI) and magnetic data storage (tapes,
hard drives).
The reservoir of interesting spin phenomena is still far from dry. Re-
4
cent advances in materials, electronics, and low-temperature technologies
have brought new untapped wells of spin physics within reach. But how
will we know where to sink a new well? Like a magnetotactic bacterium
following its magnetic needle, we will use the promise of future spin-based
technologies as our divining rod to lead us to new and interesting physics.
1.2 Spintronics
One promising area of application for spin physics is in the eld of spin-
based electronics, or spintronics. Here, the idea is to integrate magnetic
functionality with more traditional, electronic devices. Virtually all elec-
tronic devices are based on the distribution and ow of electrons within a
material. These electrons have spin of course, but in general, the spin is
ignored as far as the functionality of the device is concerned.
Recently, the rst spintronic devices have successfully integrated mag-
netic and electronic functionality. The ow of spin-polarized electrons
through layered magnetic structures has revolutionized hard-drive storage
technology through the phenomenon of giant magnetoresistance (GMR).
Also, new types of non-volatile memory (magnetic RAM: MRAM) are on
the market that integrate magnetic memory and electronic devices in a
single architecture. Future renements of this idea, possibly using spin
5
ferromagnetic
spin injector
ferromagnetic
spin analyzer
gate
V
g
semiconductor channel
I
Figure 1.1: The Datta-Das spin transistor. Spin polarized electrons are injected from
a ferromagnetic source into a semiconductor channel, where they are manipulated using
a gate voltage, and then read out using a ferromagnetic drain contact.
torque devices [6], may play a signicant role in consumer electronics.
Both MRAM and GMR devices are based on metallic magnetic mate-
rials. But it is also interesting to consider the potential of semiconductor
spintronic devices. Semiconducting materials are the backbone of modern
electronics, and oer a wide range of functionality, controllable through
combinations of material composition and doping. The further integration
of semiconductor devices with magnetic properties oers another dimen-
sion of potential applications.
The canonical semiconductor spintronic device is the spin transis-
tor as proposed by Supriyo Datta and Biswajit Das in 1990, shown in
Fig. 1.1 [7]. While this may or may not be a practical or useful device, it
captures some essential features of a spintronic device: generation, manip-
6
ulation, and detection of spin polarization. The Datta-Das spin transistor
consists of a semiconductor channel with ferromagnetic source and drain
contacts at either end. When a voltage is applied across the source and
drain, spin-polarized electrons ow from the source contact into the semi-
conductor. After traversing the semiconductor channel, the electrons ow
into the drain contact. The conductivity across the semiconductor-drain
interface will depend on the spin polarization of the current, relative to the
spin-dependent density of states in the drain contact. Therefore, by ma-
nipulating the spin of the electrons as they ow through the channel (here,
eected in some way by a gate electrode), the source-drain conductivity
can be modulated.
Most, if not all, of the components of the Datta-Das spin transistor have
been demonstrated in individual devices. However, putting them together
in a commercially viable device has so far proved elusive. Clearly, more
research is needed to nd a practical use for spins in semiconductor devices.
Just as the American Plains Indians put the scrotum of the bualo to good
use
3
, so too should we strive to use every part of the electron.
3
For ceremonial rattles [8]
7
1.3 Quantum information processing
In the typical view of spintronic devices, the spin is thought of classically,
as a vector in 3-space. That is, quantum eects such as coherence or
entanglement are ignored. But if these eects can be understood and
controlled, a number of powerful applications may become possible.
In general, exploiting uniquely quantum eects in spins or other sys-
tems goes by the name of quantum information processing. This eld can
be divided into two categories: quantum computing, and quantum com-
munication.
Quantum computing is predicted to oer exponential speedup of certain
computational problems, eectively solving problems, such as factoring
large numbers, that are currently impossible for classical computers. In
a classical computer, information is stored and processed in bits, each
of which can take on two values (0 or 1). Extending this idea into
the quantum realm, a quantum bit, or qubit, can exist in a coherent
superposition of 0 and 1 and furthermore, an ensemble of qubits can
exist in a superposition, entangled with one another. These superpositions,
in eect, allow many computations to be performed in parallel, leading to
the improvement in computing power.
In order to build a qubit, one needs a two-state quantum system that
8
can be initialized into a particular state, and subsequently measured. One
then needs to construct a system in which quantum gates can be applied
to one or more qubits, in analogy to the logical gates that act on classical
bits. Additionally, the system must be scalable up to large numbers of
qubits, and the qubits must maintain their coherence over suciently long
timescales such that a large number of gate operations can be performed
within the coherence time.
An electron spin is a popular candidate for a qubit, since it is a natural
two-state system. Electron spins in semiconductors have received much
attention for quantum information applications because 1. semiconductor
processing technology should make the scaling to large systems easier, and
2. electron spins in semiconductors have been found to have long coherence
times relative to the expected times for gate operations. In recent years, a
number of schemes have been demonstrated for achieving the requirements
of state initialization, readout, and control for spin qubits (see Chapters
5-7). Nevertheless, there is still a long way to go before these elements can
be put together in a functioning quantum computer. This is good news for
a physicist, and bad news for an engineer.
Quantum communication involves the transmission of quantum infor-
mation from one place to another. This has applications in secure commu-
9
nication (cryptography) and in teleportation of quantum states. Quantum
communication necessitates a ying qubit a carrier of quantum infor-
mation that can be moved from place to place. Though spins and other
qubit candidates such as single atoms can be moved over micron-scale dis-
tances within the coherence time, the only practical qubits for long-distance
quantum communication are photon-based. Photons make ideal carriers
of quantum information because they travel fast, and they have very long
coherence times. The ip side to the long coherence time, is that photons
interact very weakly with each other - and a strong controllable interaction
is a desired feature for quantum information processing. This leads one to
consider a hybrid system with stationary qubits used for quantum com-
puting at either end, and ying qubits communicating between the two.
This requires a way of converting stationary qubits to ying qubits. Fortu-
nately, spins in semiconductors couple to photons in a variety of ways (as
discussed below), making this an intriguing platform for a potential hybrid
quantum computing/communication system.
10
Chapter 2
Background
2.1 Spins in semiconductors
According to quantum mechanics, angular momentum can be described by
two quantum numbers: total angular momentum L, and the projection of
angular momentum on the (say) z-axis: L
z
. These quantum numbers are
derived from eigenvalues of the commuting operators

L
2
and

L
z
:

L
2
[) =
L(L + 1) h
2
[) and

L
z
[) = L
z
h[), where

L
2
=

L
2
x
+

L
2
y
+

L
2
z
, and

L

is the angular momentum operator along the -direction. L can take on


half-integer values, and for a given L, L
z
can take on values L
z
= L, L+
1, ..., L.
In the case of a particles spin, the total angular momentum S is xed,
and the projection of the spin can take on 2S + 1 values, from S to S.
An electron has total spin S = 1/2, and S
z
= 1/2. Therefore, there are
11
two eigenstates one with S = 1/2 and S
z
= +1/2 denoted [ ) and the
other with S = 1/2 and S
z
= 1/2 denoted [). A general spin state of an
electron is then given by
[) = [) +[) (2.1)
where and are complex numbers such that [[
2
+[[
2
= 1.
In this basis of states, it is convenient to write down the matrix forms
of the operators

S

S
x
=
h
2
_
_
_
_
_
0 1
1 0
_
_
_
_
_

S
y
=
h
2
_
_
_
_
_
0 i
i 0
_
_
_
_
_

S
z
=
h
2
_
_
_
_
_
1 0
0 1
_
_
_
_
_
,
where
_
_
_
_
_

_
_
_
_
_
represents the state given by Eq. 2.1.
These matrices without the factor of h/2 are known as Pauli matrices,
and the vector

S =
_

S
x
,

S
y
,

S
z
_
is the spin operator for the electron spin-
1/2.
Note that there are four degrees of freedom in the two complex coe-
cients and in Eq. 2.1. However, the normalization requirement removes
one of these degrees of freedom, and another can be ignored as an overall
phase. Thus there are only two degrees of freedom that we care about, and
Eq. 2.1 can be rewritten
[) = cos

2
[) +e
i
sin

2
[). (2.2)
12
) cos , sin sin , sin (cos S
v
+
+ i


x
y
z
Figure 2.1: The Bloch sphere. The vectors pointing to the north and south poles of
the Bloch sphere represent the up and down eigenstates, with the rest of the sphere
representing superpositions of up and down.
The two parameters and can be thought of as the polar and az-
imuthal angles dening a point on a sphere. This is known as the Bloch
sphere (Fig. 2.1), and turns out to be a very useful way of picturing a
spin-1/2 state. The usefulness of this picture can be seen by looking at
the expectation values of the spin in the x, y, and z directions. Using the
matrix forms of the

S

operators given above, it is easy to show that the


corresponding expectation values are

S
x
) =
h
2
cos sin

S
y
) =
h
2
sin sin

S
z
) =
h
2
cos .
These expectation values are equivalent to the x, y, and z components
13
0
k
heavy holes
J=3/2, J
z
=3/2
light holes
J=3/2, J
z
=1/2
split-off holes
J=1/2, J
z
=1/2
Conduction band
s-like: L=0
E
g

E
Valence band
p-like: L=1
electrons
S=1/2, S
z
=1/2
Figure 2.2: Typical direct-gap band diagram. Near the zone center, the dispersion
of the energy bands in a direct-gap semiconductor can be approximated as parabolas.
The conduction band (red) has zero orbital angular momentum (L = 0). The valence
bands (black) have L = 1, which splits the band into three subbands: heavy holes, light
holes, and split-o holes. E
g
is the band-gap, and is the spin-orbit coupling energy.
of the Bloch vector, as shown in Fig. 2.1. Therefore, it is correct in some
sense to think of the spin as actually pointing along the vector on the
Bloch sphere. This one-to-one correspondence between the quantum state
and the intuitive picture of a classical angular momentum vector is ap-
parently just a coincidence. For spin other than 1/2, there is no such
direct correspondence. But since here we are typically interested in elec-
tron spins, the Bloch sphere provides a useful and intuitive way of thinking
about quantum spin states.
14
The results described below will focus on electron spins in semiconduct-
ing materials. In a semiconductor, a completely lled electronic band (the
valence band) is separated from an empty band (the conduction band) by
an energy gap, E
g
on the order of 1 eV. Fig. 2.2 shows the typical dispersion
of the conduction and valence bands in a direct gap semiconductor (where
the valence band maximum and conduction band minimum occur at the
same quasi-momentum k). In general, the dispersion has more complex
structure as a function of k, but suciently close to the conduction band
minimum (and valence band maximum) we can approximate the curvature
as parabolic. Fortunately, we will be mainly interested in electrons residing
in this region.
In all of the semiconductors considered here, the valence band is derived
from the lled p-orbitals of the constituent atoms, and the conduction
band from the unlled s-orbitals. At the center of the Brillouin zone (k =
0), an electron in the conduction band has s-symmetry (orbital angular
momentum L = 0) and an electron in the valence band has p-symmetry
(L = 1). Away from the zone center, these pure s and p states are not
eigenstates. Matrix elements of the operator kp mix states between bands,
yielding some p character in the conduction band and some s character in
the valence band. Near the zone center, one can often ignore this mixing
15
to rst order (the conduction band states are s-like and the valence
band states are p-like). For a detailed discussion of semiconductor band
structure, see Ref. [9].
Since the conduction band states are s-like, there is no orbital angular
momentum, and the total angular momentum of an electron is just the spin:
S = 1/2. However, the situation in the valence band is more complicated.
Here, the electrons have orbital angular momentum, L = 1 in addition to
their spin angular momentum. The eigenstates of the valence band now
must be described by the quantum numbers J and J
z
, derived from the
eigenvalues of the total angular momentum operator squared

J
2
= (

L+

S)
2
,
and the projection

J
z
, respectively. According to the rules for addition
of angular momentum, J can be either [L + S[ or [L S[ and J
z
can
take on values J, J + 1, ..., J. Therefore, in the present case, there
are six angular momentum subbands in the valence band with (J, J
z
) =
(
3
2
,
3
2
), (
3
2
,
1
2
), (
1
2
,
1
2
).
Due to the interaction of the spin and orbital angular momentum in
the valence band (spin-orbit coupling), the energies and dispersions of the
dierent subbands are altered. The two J = 1/2 bands are split from the
four J = 3/2 bands by the spin-orbit interaction energy . Furthermore,
the curvature of the valence subband and conduction band dispersions
16
are altered by the k p matrix elements between the dierent bands, as
mentioned above. These curvatures are interpreted as an eective mass
for electrons in the dierent bands. Fig. 2.2 shows the conduction and
valence band dispersions near the zone center. Since the valence subbands
have negative curvature, it is more convenient to refer to holes (the absence
of an electron) with positive mass. The broadest hole subband has J = 3/2
and J
z
= 3/2, and is known as the heavy hole band, because of its larger
eective mass. The subband with J = 3/2 and J
z
= 1/2 is referred to
as the light hole band, and is degenerate with the heavy hole band at the
zone center. Finally, the subband with J = 1/2 and J
z
= 1/2 is known
as the split-o hole band.
A spin in a magnetic eld, B has a contribution to its energy from the
Zeeman Hamiltonian:
H
Z
=
g
B
h
B

S, (2.3)
where
B
= 9.274 10
24
J/T is the Bohr magneton. For an electron in
vacuum, the electron g-factor is approximately 2. However, the presence
of the spin-orbit interaction modies this quantity. Even though, there is
no orbital angular momentum in the conduction band to rst order, the
k p matrix elements between the conduction and valence bands shift the
eective g-factor of electrons in the conduction band to values less than 2.
17
The g-factor can be calculated to be
g = g
0

2
3
E
p

E
g
(E
g
+ )
, (2.4)
where g
0
is the bare electron g-factor, E
g
is the bandgap, is the spin-
orbit energy, and E
p
is a parameter specifying the strength of the interband
coupling. For example, electrons in the conduction band of GaAs have
g = 0.44. The g-factor of holes is further complicated by the orbital
angular momentum in the valence band. However, the results below are
concerned with conduction band spins, so hole spin will be disregarded for
the most part.
The result of the Zeeman eect on an electron spin is clearly seen by
choosing the z-axis to be along the magnetic eld. Now
H
Z
=
1
2
g
B
B
z
_
_
_
_
_
1 0
0 1
_
_
_
_
_
. (2.5)
The spin eigenstates [ ) and [ ) are split by the Zeeman energy E =

1
2
g
B
B
z
. If the spin is not in an eigenstate, then it evolves in time,
depending on the Zeeman splitting. For a spin in the state given by Eq. 2.2
at t = 0, the state evolves according to (again, ignoring the overall phase)
[(t)) = cos

2
[) +e
i(
L
t+)
sin

2
[), (2.6)
where h
L
= g
B
B
z
is the Zeeman splitting.
L
is known as the Larmor
frequency. In the Bloch sphere picture, this corresponds to the spin vector
18
precessing about the z-axis at the Larmor frequency:

S = (cos(
L
t +) sin , sin(
L
t +) sin , cos ) . (2.7)
This phenomenon is referred to as Larmor precession.
For a spin perfectly isolated from the environment, a spin in a static
magnetic eld would obey the dynamics of Eq. 2.6 forever. In reality,
there are a number of eects that limit the lifetime of an electron spin in
a semiconductor. These eects can be divided into two categories: those
that aect , and those that aect in Eq. 2.6. The randomization of
is referred to longitudinal spin relaxation, and is characterized by a time
T
1
. The loss of the relative phase information is referred to as transverse
spin decoherence, occurring in time T
2
. The results described below focus
on measurements of spin coherence, so we will primarily be concerned with
mechanisms aecting T
2
. Such mechanisms can often be described as an
eective magnetic eld that uctuates randomly in time. This leads to
random uctuations in the precession axis, and precession frequency. Over
time, these uctuations induce a random walk of the spin away from its
unperturbed state, as shown schematically in Fig. 2.3. The result of these
eects is typically a reduction in the spin polarization over time, with the
form exp(t/T
2
).
One source for this eective eld is spin-orbit coupling, with Hamil-
19
time
eff
B
r
S
r
Figure 2.3: Spin decoherence: a random walk. A spin

S in an initial state (represented
by the leftmost blue arrow) is inuenced by a randomly uctuating eective magnetic
eld,

B
eff
(black arrows). As the spin precesses around the random eective eld, it
undergoes a random walk away from its initial state.
tonian H
SO
= h/(4m
2
0
c
2
)(k V ) S, were V is (for the case of bulk
semiconductors) the potential seen by the electron from the lattice ions.
By comparison with the Zeeman Hamiltonian, it is clear that this can
be seen as a k-dependent eective magnetic eld acting on the spin. In
most semiconductors, the momentum scattering time for electrons is very
short ( 10 fs). That is, the electrons quasimomentum, k is randomly
changed every 10 fs, and the spin feels an eective eld uctuating on this
timescale. This spin decoherence mechanism is known as the Dyakonov-
Perel mechanism. In the valence band, the spin-orbit coupling is strong,
and therefore the hole spin lifetime is often very short (on the same or-
der as the momentum scattering time). This short lifetime means that
20
in many cases, the hole spins can be ignored. In the conduction band,
spin-orbit coupling is only present due to the k p terms coupling the con-
duction and valence bands. Therefore, conduction band spin lifetimes can
be signicantly longer (exceeding 100 ns in bulk n-type GaAs [10]).
Additionally, spin-orbit coupling can cause spin relaxation or decoher-
ence in the conduction band more directly. Even though the spin-orbit
coupling in the conduction band is small, just as in the valence band, S
and S
z
are no longer good quantum numbers. That is to say, [) and [)
are not eigenstates there is a little mixing due to spin-orbit coupling.
This mixing can often be ignored, but it means that interactions such as
between the electron and phonons can induce spin-ips. This is known as
the Elliot-Yafet mechanism.
Another source of a randomly uctuating eective eld is the spin of
the atomic nuclei in the semiconductor. As is discussed below, the hyper-
ne interaction governing the electron-nuclear spin interaction also has the
form of an eective magnetic eld, seen by the electron spin. This inter-
action is proportional to the electron wavefunction squared at the position
of the nucleus typically each electron interacts with a number of nuclei.
If the electron has signicant overlap with N randomly oriented nuclear
spins, then there will be a randomly oriented net nuclear spin polariza-
21
tion proportional to

N. Due to dipole-dipole interactions, nuclear spins


uctuate with timescales on the order of hundreds of microseconds [11].
This timescale is typically longer than the spin decoherence time caused
by other (e.g. spin-orbit-related) mechanisms. Therefore, the nuclear spin
polarization is essentially constant within the lifetime of an individual spin,
and the precession axis and frequency are modied accordingly. However,
in most measurements, the spin is reinitialized and measured repeatedly,
or equivalently, an ensemble of spins is measured. In this case, each spin
experiences a dierent nuclear spin polarization, and the cumulative eect
is a reduced observed spin lifetime. Unlike the decoherence mechanisms
discussed above, this nuclear dephasing mechanism is not predicted to re-
sult in a simple exponential decay, but instead has Gaussian and power-law
terms [12, 13]. This type of mechanism is referred to dephasing, as opposed
to decoherence, since the coherence of individual spins is not aected. The
timescale for decoherence plus dephasing is often referred to as the inho-
mogeneous or eective transverse spin lifetime, T

2
. In principle, the eects
of dephasing can be circumvented, as will be mentioned again in Chapter
7.
As mentioned in the preceding paragraph, electron spins arent the only
spins present in a semiconductor. Many nuclear species present in common
22
semiconductors have non-zero nuclear spin. The contact hyperne Hamil-
tonian governs the electron-nuclear spin interaction: H = A

S, where I
is the spin operator for the nuclear spin, analogous to the electron spin op-
erator

S, though not necessarily spin 1/2. The constant A is proportional
to the overlap integral of the electron and nuclear wavefunction.
An important aspect of the hyperne interaction is that it allows elec-
tron and nuclear spins to ip each other. This can be seen by rewriting
the Hamiltonian in terms of raising and lowering operators: H = A

S =
A[

I
z

S
z
+(

I
+

S

+

I


S
+
)/2] where

I

=

I
x
i

I
y
and

S

=

S
x
i

S
y
. Here,

,

I

act as raising/lowering operators in that they change the quantum


number I
z
or S
z
by 1. For example, if the electron spin is in a state
with S
z
= 1/2 and the nuclear spin is in a state with S
z
= 3/2, the term
with the raising/lowering operators would yield a state with S
z
= 1/2 and
I
z
= 1/2. In this way, electron and nuclear spins ip each other.
In the experiments described below, a non-equilibrium electron spin po-
larization will be generated in a semiconductor structure. As a result, this
leads to a net polarization of the nuclear spin parallel to the electron spin
polarization. The component of the nuclear spin perpendicular to any ap-
plied magnetic eld will precess, similar to the electron spins, and therefore
there will be no steady-state nuclear spin polarization. However, if there is
23
a component of the spin along the magnetic eld, than this non-precessing
component can build up over time. Since the longitudinal relaxation time
for nuclear spins can be quite long, this steady state nuclear polarization
can take seconds or minutes to reach equilibrium. This phenomenon is
known as dynamic nuclear polarization (DNP).
As mentioned above, it is also evident from the hyperne Hamiltonian
that the nuclear spin can be thought of as an eective eld acting on the
electron spins. If the nuclear spins are thought of in a mean eld picture,
then the eect of the nuclei on an electron spin can be expressed as an
eective magnetic eld
B
n
=
h
g
B
A

I, (2.8)
where

I is the mean nuclear spin polarization, and A is appropriately
weighted to reect the strength of the interaction. This eective eld can
be quite large. In GaAs, the eective eld for 100% nuclear spin polar-
ization has been estimated to be B
n
6 T [11]. This eect can readily
be observed in a number of measurements, such as time-resolved Faraday
rotation (Chapter 6) or Hanle measurements [14].
24
2.2 Quantum dots
When electrons are conned within a semiconductor structure with one or
more dimensions smaller than the extent of the bulk electron wavefunction,
the electronic properties are drastically modied. For connement in one,
two, or three dimensions, this type of structure is known as a quantum
well, quantum wire, or quantum dot respectively. Like a particle-in-a-box,
the energy levels of electrons in a quantum dot (QD) are quantized into
discrete levels, as illustrated in Fig. 2.4. Instead of a continuous band of
conduction band states, the energy eigenstates are now spatially localized
within the QD, and separated by an energy that increases with increasing
connement. The lowest conduction band state is blue-shifted from the
conduction band minimum due to this connement.
When the temperature is low enough such that k
B
T < E, where E
is the QD energy level spacing, then the quantized nature of the energy
levels becomes apparent. For temperatures around 4 K, this requires a QD
size on the order of 100 nm. In this regime, QDs will exhibit an atom-like
spectrum of absorption and emission lines.
The spin physics are also altered by quantum connement. The blue-
shift of the conduction band ground state eectively increases the gap
between the conduction and valence band states, reducing the eect of the
25
conduction band
valence band
E
QD
Figure 2.4: Schematic of conning potential, energy levels, and wavefunctions in a
quantum dot. A quantum dot is formed in a semiconductor at a local energy mini-
mum. Quantized energy levels and particle-in-a-box wavefunctions are illustrated. The
conning potential exists in three dimensions, though only one dimension is shown.
k p matrix elements that couple the two bands. This primarily has the
eect of shifting the electron g-factor towards the bare electron g-factor,
g = 2. Modifying Eq. 2.4 to account for this energy shift, E
0
,
g = g
0

2
3
E
p

(E
g
+E
0
)(E
g
+E
0
+ )
. (2.9)
Additionally, momentum scattering is now suppressed and along with it,
the various spin-orbit-related decoherence mechanisms that are caused by
26
momentum scattering. However, since the electron wavefunction is now
more spatially concentrated, it interacts more strongly with the nuclear
spins, making nuclear spin dephasing much more signicant.
There are a number of ways of physically realizing semiconductor QDs.
These fall into three categories: 1. Chemically synthesized QDs, 2. QDs
grown by molecular beam epitaxy (MBE), and 3. QDs dened by selec-
tively depleting a two-dimensional electron gas (2DEG). Figure 2.5 shows
examples of each type.
Chemically synthesized QDs can by made in a number of ways, though
primarily through colloidal chemistry methods. In this technique, a ask
containing a solution of, for example, Cd and Se ions is heated and cooled in
a specically controlled protocol to nucleate CdSe nanocrystals and then
grow them to the desired size. CdSe is one common choice of material,
though many materials have been used with this method including II-VI,
III-V, and oxide semiconductors. The result of this process is a large
number of nanocrystal quantum dots in solution with a diameter tunable
from about 1 - 10 nm (Fig. 2.5 a).
One drawback to chemically synthesized QDs is that they live in a hard-
to-control chemical environment. (On the other hand, this could be seen
as a bonus the chemical environment leads to interesting phenomena and
27
5 nm
(a)
InAs
GaAs
(b)
(c)
Figure 2.5: Types of quantum dots. (a) High-resolution transmission electron micro-
graph of a CdSe nanocrystal QD, showing the regular lattice of Cd and Se atoms (X.
Peng et al., J. Am. Chem. Soc. 119 2007). (b) Cross-sectional transmission electron
micrograph (55 55 nm) of a vertical stack of self-assembled InAs QDs (P. M. Koen-
raad et al., Physica E 17 2003). (c) Electron micrograph of a gate-dened 2DEG QD
structure, dashed circle is the QD region (J. M. Elzerman et al., Nature 430 2004).
is another knob to turn.) Nevertheless, it is often nice to simplify things a
bit, and better isolate the system being studied. This can be achieved by
fabricating QDs that are embedded in an epitaxial semiconductor mate-
rial. One method is to fabricate a quantum well out of GaAs, with AlGaAs
barriers. Under appropriate growth conditions, the GaAs/AlGaAs inter-
faces will have roughness of 1 atomic layer, with lateral length scales on
the order of 100 nm. These uctuations in the thickness of the QW form
28
localized potential minima that act as QDs. Another type of MBE grown
QD is grown using strain-induced self-assembly. If InAs is deposited on a
GaAs surface under the right conditions, the InAs will not form a smooth
layer, but instead will bead up into droplets with dimensions of tens of
nanometers. After these droplets have formed, GaAs is grown over the top
of the InAs (Fig. 2.5 b). Since the bandgap of InAs is lower than that of
GaAs, these droplets then serve as QDs.
Finally, a 2DEG (essentially an electron-doped QW with high electron
mobility) can be grown via MBE. The two-dimensional electrons in this
structure can then be further conned by applying a bias to electrical
gates deposited on the surface of the sample. These gates deplete the
nearby electrons in the 2DEG, creating a potential minimum surrounded
by depleted regions (Fig. 2.5 c). This forms a QD well-suited for electrical
contacting a small gap in the gates creates a tunable tunneling barrier
to an electrical lead.
2.3 Optical properties of spins in
semiconductors
The connection between spins in semiconductors and light is interesting for
practical applications, experimental techniques, and new physical phenom-
29
ena: An example of an application might be the hybrid solid state/optical
platform for quantum information processing described above; Several op-
tical experimental techniques for probing spins in semiconductors will be
described below; and an example of an interesting physical phenomenon
would be the o-resonant interaction between a single spin and light de-
scribed in Chapter 7.
Abraham Lincoln told a story about an Eastern monarch [who] once
charged his wise men to invent him a sentence, to be ever in view, and which
should be true and appropriate in all times and situations. If they were
truly wise, they would have said, Write down the Schrodinger equation,
and solve it. For an electron interacting with an electromagnetic eld,
the Schrodinger equation for the electron wavefunction is
_
1
2m
_
p +
e
c
A
_
2
+V (r)
_
= E (2.10)
where A is the vector potential of the electromagnetic eld. For a plane
wave with frequency and polarization along , we can take
A = (c/)E
0
exp(i(t k r)) +c.c.. (2.11)
Treating the A-dependent terms as a perturbation and making the usual
dipole approximation that k r is small, then to leading order we have a
30
time-dependent perturbation
V

=
eE
0
p
m
_
e
it
+e
it
_
. (2.12)
Using time-dependent perturbation theory, it is not too hard to show
that this perturbation induces transitions from an initial state [i) to a nal
state [f) with probability proportional to
_
eE
0
m
_
2
[f[ p[i)[
2
. (2.13)
For electrons in a semiconductor, the initial states [i) are states in any
of the valence subbands, and the nal states [f) are in the conduction
band. Because of the symmetry of the momentum operator in Eq. 2.13,
there are certain selection rules for these interband transitions. Namely,
for circularly polarized light, = ( x i y)/

2, the momentum operator in


the matrix element in Eq. 2.13 transforms like a spherical tensor of rank 1,
with q = 1. This means that these matrix elements are non-zero only for
initial and nal states with angular momentum projection L
z
diering by
1. Here, the z-axis is dened by the propagation direction of the light.
This can also be seen in terms of photons: the absorption of a circularly
polarized photon transfers its angular momentum L
z
= h to the spin.
Looking at Fig. 2.2, it is clear that these selection rules determine
which transitions can take place between various valence subbands, and
31
S=1/2:
J=3/2:
J=1/2: J
z
=-1/2 J
z
=+1/2
J
z
=-3/2 J
z
=-1/2 J
z
=+1/2 J
z
=+3/2
S
z
=+1/2 S
z
=-1/2
3
1
2
2
1
3
Figure 2.6: Optically allowed transitions from the valence band (heavy and light holes
with J = 3/2 and split-o holes with J = 1/2) to the conduction band (with S = 1/2).
Red and blue indicate left and right circularly polarized light, and the thickness of the
arrow indicates the strength of the transition.
the conduction band. Specically, the allowed transitions from the heavy
hole, light hole, and split-o hole bands are shown in Fig. 2.6 for right and
left circularly polarized light.
If the matrix elements are actually calculated [9], it is found that the
dierent transitions shown in Fig. 2.6 have dierent matrix elements. That
is, some transitions are more likely than others, for the same light intensity.
These relative transition strengths are also shown in the gure: the heavy
hole transition is three times as strong as the light hole transition, and the
split-o hole transition is twice as strong as the light hole transition. This
provides a useful way to generate a net spin polarization in the conduction
band of a semiconductor. If light is incident on the semiconductor at an
32
energy E, such that E
g
< E < + E
g
, then only transitions from the
heavy hole or light hole states can be excited. If this light is circularly
polarized, then only transitions from the, say, J
z
= +3/2 heavy hole and
J
z
= +1/2 light hole are allowed. Since the heavy hole transition is three
times as likely as the light hole transition, the result is three times as many
conduction band spins with S
z
= +1/2 than with S
z
= 1/2. This results
in a net spin polarization in the conduction band of 50%.
The situation can be improved further if the semiconductor is strained,
or if the electrons and holes are conned in one or more dimensions. In
this case, the degeneracy of the heavy and light hole bands is lifted, and
transitions can be pumped from the heavy hole band only, resulting in
ideally 100% conduction band spin polarization.
The inverse of this optical spin injection provides a means for detecting
the spin polarization of electrons in a semiconductor. When an electron
and hole recombine, light is emitted with circular polarization reecting
the spin state of the electron and hole. Thus by measuring the degree
of circular polarization of this luminescence, one can measure the spin
polarization of the electrons and holes.
Using the polarization of the luminescence as a probe of the spin polar-
ization, one can get some information about the spin dynamics in a mag-
33
laser
polarizers
/4-plates
circ. pol.
P
L
To
spectrometer
sample
B
Figure 2.7: Schematic of a Hanle measurement setup. Above-bandgap laser light is
circularly polarized using a quarter-wave plate. This light excites spin polarized elec-
trons into the conduction band of the sample. The PL is collected, and its polarization
is analyzed using a quarter waveplate and a polarizer. The angle between injection and
detection may be made smaller than shown here.
netic eld. This type of experiment is known as a Hanle measurement, and
the setup is shown schematically in Fig. 2.7. A circularly polarized laser is
incident on the sample in the +x-direction, serving to inject spin-polarized
electrons and holes. A magnetic eld B
z
is applied in the z-direction.
The polarization of the subsequent photoluminescence (PL) collected back
along the x-direction reveals the steady-state spin polarization along the
measurement direction.
This steady-state spin polarization can be calculated by taking an ini-
tial electron spin polarization S
0
at t = 0 along the +x-direction. (For
simplicity, we will assume that the hole spin lifetime is very short and
can be ignored.) The spin then precesses as given by Eq. 2.6, and the x
34
increasing magnetic field, B
z
S
r
x
S
x
y
z
Figure 2.8: Illustration of the Hanle eect. At zero magnetic eld, the spin decays
without precessing (left). As the magnetic eld increases, the spin precesses through
a larger angle before decaying. As the maximum precession angle increases, the time-
averaged spin polarization decreases, as indicated by the red arrows.
component of the spin as a function of time is given by
S
x
(t) =
_

_
0 t < 0
S
0
cos(
L
t) exp(t/T

2
) t 0
. (2.14)
As before,
L
= g
B
B/ h is the Larmor precession frequency, and for gener-
ality both spin decoherence and dephasing are included in the spin lifetime
T

2
. The steady state spin polarization S
x
is then found by integrating
Eq. 2.14 from t = (, ):
S
x

_

0
S
0
cos(
L
t) exp(t/T

2
) dt = S
0
_
1/T

2
L
+ (1/T

2
)
2
_
. (2.15)
Thus as a function of magnetic eld (proportional to
L
), the mea-
sured PL polarization sweeps out a Lorentzian function with width B
1/2
=
h/(g
B
T

2
). This eect is illustrated in Fig. 2.8. If the g-factor is known,
35
then this measurement reveals the eective transverse spin lifetime, T

2
.
The analysis of the Hanle measurement can be done more rigorously and
quantitatively by setting up and solving rate equations for spin injection,
decay, and recombination [14]. The theory can be extended to the case
of nuclear polarization, non-exponential decay, doped semiconductors, etc.
This was the technique used for much of the initial exploration of spin
physics in semiconductors. However, there is only so much information
that can be extracted from a Hanle-type measurement. In contrast, the
Faraday rotation measurements described in the next section provide a
more powerful probe of semiconductor spin dynamics.
2.4 Faraday rotation: theory
The Faraday eect provides a useful optical probe of spin polarization in
semiconductors. This eect can be observed as a rotation of the plane
of polarization of linearly polarized light as it is transmitted through a
material (Fig. 2.9). The angle of this rotation is dependent on the spin
polarization in the material. The Kerr eect is directly analogous, but
refers to reection of the light o of the sample, instead of transmission.
The eect arises from the spin-dependence of the index of refraction for
right and left circularly polarized light (
+
or

polarization). This spin-


36
S
r
S
x
x F
S
Figure 2.9: Diagram of the Faraday eect. As linearly polarized light is transmitted
through the sample, its plane of polarization is rotated through an angle,
F
, propor-
tional to the projection of the spin polarization in the material along the light propa-
gation direction.
dependent change in the index of refraction results in a dierent phase shift
in the light for
+
or

polarized light transmitted through the sample.


Following Eq. 2.11, the vector potential for

polarized light propagating


in the z-direction is given by
A

= A
0
(x iy) exp[i(t k r
F
)] +c.c., (2.16)
where
F
is the phase shift caused by the Faraday eect, and k = [k[z.
Linearly polarized light can be written as the sum of
+
and

polariza-
tions: (x + iy) + (x iy) = 2x. After transmission through the sample,
linearly polarized light is given by the vector potential
A
+
+A

= 2A
0
(cos
F
x sin
F
y) exp[i(t k r)] +c.c.. (2.17)
This is the vector potential for light propagating in the z-direction with
37
linear polarization rotated through an angle
F
with respect to the x-axis.
There are several ways to see where this phase shift comes from. Here,
I will sketch a quantum mechanical derivation with the radiation treated
classically. Appendix C outlines two other derivations: a classical view,
and a fully quantized picture.
Following Ref. [15], we start with the Schrodinger equation for an elec-
tron in an electromagnetic eld, given in Eq. 2.10. Again, using the dipole
approximation and time-dependent perturbation theory, we can calculate
the dielectric function (E) for the electron, in terms of the available initial
and nal states. The imaginary part of

yields the absorption spectrum,


and the real part the index of refraction as a function of energy. For a
single set of initial and nal states, the index of refraction for

polarized
light is found to be
n

[c[ p
x
i p
y
[v)[
2
E E
c,v
(E E
c,v
)
2
+
2
c,v
, (2.18)
where [c) and [v) are conduction and valence band states, and E
c,v
and

c,v
are the energy and linewidth of the transition from [v) to [c).
Since the velocity of light in a material is given by c/n, the phase shift
for circularly polarized light caused by transmission through a material
with index n is = Ln/c, where L is the path length through the
38
material. From this we have the Faraday rotation angle

F
(E) =
LE
hc
(n
+
n

)
= CE

c,v
_
[c[ p
+
[v)[
2
[c[ p

[v)[
2
_
E E
c,v
(E E
c,v
)
2
+
2
c,v
,
(2.19)
where p

= p
x
i p
y
, C is a constant that depends on the geometry of
the experiment, and we have summed over all valence band to conduction
band transitions.
Note that if the momentum matrix elements are equal for right and
left circular polarization, then the Faraday rotation angle is zero. This is
where the spin dependence comes in. If there is an electron in the conduc-
tion band in a particular spin state, then the transition to this state will be
forbidden by the Pauli exclusion principle. As discussed above, the momen-
tum matrix elements for a given circular polarization couple more strongly
to one conduction band spin state than the other (Fig. 2.6). Therefore,
the magnitude of these matrix elements for right and left circular polariza-
tion depend on the spin polarization in the conduction band. Figure 2.10
shows the predicted Faraday rotation spectrum from Eq. 2.19 for a single
interband transition, and integrated over a typical density of states with
a spin-polarized electron population given by the Fermi distribution near
the conduction band edge.
It should be noted that Faraday rotation can also arise when the am-
39






FR
DOS
Energy (a.u.)
F
a
r
a
d
a
y

r
o
t
a
t
i
o
n

(
a
.
u
.
)
Filled spin-
up states
(a)
(b)
Figure 2.10: Theoretical Faraday rotation spectra. (a) Faraday rotation spectrum for
a single interband transition. (b) The Faraday rotation spectrum integrated over an
ensemble of states. The dashed red line is (qualitatively) the density of states (DOS)
near the absorption edge, and the blue shaded region indicates the occupied electron
states for a given Fermi energy and temperature.
plitude of the matrix elements are equal, but the energy of the transitions
are spin-split (e.g. by a magnetic eld). In the measurements described
here, the magnetic eld is suciently small that this eect is negligible.
40
2.5 Faraday rotation: experiment
Since the initial interest in Faraday rotation in semiconductors in the
1960s, a number of experimental techniques have been developed for this
type of measurement. A typical scheme is to use a pump laser to optically
inject spins into the conduction band, and a probe laser to measure Fara-
day rotation (or Kerr rotation). In this type of pump/probe spectroscopy,
the two lasers may have the same or dierent energies.
If the pump and probe lasers are continuous wave (cw) then this pro-
vides information about the steady-state spin polarization, similar to the
Hanle measurement described above. However, Faraday rotation measure-
ments can be extended into the time domain to provide a more direct look
at the spin dynamics.
In a time-resolved Faraday rotation experiment, pulsed pump and probe
lasers are used, shown schematically in Fig. 2.11. Often, the pump and
probe lasers are derived from a single mode-locked Titanium:Sapphire
(Ti:Sa) laser. Such a laser typically outputs an optical pulse every 13 ns
with duration 150 fs. The pump pulse is circularly polarized, and excites
spin polarized electrons into the conduction band, as discussed above. The
probe pulse can be made to arrive at the sample with a xed time delay,
t, by controlling the relative path length traveled by the two pulses. This
41
Ti:Sapphire
t
Circ.
pol
lin.
pol
pump
probe
~150 fs
sample
to diode
bridge
(not to scale)
Figure 2.11: Typical time-resolved Faraday rotation setup. A mode-locked
Ti:Sapphire laser produces a train of pulses with 150 fs duration. The pulse train
is split into a circularly polarized pump and linearly polarized probe. The delay, t,
between pulses is controlled with a delay line in one path. The pump excites spins in
the sample, and the probe measures the spin a time t later, via the Faraday eect.
can be achieved by reecting one of the pulses o of a delay line that
is, a translatable mirror (or retroreector). Given the speed of light in air
(about 1 ft./ns), changing the mirror position by one foot changes t by
about 2 ns. After transmission through (or reection o of) the sample,
the Faraday (or Kerr) rotation of the linear polarization of the probe pulse
is measured. This provides a snapshot of the spin polarization precisely t
after the arrival of the pump pulse. By varying the delay, t, the evolution
of the spin dynamics can be mapped out in time.
The polarization rotation of the probe beam is detected with a balanced
photodiode bridge. As shown in Fig. 2.12, the polarization of the probe
42

F
/2 plate
22.5
45-
F
Glan-thomson
(polarizing)
beamsplitter
I
0
I
0 ( ) 2 1
0 F
I +
( ) 2 1
0 F
I
photodiodes:
A B
-
bridge
circuit
F B A out
I I V
Figure 2.12: Sensitive polarization detection with a balanced photodiode bridge. In-
coming light (from the right) is linearly polarized at a small angle,
F
to the vertical.
A /2 plate reects the polarization about its axis at 22.5

, vertical and horizontal


components are split, and the dierence is measured using the diode bridge circuit.
beam, initially vertical, is rotated through an angle
F
. The probe beam is
then passed through a half-wave plate which reects the polarization about
its axis at 22.5

, resulting in the polarization at an angle = /4


F
.
Next, the probe is passed through a polarizing beamsplitter, which sepa-
rates the horizontally and vertically polarized components of the light. As-
suming
F
is small (typically on the order of milliradians or less), then the
horizontal and vertical components have intensity I
H(V )
I
0
(1
F
)/

2
where the incoming probe beam has intensity I
0
. These two beams are
then focused onto a pair of photodiodes, and the dierence between the
two photocurrents is measured by the diode bridge circuit. (For a detailed
discussion of diode bridge circuits, see Katos thesis [16]). This dierence
43
signal S
HV
I
H
I
V
=

2I
0

F
, and is thus proportional to the Faraday
rotation angle.
Since the Faraday rotation angles are typically small, the pump and
probe beams are often modulated at frequencies ranging from Hz to kHz
and lock-in detection is used to isolate the desired signal and to reduce
noise (specic details will be given for the particular experiments described
below). The bandwidth of the diode bridge must exceed these modulation
frequencies, though for noise reduction purposes it is desirable to keep
the bandwidth as low as possible. In particular, the bandwidth will be
much lower than the bandwidth of the probe pulses (which have bandwidth
as high as 1/100fs = 10 THz). This is the great advantage of this sort
of stroboscopic technique it allows measurements with very high time
resolution with low bandwidth detection.
The experiments described below all make use of some variation on the
theme of time-resolved Faraday rotation. More experimental details will
be given throughout.
44
Chapter 3
Quantum Shells in
Semiconductor Nanocrystals
3.1 Motivation
In the biochemical process used by magnetotactic bacteria to build their
compass (see Chapter 1), magnetic particles are nucleated and grown from
metallic precursor chemicals [2]. Once the particle has reached the desired
size (on the order of 10s of nanometers) the growth is arrested. This
ancient, natural process is remarkably similar to the methods recently
developed by human scientists to grow both magnetic and non-magnetic
nanocrystals through colloidal chemistry.
Here we are focusing on semiconductor nanocrystals which act as quan-
tum dots (QDs), conning electrons on nanometer length scales. Nanocrys-
45
ZnS
CdSe
(a)
V
r
e
h
V
r
e
h
CdS
CdSe
(b)
Figure 3.1: Cutaway schematic of layered nanocrystals. (a) A low-bandgap core sur-
rounded by a high-bandgap shell connes electrons and holes to the interior, protecting
them from the environment. (b) A high-bandgap core, a low-bandgap shell, surrounded
by a high-bandgap cap connes electrons and holes to the spherical shell layer. The di-
agrams at the right illustrate the conduction and valence band proles on a nanocrystal
in the radial direction, with schematic electron and hole wavefunctions.
tal QDs are interesting systems for a number of reasons. First, the strong
connement of the electrons and holes means that the connement energy
exceeds k
B
T even at room temperature. Additionally, the simple chemical
process used to fabricate these nanocrystals (discussed below) allows great
exibility in choice of materials, the chemical environment of the QDs, the
size, and even the shape of the nanocrystals.
Furthermore, not only can nanocrystal QDs be made of a variety of
46
semiconducting materials, but they can have multiple layers of dierent
materials within a single nanocrystal. By taking advantage of the dif-
ferent bandgaps of dierent semiconductors, this allows one to tailor the
conning potential for the electron in an individual nanostructure. For
example, in a core-shell structure with a low bandgap core, surrounded
by a larger bandgap shell, the electron will sit mainly in the core and
thus be protected from the environment. Perhaps more interestingly, a
high-bandgap core can be surrounded by a low-bandgap shell, and then a
second high-bandgap shell. Such structures, referred to as quantum-dot
quantum wells (QDQWs), are essentially a quantum well in a spherical ge-
ometry, within a single nanocrystal. A cutaway schematic of such a struc-
ture is shown in Fig. 3.1. Both core-shell quantum dots [17, 18, 19] and
QDQWs [20, 21, 22, 23, 24, 25] have been synthesized during the past years.
Although both CdS/HgS/CdS [21, 26, 22, 27, 28] and CdS/CdSe/CdS [25]
QDQWs were well characterized by photoluminescence (PL) and absorp-
tion spectroscopy, a detailed investigation of the quantum size levels was
not initially possible, mainly because the PL peak is spectroscopically
broad and the absorption spectrum shows a featureless increase. This
is due to the distribution of size and shape among the ensemble of QDs,
and therefore the distribution of QD connement energies. In the work
47
described below, the spin dynamics in these systems is addressed using the
Faraday eect, which also provides a more sensitive measurement of the
energy level structure.
The following sections describe time-resolved Faraday rotation (TRFR)
for CdS/CdSe/CdS QDQWs with varying CdSe quantum well width (n
CdSe
=
1 5 monolayers). The spin lifetime is of order 2 3 ns even at room tem-
perature. The QDQWs exhibit g-factors that vary with quantum well
width. TRFR is not only a unique experimental probe for the spin dynam-
ics, but also a sensitive spectroscopic technique. In contrast to absorption
spectra, the amplitude of the TRFR signal as a function of probe energy
exhibits three distinct resonances close to the absorption edge, because
optical transitions to the lowest conduction band level are probed selec-
tively. From the level scheme and dielectric response functions evaluated
with k p calculations [27, 29], we show that, while several peaks are ex-
pected to emerge in the TRFR signal, the large number of resonances with
comparable spectral weight is not reproduced by the quantum size levels
of a spherical QDQW. One possible mechanism that explains the exper-
imental data is broken spherical symmetry, which mixes dierent valence
band multiplets.
48
3.2 Sample structure and characterization
Colloidal QDQWs with varying width of the CdSe quantum well were syn-
thesized by a successive ion layer adsorption and reaction (SILAR) tech-
nique to produce nanocrystals with accurate control over the quantum well
width [19, 25]. First, a CdS (E
g
= 2.48 eV) core is grown using standard
colloidal chemistry. Above a critical temperature (T > 350

C) nanocrys-
tals are nucleated in a solution of Cd- and S-containing precursor molecules.
The temperature is then lowered below the nucleation temperature, and
the nanocrystals then begin to grow. The size of the nanocrystals can be
monitored in situ using the optical absorption. Once the nanocrystals have
grown to the desired size, the growth is stopped by lowering the tempera-
ture further. In this work, a 3.4 nm core diameter was used. Subsequent
layers are then grown by immersing the nanocrystals in alternating cation
or anion containing solutions, adding one monolayer at a time. In this
way, the thickness and composition of the nanocrystals can be controlled
with atomic precision. Samples were made with 1 to 5 layers of CdSe
(E
g
= 1.74), followed by a 1.6-nm thick CdS cap. A schematic represen-
tation of the structure is shown in Fig. 3.1, along with a diagram of the
conduction and valence band proles in the radial direction. Electrons and
holes will be conned to the CdSe shell, as indicated in the gure. The
49




Wavelength (nm)
P
L

I
n
t
e
n
s
i
t
y

(
n
o
r
m
a
l
i
z
e
d
)
1 2 3 4 5 CdSe layers
500 600 700
0
(a)
(b)
Figure 3.2: Photoluminescence of quantum shell samples. (a) Photograph of single
shell samples under UV illumination with 1 (right) to 5 (left) CdSe layers. (b) PL
spectra of the same samples. The spectra have been normalized for clarity.
QDQWs were dissolved in toluene and all measurements were carried out
on a large ensemble of QDQWs in solution at room temperature unless
otherwise specied.
A photograph of the samples used in these experiments is shown in
Fig. 3.2, illuminated by a handheld UV lamp. As the shell thickness is
50
Ti:Sapphire
t
Circ.
Pol
(PEM)
lin.
pol
pump
~150 fs
sample
to diode
bridge
(not to scale)
Regenerative
amplifier
~150 fs
OPA2 OPA1
chopper
permanent
magnets
monochromator
Figure 3.3: Setup for quantum shell TRFR. Two optical parametric ampliers (OPAs)
provide synchronized, independently tunable pump and probe pulses. The pump has its
circular polarization modulated with a photo-elastic modulator (PEM), and the linearly
polarized probe is modulated with a chopper. The time delay between arrival of the
pump and probe at the sample is controlled by a delay line, and the Faraday rotation
of the probe is measured by a photodiode bridge. A monochromator can be inserted in
the probe path to enhance the spectral resolution.
increased, a redshift of the luminescence can be clearly seen. The corre-
sponding PL spectra are shown in the gure.
3.3 Experimental setup
The experimental setup is shown in Fig. 3.3. A regeneratively amplied
Ti:Sapphire laser (Coherent RegA) outputting 200 fs duration pulses at
51
a repetition rate of 250 kHz was used to drive two optical parametric am-
pliers (OPAs). These OPAs, outputting equally short pulses with wave-
length independently tunable from about 500-700 nm, are used as pump
and probe lasers. The pump and probe pulses were both focused to an
overlapped spot within the QDQW solution. The wavelength of the pump
laser was xed at
pump
= 505 nm (E
pump
= 2.46 eV). This pulse train was
circularly polarized, exciting spin polarized electrons into the conduction
band states of the QDQWs. Relaxation of the electron and hole to the
lowest exciton state presumably occurs on a picosecond time-scale, as in
similar systems such as CdS/HgS/CdS QDQWs [30]. The linearly polar-
ized probe pulse then passes through the QDQW solution a time t later,
where t is set using a mechanical delay line in the pump beam path. The
Faraday eect causes the polarization of the probe pulse to be rotated by
an angle,
F
, proportional to the component of the net spin polarization
along the probe beam direction (as described in Section 2.4). By recording

F
for varying t, we detect the time evolution of the optically injected
electron spins in the QDQWs.
52
t (ns)


0 1 2 3

F
(
a
.
u
.
)



F
T

p
o
w
e
r
(
a
.
u
.
)
0 5 10
Frequency (GHz)
(a)
B
app
=3 kG
(b)
Figure 3.4: Spin precession in quantum shells. (a) Faraday rotation as a function of
pump-probe delay, t, in a 3 kG magnetic eld. The red arrow indicates the amplitude
of the FR, as plotted in the FR spectra below. (b) Fourier transform of the data in (a).
The apparent rise at short delay is actually due to the second frequency component,
visible as the high frequency shoulder in the FT.
3.4 Spin dynamics in quantum shells
Two permanent magnets with adjustable separation were used to apply a
magnetic eld, B
app
, to the sample perpendicular to the pump and probe
direction. Spins that were initially polarized along the pump beam pre-
cess around the magnetic eld at the Larmor frequency,
L
= g
B
B
app
/h
where g is the electron g-factor,
B
the Bohr magneton, and h the Planck
constant. Figure 3.4 shows typical data from a sample with a quantum
well width of n
CdSe
= 3 monolayers and B
app
= 0.3 T. The inset shows
the Fourier transform (FT) power spectrum of the time-domain data. A
second precession frequency was observed, as indicated both by the small
53
shoulder in the FT spectrum and the beating in the time-resolved data.
While the origin of this second frequency is unclear in the present case,
similar behavior has been observed in CdSe nanocrystals [31, 32, 33, 34].
This second frequency may be associated with charging of defect states at
the surface of the nanocrystals. Dangling bonds at the nanocrystal surface
give rise to electron states within the bandgap. These states have been
observed to randomly charge and discharge leading to blinking of the PL
and shifts in the PL energy. Measurements of spin dynamics in electro-
chemically gated nanocrystal QDs have shown a dependence of the relative
prominence of the two precession frequencies on the QD charging [35].
There is also a non-oscillating component to the TRFR signal which
was also seen in previous measurements on CdSe nanocrystals [31]. In some
samples, particularly for n
CdSe
= 5, the magnitude of the non-oscillating
component is comparable to that of the oscillating component. However,
for the purposes of this work we focus only on the oscillating component
(indicated by the arrow in Fig. 3.4). The eective transverse spin lifetime,
T

2
, was of order 2 or 3 ns for all samples measured. Remarkably, the spin
lifetime was essentially temperature-independent between room tempera-
ture and 5 K, as shown in Fig. 3.5. (Low temperature measurements were
performed on QDQWs embedded in a polyvinyl butyral (PVB) matrix in
54



0 100 200
0
1
2
Temperature (K)
S
p
i
n

l
i
f
e
t
i
m
e
,

T
2
*
(
n
s
)
Figure 3.5: Temperature dependence of the quantum shell spin lifetime. Between 5
and 200 K, the measured spin lifetime, T

2
is essentially constant around 1.5 ns. There
might even be a slight increase in spin lifetime at higher temperature. This data is
from QDQWs in a PVB matrix, but the measured lifetime is consistent with the room
temperature spin lifetime in solution.
a magneto-optical cryostat, as further described in Chapter 4.)
We have performed TRFR measurements as a function of B
app
on sam-
ples with CdSe quantum well width from n
CdSe
= 1 5 monolayers. In
all cases, the results show either one or two precession frequencies that
increase linearly with B
app
. The component with higher frequency always
had a substantially smaller amplitude and will not be addressed further
here. The top panel of Fig. 3.6 shows the FT of the spin precession for
55
shells with n
CdSe
= 1, 3, and 5 monolayers, with a clear shift in precession
frequency with shell thickness. The inset to Fig. 3.6 shows the linearly
increasing main precession frequency as a function of B
app
for n
CdSe
= 1, 3,
and 5 monolayers. The measured g-factor for each sample is shown in the
lower panel of Fig. 3.6 (circles) in comparison with the theoretical values
(crosses) obtained from an weighted average of the CdSe and CdS g-factors
(see below). Within the experimental error, the g-factor did not show any
dependence on temperature from 5 K to room temperature or on the probe
wavelength.
3.5 Energy levels in quantum shells
In order to investigate the QDQW energy levels, we have measured the
dependence of the TRFR amplitude on probe wavelength in the samples
with n
CdSe
= 3, 4, and 5. The probe beam coming out of the OPA typically
had a full width at half maximum (FWHM) of 10 nm. A monochromator
after the sample (as shown in Fig. 3.3) was used to select a 2 nm FWHM
slice of the probe beam. The measurement was performed by setting the
probe OPA near the desired wavelength, then setting the monochroma-
tor to the desired wavelength within the OPA spectrum. A TRFR scan
was performed at this probe wavelength, and then the monochromator
56



1 3 5 CdSe layers
Frequency (GHz)
F
T

p
o
w
e
r

(
a
.
u
.
)
0 5 10
B
app
=3kG




0
0.6
1.2
1.8
1 2 3 4 5
g
-
f
a
c
t
o
r
experiment
theory
CdSe monolayers



8
0
0 4 B
app
(kG)

(
G
H
z
)
Figure 3.6: Precession frequency and g-factors as a function of shell thickness. Top:
Fourier transforms of the TRFR data for samples with shell thickness of 1, 3, and 5
CdSe monolayers, in a magnetic eld of 3 kG. Bottom: quantum shell g-factor as a
function of shell thickness. Circles are the measured values, and crosses are calculated
as discussed below. The inset shows the precession frequency as a function of magnetic
eld for 1, 3, and 5 CdSe monolayers.
wavelength was adjusted, and the next scan performed. Once the desired
wavelength was outside of the probe OPA spectrum, then the probe OPA
wavelength was adjusted. In this way, a large range of probe wavelength
57







F
a
r
a
d
a
y

r
o
t
a
t
i
o
n

a
m
p
l
i
t
u
d
e

(
a
.
u
.
)
O
p
t
i
c
a
l

a
b
s
o
r
p
t
i
o
n

(
a
.
u
.
)
550 600
650 700
Probe wavelength (nm)
5 monolayer
4 monolayer
3 monolayer
0
0
0
Figure 3.7: Quantum shell Faraday rotation spectra. Faraday rotation amplitude as
a function of probe wavelength for samples with 3, 4, and 5 CdSe monolayers. Also
shown is the optical absorption for each sample. The FR amplitude was dened as
the dierence between the local maximum and minimum closest to t = 500 ps in the
TRFR data.
was swept out.
Figure 3.7 shows the TRFR oscillation amplitude as a function of probe
wavelength for the dierent samples, together with optical absorption data.
The amplitude was determined by taking the dierence between the max-
58
imum and minimum signal for scans like the one shown in Fig. 3.4. While
the absorption signal only shows a featureless staircase-like behavior with
no distinct resonances, the amplitude of the TRFR signal exhibits several
pronounced resonances close to the absorption edge. The results in Fig. 3.7
show that TRFR does not only provide information on the spin dynam-
ics, but also is a more sensitive spectroscopic technique than absorption
spectroscopy and allows one to identify individual exciton transitions in
QDQWs.
3.6 Theoretical description
These results were modeled theoretically in Meier et al. [36]. In that work,
the conduction and valence band level scheme of spherical QDQWs is cal-
culated with k p theory [27, 29], using a two-band description for the
conduction band and the four-band Luttinger Hamiltonian in the spheri-
cal approximation for the valence band. The radial potential for electrons
and holes is determined by the oset of the CdS conduction and valence
band edge relative to CdSe, 0.32 eV and 0.42 eV, respectively [37]. The
electron and hole wavefunctions are then given by an envelope function,
which is a solution to the Schrodinger equation in the given potential,
multiplied by the underlying Bloch wavefunctions.
59
(a)
(b)
Figure 3.8: Calculated electron and hole energy levels in quantum shells. Low-lying
electron (a) and hole (b) energy levels as a function of shell thickness. Energies are
given relative to the CdSe band-gap. The unlabeled hole states are 1D
5/2
, 1P
5/2
, and
1D
7/2
from the bottom.
The calculated energies of the lowest conduction and valence band
states are shown in Fig. 3.8. Dierent valence band multiplets are de-
noted by L
F
[38, 39], where L is the smallest angular momentum of the
envelope wave function and F = [LJ[ the total angular momentum. (The
valence band eigenstates are superpositions of envelope functions with an-
gular momentum L and L + 2 [36].) Figure 3.9 shows the radial wave
function envelope of the conduction band ground state 1S
e
(solid line) and
of 1S
3/2
(dashed lines) for n
CdSe
= 3. Both wavefunctions show some degree
of localization within the spherical quantum well. Because of the larger va-
lence band mass, the valence band states are much better localized in the
quantum well. The valence band ground state, 1P
3/2
, has a p-type envelope
60
r
1
=1.7nm
r
2
-r
1
=(0.43nm)n
CdSe
r
3
-r
2
=1.6nm
1S
e
1S
3/2
Figure 3.9: Calculated electron and hole wavefunctions in a quantum shell. The radial
component of the electron (red) and hole (black) wavefunctions in the 1S
e
and 1S
3/2
states. The dashed line is the L = 0 component, and the dot-dashed line is the L = 2
component of the hole wavefunction.
wave function, which is consistent with a dark exciton ground state. This
characteristic of a p-type envelope wavefunction for the lowest energy hole
state is often found in nanocrystal QDs, and leads to very long radiative
lifetimes (> 10 ns at room temperature).
From the energy and wavefunction of the conduction band ground state
1S
e
, the electron g-factor is calculated as weighted average over the CdSe
and CdS g-factors, g = 2 2E
p

so
/3(E
p
+
so
+ E
1Se
)(E
p
+ E
1Se
) (see
Eq. 2.9). The numerical values obtained with standard parameters [40]
61
for the Kane interband energy E
p
, energy gap E
g
, and spin-orbit energy

so
are shown in Fig. 3.6 (crosses). The quantitative discrepancy to the
experimental values is attributed to the fact that the two-band model un-
derestimates the conduction band energies of larger nanocrystals [31].
From the single-particle spectrum, we evaluate the amplitude of the
TRFR signal as a function of probe energy,
F
(E), which is proportional to
the dierence of the dynamic dielectric response functions for

circularly
polarized light, as in Eq. 2.19 [41, 42, 43]. The conduction band electron
with S
z
= 1/2 created by the pump pulse relaxes rapidly to 1S
e
, such that

F
(E) is determined by optical transitions to the unoccupied 1S
e
state,
[1S
e
; ), [44]

F
(E) = CE

=1;|v
[1S
e
; [ p
x
+i p
y
[
v
)[
2
(3.1)

E E
X,v
(E E
X,v
)
2
+
2
v
.
The sum extends over all valence band states [
v
), E
X,v
(
v
) denotes the
energy (linewidth) of the 1S
e
-
v
exciton transition, and C is a constant.
Equation 3.1 implies that only transitions to the conduction band ground
state contribute to
F
(E). The transition matrix element is nite for S
3/2
valence band multiplets, i.e., for 1S
3/2
and 2S
3/2
in Fig. 3.8 [45]. Because
the energy splitting between 1S
3/2
and 2S
3/2
is of order 0.15 eV,
F
(E)
62
shows two well-dened resonances within 0.15 eV of the absorption edge. If
the crystal anisotropy is taken into account [39], both resonances split into
doublets, but the characteristic energy splitting is smaller than 25 meV.

F
(E) (displayed in Fig. 3.10 b for n
CdSe
= 3 and
v
= 10 meV) exhibits
only two distinct resonances, with a spectral weight that is larger for 1S
3/2
than for 2S
3/2
because of the larger overlap with the envelope wavefunction
of 1S
e
. The functional behavior shown in Fig. 3.10 b is in stark contrast to
the experimental data in Fig. 3.10 a, where at least three resonances with
comparable spectral weight can be identied close to the absorption edge.
Mixing of the 1S
3/2
and 1P
3/2
valence band multiplets is one possi-
ble mechanism which would explain both the additional resonances in

F
(E) and the large Stokes shift between the PL peak and the absorp-
tion edge [25]. Microscopically, mixing of 1S
3/2
and 1P
3/2
is eected by
broken spherical symmetry caused, e.g., by a uctuation of the quantum
well width. For illustration, we consider a perturbation to the valence band
Hamiltonian,
V (r) = v
0
sin (1 + cos ) , (3.2)
where (, ) are the azimuthal and polar angle of r relative to the crys-
tal symmetry axis. For a monolayer variation of the quantum well, v
0
is of order 0.1 0.2 eV for n
CdSe
= 3, 4. We diagonalize V (r) in the
63
Figure 3.10: Calculated FR spectra of quantum shells. (a) Measured FR spectrum
for QDQWs with a 3-monolayer CdSe shell. (b) Calculated FR spectrum for a spherical
QDQW. (c) Calculated FR spectrum for a spherical QDQW with an asymmetrical
perturbation. (d) the same as c, but with a stronger perturbation. Also shown is the
measured and calculated optical absorption for each case.
subspace spanned by the 1S
3/2
and 1P
3/2
multiplets for v
0
= 0.2 eV, ne-
glecting the perturbation V (r) for higher valence band multiplets. All
resulting eigenstates are superpositions of 1S
3/2
and 1P
3/2
states and have
64
a nite transition matrix element to the conduction band ground state.
The spectral weight in Eq. (3.1) is re-distributed among all eight states
and several additional resonances emerge in
F
(E) close to the absorption
edge. In Fig. 3.10 c and d, we show
F
(E) calculated for QDQWs with
broken spherical symmetry for two strengths of the perturbation for ran-
domly oriented nanocrystals, v
0
= 40 meV and 70 meV, with
v
= 12 meV.
Compared to spherical QDQWs (Fig. 3.10 b), the qualitative agreement
with experimental data improves when mixing of the 1S
3/2
and 1P
3/2
mul-
tiplets is taken into account. The anisotropy potential in Eq. (3.2) does
not provide a microscopic description for broken spherical symmetry, but
represents a simple model which allows for an analytical calculation of

F
(E). However, a perturbation of the form v
0
sin cos (Y
1,1
Y
1,1
)
is the lowest term in a multipole expansion for any realistic deformation
of the QDQW, such that Eq. (3.2) is expected to qualitatively capture
the essential features as long as the corresponding expansion coecient is
nite.
These results have been further modeled using density functional the-
ory in Schreier et al.. There is some concern that the k p description
above may be inaccurate, since it is assuming bulk crystal properties for
a structure just tens of atoms across. In this work, the crystal lattice of
65
the nanocrystal is simulated by building up the crystal atom by atom, ac-
cording to the wurtzite crystal structure. The structure is then allowed to
relax, minimizing the strain energy. This can lead to signicant deviations
from the bulk crystal structure, since a relatively large proportion of atoms
in the nanocrystal are at the surface. Then, density functional theory is
used to calculate the electron and hole wavefunctions in this crystal struc-
ture. The results for the energy levels are somewhat dierent than the k p
calculation, though it is hard to say whether the DFT calculation provides
a better match to the experimental results. Regardless, it is clear that
modeling these nanocrystal QDs is not a simple task. Deviations in the
shape, large amounts of strain, and surface defects make this a complicated
system to treat theoretically.
In this chapter, we have studied the spin dynamics and quantum size
levels in QDQWs using TRFR. The variation of the energy levels and the
electron g-factor with quantum well width allows one to selectively address
quantum wells using optical techniques. This yields an additional knob to
turn in engineering nanocrystal QD structures, in that the energy and g-
factor can be controlled while maintaining the same overall nanocrystal
diameter. The observed Faraday rotation spectrum has a complex line-
shape, in contrast to the smooth absorption spectrum. This reects the
66
spectrum of electron and hole energy levels in the nanocrystal, though the
complicated and irregular structure of these QDQWs make a quantitative
comparison to theory dicult. As we will see in the next chapter, this idea
can be taken a step further, and multiple quantum shells can be incorpo-
rated in each nanocrystal.
67
Chapter 4
Coupled Shells in Layered
Colloidal Nanocrystals
4.1 Motivation
In the eld of quantum information science, semiconductor quantum dots
(QDs) are of signicant interest for their ability to conne a single electron
for use as a qubit [46, 47]. However, to realize the potential oered by
quantum information processing, it is necessary to couple two or more
qubits. For nanocrystal QDs interconnected by conjugated molecules,
spin-conserving electron transfer between nanocrystals has been demon-
strated [48]. In contrast to coupling individual quantum dots, here we
demonstrate the integration of two coupled electronic states within a single
quantum dot heterostructure. These chemically-synthesized nanocrystals,
68
known as quantum dot quantum wells (QDQWs) [49, 21, 25, 50, 51], are
comprised of concentric layers of dierent semiconducting materials. In
the previous chapter, QDQWs were investigated with a low-bandgap shell
sandwiched between a high-bandgap core and outer shell. In contrast, now
we will look at structures with a low-bandgap core and outer shell, sepa-
rated by a high-bandgap barrier. We investigate carrier and spin dynamics
in these structures using transient absorption (TA) and time-resolved Fara-
day rotation (TRFR) measurements. By tuning the excitation and probe
energies, we nd that we can selectively initialize and read out spins in
dierent coupled states within the QDQW. These results open a pathway
for engineering coupled qubits within a single nanostructure.
4.2 Sample structure and characterization
The samples studied in this chapter are ensembles consisting of nanocrys-
tals with a 5.5 nm diameter, low-bandgap (E
g
= 1.74 eV) CdSe core, sur-
rounded by a 3 monolayer (ML), high-bandgap (E
g
= 3.68 eV) ZnS barrier,
and a 4 ML outer CdSe shell [51]. The nanocrystals were prepared with
the same type of colloidal chemistry techniques as described in Chapter 3.
Transmission electron microscopy (TEM) images show that the nanopar-
ticles are fairly uniform in shape and size and are roughly spherical with
69
ZnS barrier
CdSe core
CdSe
shell
V
r
e
h
Figure 4.1: Diagram of a coupled core-shell nanocrystal. In the core-shell nanocrystal
with a low band-gap core and outer shell, separated by a high band-gap barrier, electrons
and holes are conned to the core or the shell. The radial conduction and valence band
prole is shown, with schematic electron and hole wavefunctions.
some faceting at the surface [51]. A cut-away illustration of the sample
structure is shown in Fig. 4.1. Qualitatively similar results were also ob-
tained on a sample with 6.4 nm core, 2 ML barrier, and 4 ML shell. For
comparison, a control sample of 6.8 nm-diameter CdSe QDs (peak emission
at 5 K = 2.03 eV) was also prepared with nanocrystals purchased from Ev-
ident Technologies. For all of the samples, the nanocrystals were dispersed
in a solution of polyvinyl butyral (PVB) in dichloromethane. This mixture
was drop-cast into solid polymer lms, following the method of Ref. [52].
The optical density of the lms was around O.D. 1, giving them a slightly
translucent orange or brown appearance. The samples were placed in an
Oxford Spectromag magneto-optical cryostat. This allows measurements
at temperatures down to 4 K, and magnetic elds up to 6 T.
70
Energy (eV)



E
p
u
m
p
=
2
.
0
1
e
V
E
p
u
m
p
=
2
.
4
3
e
V
P
L

I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
A
b
s
o
r
p
t
i
o
n

(
O
.
D
.
)
1.75 2.0 2.25 2.5
0.0
0.4
0.8
T=295 K
Figure 4.2: Photoluminescence of coupled core-shell nanocrystals. With a high energy
pump, two PL peaks are seen from the core and the shell. With lower pump energy,
only the core PL is present. The optical absorption is also shown. The QDQWs are in
toluene solution at room temperature.
Figure 4.1 shows the radial potential of the core-shell structure along
with the conduction- (c) and valence- (v) band wavefunctions schemati-
cally illustrated (discussed more quantitatively below). The band prole is
analogous to a pair of coupled quantum wells in which the core corresponds
to one well and the shell to the other. Indeed, under 2.43 eV excitation the
PL spectrum of these QDQWs (Fig. 4.2) shows two peaks at 2.18 eV and
1.92 eV, which have been previously attributed to radiative recombination
from an electron-hole pair in the shell and in the core, respectively [51].
When the excitation energy is tuned between the core and shell emission to
71
2.01 eV, only the lower energy (core) emission is observed. This behavior of
the PL indicates that two optically-active, metastable exciton states exist
in the QDQWs, and that by changing the pump energy either the core or
both the core and the shell can be selectively excited. We refer to these as
metastable states since they both exist for a time-scale of tens of nanosec-
onds (see below). This is typical for the lowest energy exciton state in
most nanocrystal QDs, but usually any higher energy exciton states relax
to the exciton ground state on picosecond time-scales. For some reason,
the higher energy shell state cannot rapidly relax to the core state. This
may have something to do with the spatial separation, but the specic
details of energy relaxation in these QDQWs is still an open question.
Further evidence that these two PL peaks come from states localized in
the core and the shell can be seen by varying the core and shell dimensions.
Figure 4.3 shows PL spectra from a series of samples with xed (5.6 nm)
core diameter and varying shell thickness, and with xed (3 monolayer)
shell thickness and varying core diameter. The dashed lines provide a
guide to the eye showing mainly a shift in the lower energy peak as the
core diameter is changed, and a shift in the higher energy peak as the shell
thickness is changed.
72
core dia. shell
3.0 nm
3.5 nm
5.6 nm
6.4 nm
1 ML
2 ML
3 ML
4 ML
Figure 4.3: Coupled core-shell PL as a function of core and shell thickness. Left
column: PL from QDQWs with a 3 monolayer shell, and core diameter varied from
3.0 to 6.4 nm. Right column: PL from QDQWs with a 5.6-nm-diameter core, and
shell thickness varied from 1 to 4 monolayers. The dashed lines are guides to the eye.
Adapted from D. Battaglia et al., J. Am. Chem. Soc. 127, 10889 (2005).
4.3 Experimental setup
To perform time-resolved measurements of the spin and carrier dynamics
in these QDQWs (TRFR and TA, respectively), we use two optical para-
metric ampliers (OPAs) seeded and pumped by an amplied Ti:Sapphire
laser, as in Chapter 3. This allows pump-probe spectroscopy using two
independently tunable, 200 fs duration pulses ( 30 meV linewidth).
73
Ti:Sapphire
t
Circ. Pol.
(PEM for FR)
lin.
pol
pump
probe
~150 fs
sample
to diode
bridge for
(KR)
(not to scale)
Regenerative
amplifier
~150 fs
OPA2
choppers
monochromator
OPA1
white light
optional
mirror
for TA
(for TA)
to spectrometer
and PMT for TA
magneto-optical
cryostat
5 K
B
Figure 4.4: Setup for TRFR and TA measurements on coupled core-shell nanocrystals.
This setup is similar to the one for single shell measurements. Here, there is additionally
a white light pulse that can be used instead of the probe for TA measurements. The
white light transmission is measured using a spectrometer and photomultiplier tube
(PMT). Also, the sample is now in a magneto-optical cryostat, allowing for measure-
ments down to 4 K and up to 6 T.
Additionally, a broad-spectrum white light pulse is produced via super-
continuum generation. The output of one OPA is focused on the sample
as a pump pulse ( 50 m diameter, 0.5-1.0 mW) , and the output of the
other OPA or the white light is overlapped with the pump to serve as a
probe pulse, which is delayed in time using a mechanical delay line. A
schematic of the setup is shown in Fig. 4.4.
74
For TA measurements, the white light is used as the probe. After the
probe passes through the sample it is dispersed in a 0.5 m spectrometer and
detected with a photomultiplier tube. By mechanically chopping both the
pump and probe beams at dierent frequencies f
1
and f
2
, lock-in detection
at the sum frequency f
1
+f
2
can be used to isolate the TA signal.
To excite spin-polarized electrons into the QDQWs, the pump beam
polarization is modulated between right and left circular polarization at
a frequency f
PEM
= 42 kHz using a photoelastic modulator (PEM). As
the linearly polarized probe pulse is transmitted through the sample, its
polarization is rotated through an angle proportional to the spin polariza-
tion, due to the Faraday eect, as described in Section 2.4. The probe
beam is modulated by a mechanical chopper at a frequency, f
2
, of several
hundred Hz. The polarization rotation is measured by a balanced pho-
todiode bridge, and lock-in detection is used at both the PEM frequency
and the frequency of the mechanically chopped probe beam to isolate the
pump-induced spin polarization signal. Specically, two lock-ins are used.
First, the diode bridge dierence signal is sent to a lock-in amplier with
reference frequency at f
PEM
, and a time constant of 640 s. The Fast X
output of this lock-in is sent to a second lock-in, with reference frequency
f
2
.
75
4.4 Theoretical description
As in Chapter 3, we model the QDQWs using k p theory [36, 53, 38, 39].
While such a model is not rigorously applicable to structures with layers of
only a few monolayers, we nd that it still provides a surprisingly accurate
description of the observed spectra. The c-band states are described in a
two-band approximation, and we assume for the v-band states a four-band
Luttinger Hamiltonian in the spherical approximation. For the c band we
use the eective masses m
CdSe
= 0.12m
0
for CdSe, m
ZnS
= 0.28m
0
for
ZnS, and m
vac
= m
0
for the surrounding material (which we assume to be
vacuum), where m
0
is the bare electron mass. The Luttinger parameters
for the v band are given by
1,CdSe
,
CdSe
= 1.67, 0.56 for CdSe and

1,ZnS
,
ZnS
= 2.00, 0.75 for ZnS. For simplicity, we assume a spherical
shape of the coupled core-shell structure, with a CdSe core extending from
r = 0 to r = r
1
, a ZnS barrier extending from r = r
1
to r = r
2
, and
a CdSe shell extending from r = r
2
to r = r
3
(see Fig. 4.1). Under the
assumption of spherical symmetry, the c-band and v-band states can be
classied by the quantum numbers nL
e
and nL
F
, respectively [38, 39, 36,
53]. Here, the total angular momentum F = [J L[ is the sum of the
angular momentum of the Bloch function, J, and the angular momentum
of the envelope function, L. The radial quantum number is n. For both
76
the c band and the v band we assume a potential barrier of 0.9 eV for ZnS
with respect to CdSe. The barrier height to the surrounding material is
taken as 4 eV (innite) for the c (v) band. We take the ML thickness to
be 0.31 nm for ZnS and 0.35 nm for CdSe (the interplane distances in the
[002] direction of the wurtzite crystal) [18]. We solve the radial Schrodinger
equation piecewise and take into account the boundary conditions at r = r
i
as explained in Reference [36]. To calculate exciton energies we include the
Coulomb interaction to rst order.
The calculated radial wavefunctions (r) are shown in Fig. 4.5 for the
two c-band states, 1S
e
and 2S
e
(black) and the two v-band states, 1S
3/2
and 2S
3/2
(red) for r
1
= 2.75 nm, r
2
= 3.68 nm, and r
3
= 5.08 nm. It is
clear from Fig. 4.5 that the electron and hole are localized mainly in the
core of the QDQW for the 1S
e
1S
3/2
exciton, and mainly in the shell for
the 2S
e
2S
3/2
exciton. This latter state is the energetically lowest state
with the electron and hole strongly localized in the shell. We therefore
assign these two states to the core and shell PL peaks.
The wave functions and energy spectra vary signicantly when changing
the radii r
i
. We obtain fairly good agreement by comparing the calculated
energies of the lowest core and shell states to the PL data by Battaglia et
al. (Ref. [51], Fig. 4.3) for several sets of r
i
reported in their work. The
77


0 2 4 6
-0.4
0
0.4
-0.4
0
0.4
r (nm)
r

(
n
m
-
2
)
core shell

2S
e
2S
3/2
1S
e
1S
3/2
Figure 4.5: Calculated core and shell wavefunctions. Top: the calculated 1S
e
electron
(black) and 1S
3/2
hole (red) radial wavefunctions, localized in the core. Bottom: the
calculated 2S
e
electron (black) and 2S
3/2
hole (red) radial wavefunctions, localized in
the shell. The two red curves are the L = 0 and L = 2 components of the wavefunction.
comparison of these energies is shown in Fig. 4.6. This indicates that the
k p model we apply here reasonably describes the size-dependent trends of
the experimentally observed low-lying exciton energies. Specically, when
the core diameter is increased, the core state PL redshifts and the shell
state PL remains roughly constant. On the other hand, increasing the shell
thickness causes a redshift in the shell PL, leaving the core PL essentially
the same.
78







theory
experiment
core
shell
2 3 4 5
Shell thickness (ML) Core diam. (nm)
4 5 6 3
2.8
2.4
2.0
E
n
e
r
g
y

(
e
V
)
(a) (b)
5.6 nm core, 3 ML barrier
3 ML shell, 3 ML barrier
Figure 4.6: Measured and calculated energy levels. (a) Calculated and measured
core and shell energy levels for a 5.6-nm-diameter core as a function of shell thickness.
(b) Calculated and measured core and shell energy levels for a 3 monolayer shell as a
function of core diameter.
As in Chapter 3, in order to calculate the g-factor for the lowest c-band
core and shell states [54], we rst determine the g-factor (as in Eq. 2.9) in
the two materials,
g
CdSe,ZnS
= 2
2
3
E
p

so
(E
g
+
so
+E) (E
g
+E)
(4.1)
where E is the quantum connement energy, and the band gap energy, spin-
orbit splitting, and Kane energy for bulk crystals are given by E
g
,
so
, E
p
=
1.75, 0.4, 20.3 eV for CdSe and E
g
,
so
, E
p
= 3.55, 0.09, 20.4 eV for
ZnS [40]. The g-factor is then approximated by g = (p
core
+ p
shell
)g
CdSe
+
p
barrier
g
ZnS
+p
vac
g
vac
, where we have weighted the g-factors of CdSe, ZnS,
and the surrounding vacuum (g
vac
= 2) with the probability p

of the elec-
79
tron to be in the region . Here we have neglected terms that arise due to
the interfaces [55]. For the 1S
e
state, which is mainly localized in the core,
we obtain g
CdSe
, g
ZnS
= 0.966, 1.918 and p
core
, p
barrier
, p
shell
, p
vac
=
0.899, 0.076, 0.024, 0.000, yielding g = 1.04. In contrast, the param-
eters for the 2S
e
state, which is mainly localized in the shell, are ob-
tained as g
CdSe
, g
ZnS
= 1.101, 1.922 and p
core
, p
barrier
, p
shell
, p
vac
=
0.043, 0.084, 0.851, 0.021, yielding g = 1.19.
4.5 Transient absorption and luminescence
measurements
We employ transient absorption (TA) measurements to probe the carrier
dynamics in these QDQWs. Here, a pump pulse excites carriers within the
QDQWs, and a spectrally broad probe pulse then measures the change in
the optical absorption spectrum induced by the pump. The left panel of
Fig. 4.7 shows the TA spectrum from the CdSe QD control sample at a xed
pump-probe delay of 20 ps, with a pump energy of 2.30 eV (red) and 2.03
eV (black). Two eects contribute to the observed spectrum: strictly pos-
itive signal due to bleaching of interband transitions, and signal caused by
shifts in the spectrum due to multiparticle eects [56, 30]. (Here, positive
signal refers to reduced absorption.) The rst eect is simply due to the
80
Pauli exclusion principle. If the pump excites an electron into a particular
conduction band state, then the probe cannot excite a second electron into
the same state, and the absorption is reduced. The second eect results
in both positive and negative contributions. This eect is due to a shift
induced in the energy level spectrum due to the pump pulse. For example,
if the pump pulse creates an exciton in a QD, the energy for the probe to
add a second exciton is not the same as to create the rst one, because
of the exciton-exciton binding energy. Since we are measuring the change
in the absorption spectrum due to the pump, this pump-induced shift in
energy shows up as something like the rst derivative of the absorption
spectrum.
The large peak at 2.05 eV is attributed to bleaching of the lowest opti-
cally active transition. It is apparent that the TA spectrum in the control
sample is largely independent of pump energy. The time-dependence of
the TA signal at 2.03 eV up to the maximum pump-probe separation of
2.5 ns is shown in the inset of Fig. 4.7, and is essentially the same in both
samples at all energies at the time-scales shown here. This indicates that
the lifetime of the electrons and holes is much longer than 2.5 ns, which is
usually the case in this type of QD, and which we reconrm below.
The results of the same TA measurements performed on the QDQW
81





T
A

(
a
r
b
.

u
n
i
t
s
)
1.75 2.25 2.75 1.75 2.25 2.75
Energy (eV)
E
pump
:
CdSe QDs (control) QDQWs
t=20 ps
T=5 K
0
T
A
2 1 0


Delay (ns)
Figure 4.7: Transient absorption measurements on coupled core-shell nanocrystals.
Left: TA at a delay t = 20 ps on the CdSe QD control sample, with pump energy
2.03 eV (black) and 2.30 eV (red). Right: The same measurement on the coupled core-
shell QDQWs. Inset: TA as a function of delay on the QD sample at a probe energy
2.03 eV. The black triangles indicate the calculated core and shell energies.
sample are shown in the right panel of Fig. 4.7. At the low pump en-
ergy (2.03 eV), where only the core state is excited, the TA spectrum is
very similar to the spectrum of the CdSe QDs. However, at higher pump
energy (2.30 eV) a second large positive peak appears due to bleaching
of transitions to the shell state. The black triangles in Fig. 4.7 indicate
the calculated energies of the lowest core and shell transitions, 1S
e
1S
3/2
(2.010 eV) and 2S
e
2S
3/2
(2.254 eV), in reasonable agreement with the two
bleaching peaks in the TA spectrum. This provides further conrmation
that a second, higher-energy metastable (lifetime 3 ns) state exists in
82


0 50 100 150
Time (ns)
N
o
r
m
a
l
i
z
e
d

P
L

(
a
.
u
.
)
2.18 eV filter, = 37 ns
1.94 eV filter, = 20 ns
T = 295 K
Figure 4.8: Time-resolved PL from coupled core-shell QDQWs in toluene solution
taken using time-correlated photon counting. The black curve was taken with an inter-
ference lter centered at 2.18 eV (shell PL), and the red curve centered at 1.94 eV (core
PL). The band-pass FWHM of the lters was about 40 meV. The kinks at 13 ns are
artifacts of the laser pulses.
the QDQWs. In fact, time-resolved PL measurements (shown in Fig. 4.8),
performed by time-correlated photon counting on QDQWs in toluene solu-
tion show both core and shell states to have a room-temperature radiative
lifetime of 20 ns, similar to previously measured radiative lifetimes in
CdSe QDs [57]. (The TRPL measurements were carried out on the same
sample as the PL in Fig. 4.2.) The reason why these PL measurements
were performed at room temperature and in solution is that the process
of making the PVB lms substantially degrades the PL. It is likely that
additional defect states at the surface of the nanocrystals are introduced,
83
which can lead to bleaching of the PL, particularly the PL from the shell
state, as was observed. Nevertheless, even if the radiative recombination
is quenched, the TA measurements still conrm that the lifetime of both
states is much greater than 3 ns at low temperature as well.
4.6 Time-resolved Faraday rotation
spectroscopy
Using a circularly polarized pump pulse, we can excite spin-polarized car-
riers into these two metastable states, and then probe the resulting spin
dynamics through TRFR. A 2-T magnetic eld is applied perpendicular to
the pump and probe direction (Voigt geometry). The left column of Fig. 4.9
shows characteristic spin precession in the QDQW sample measured by
TRFR at several pump and probe energies. The corresponding Fourier
transforms (FTs) are shown in the right column. (The non-precessing
component of the signal has been subtracted before calculating the FT.)
At this magnetic eld, the spin lifetime is dominated by inhomogeneous
dephasing due to the size and shape distribution of the nanocrystals. It
has been previously observed for CdSe QDs [31], as well as QDQWs with
a single well [54], that the spin dynamics exhibit two distinct precession
frequencies. The origin of these two precession frequencies (or g-factors) is
84









F
T

a
m
p
l
i
t
u
d
e

(
a
r
b
.

u
n
i
t
s
)
0 50 100
Frequency (ns
-1
)










E
pump
E
probe
2.07 eV 1.97 eV
2.07 eV 2.30 eV
1.97 eV
2.30 eV
2.43 eV
2.43 eV
F
a
r
a
d
a
y

r
o
t
a
t
i
o
n

(
a
r
b
.

u
n
i
t
s
)
0
0
0
0
0 200 400
Time (ps)
(a)
(b)
(c)
(d)
B=2.0 T, T=5 K
g=1.34
g=1.48
g=1.70
Data
Fit
Data
Fit
Figure 4.9: Time-resolved Faraday rotation on coupled core-shell nanocrystals. (a)
TRFR with low energy pump and low energy probe. (b) TRFR with low energy pump
and high energy probe. (c) TRFR with high energy pump and low energy probe. (d)
TRFR with high energy pump and high energy probe. Black is data, and red is a t
(a-c are two frequency ts, d is a three frequency t). The right column shows the
corresponding Fourier transforms (the non-precessing component is subtracted before
taking the FT). The three precession frequencies are indicated by the dotted lines.
85
still unclear, though it has been suggested that one precession frequency
may be due to lone electron spins and the other to exciton spins [31, 35].
These two behaviors might be due to dierent defect charge states (see the
discussion in Chapter 3). It has been shown for CdSe QDs that the relative
amplitude of the two precession components depends on the pump energy,
but is essentially independent of the probe energy [31]. In the core-shell
QDQWs, with the pump energy tuned to excite carriers only into the core
(E
pump
= 2.07 eV) as shown in Fig. 4.9 a and 4.9 b, two frequencies are
observed with only small changes in their relative amplitude as the probe
energy is varied. This is similar to the case in CdSe QDs. When pumping
at higher energy, however, where both the core and the shell are excited,
signicant change in the ratio of the two frequency components is observed
with changing probe energy (Fig. 4.9 c and d). Furthermore, a third pre-
cession frequency appears at high probe energy. In Fig. 4.10, this is shown
in a close-up view of the FT in Fig. 4.9 d. All three of these frequencies
vary linearly with the applied eld, and correspond to g-factors of 1.34,
1.48, and 1.70.
For a more detailed look at the probe energy dependence of the spin
dynamics, the probe beam was passed through a monochromator after the
sample to narrow the probe linewidth to 5 meV and TRFR measure-
86



30
40 50
Frequency (GHz)
F
T

a
m
p
l
i
t
u
d
e
Figure 4.10: Three precession frequencies in coupled-core shell nanocrystals. The FT
of spin precession in QDQWs with E
pump
= 2.43 eV and E
probe
= 2.30 eV. The red line
is a t to the sum of three Lorentzians, shown individually in green.
ments were performed over a range of probe energies. The resulting curves
were normalized by the probe power and t by

F
(t) = C +A
0
e
t/
0
+
n

i=1
A
i
e
t/
i
cos (
i
t +
i
) , (4.2)
where n = 2 or 3 as needed to provide a good t. The precession frequen-
cies,
i
, and the lifetimes,
i
, were xed for all the ts for a given probe
energy scan. That is, for one probe energy where the signal was good, all
the parameters were allowed to vary in the t and then the resulting values
of
i
and
i
were xed for the rest of the ts in that probe energy scan. The
red curves in Fig. 4.9 show ts to the data. The origin of the non-oscillating
components C and A
0
is unclear and will not be discussed further here.
The amplitudes A
i
(i = 13) correspond to the dierent frequency compo-
87
nents seen in the FTs in Fig. 4.9. As was previously observed, in the CdSe
QD control sample the ratio of amplitudes of the frequency components
is largely independent of probe energy, though dependent on pump energy
(Figs. 4.11 a and 4.11 b). Furthermore, the probe energy spectrum has es-
sentially the same shape at both pump energies (though A
1
is quite small
at the higher pump energy). This is expected since the carriers rapidly
relax to the lowest exciton state [56], regardless of pump energy.
As before, the QDQW sample shows similar behavior to the control
sample when only the core state is excited (Fig. 4.11 c). However, when
both the core and the shell are excited a third precession frequency ap-
pears only at high probe energy, as shown in Fig. 4.11 d. The amplitude
of Faraday rotation is a maximum for probe energies near interband tran-
sitions involving the occupied energy levels (see Eqs. 2.19, 3.1, and Ref.
[36]). This probe energy dependence implies that the spins precessing at
the third frequency are in a higher energy state than those precessing at
the other two frequencies. Since the position of the third frequency peak
corresponds to the higher energy peak in the TA spectrum in Fig. 4.7, it
is reasonable to assign this precession frequency to electrons in the shell
state.
88

1
=118 GHz

2
=150 GHz

1
=99 GHz

2
=142 GHz



1
=101 GHz

2
=141 GHz



1
=119 GHz

2
=151 GHz

3
=131 GHz
1.9 2.1 2.3 1.9 2.1 2.3
0
0
0
0
Energy (eV) Energy (eV)
F
a
r
a
d
a
y

r
o
t
a
t
i
o
n

a
m
p
l
i
t
u
d
e

(
a
r
b
.

u
n
i
t
s
)
E
pump
E
pump
CdSe QDs
(control)
QDQWs
(a) (b)
(c) (d)
B = 1 T, T = 5 K
Figure 4.11: Faraday rotation spectra of coupled core-shell nanocrystals. Amplitude
of the two FR frequency components versus probe energy in the CdSe QD control
sample, with (a) low energy pump (E
pump
= 2.03 eV), and (b) high energy pump
(E
pump
= 2.30 eV). Amplitude of the FR frequency components versus probe energy
in the coupled core-shell QDQW sample, with (a) low energy pump (E
pump
= 2.03 eV,
two frequencies), and (b) high energy pump (E
pump
= 2.30 eV, three frequencies).
4.7 Analysis of spin dynamics and
core-shell coupling
As discussed above, based on our calculated wavefunctions we have esti-
mated the eective g-factor of the lowest core and shell c-band states by
89
g = (p
core
+p
shell
)g
CdSe
+p
barrier
g
ZnS
+p
vac
g
vac
, where we have weighted the
g-factors in the CdSe, ZnS and surrounding vacuum with the probability
p

of the electron to be in the region . For the state 1S


e
, which is mainly
localized in the core, we obtained g = 1.04, and for the 2S
e
state, which
is mainly localized in the shell, we obtained g = 1.19. Comparing these
two calculated values to the observed g-factors, we see that the calculated
values are a bit smaller than any of the observed g-factors, but the dier-
ence of the two calculated g-factors is in good agreement with that of the
two lowest measured g-factors. This underestimation of the g-factors by
about 20% might not be too surprising since this k p model used here is
not particularly rigorous. A more accurate theoretical description of these
structures could be gained through detailed rst-principles calculations [58]
or by taking into account interface terms [55].
In order to obtain an estimate for the magnitude of the electron ex-
change interaction that may be expected in QDQW structures, we com-
pare our theoretical results for the conduction band states, 1S
e
and 2S
e
, of
the core-shell structure with a two-site Hubbard model to determine the
tunneling matrix element between the core and the shell.
For the conduction band, we consider the ground state
core
of a CdSe
core which is surrounded by a ZnS barrier and the ground state
shell
of a
90
CdSe shell that is surrounded by vacuum and encloses a ZnS core. Using
the same r
i
as above, we obtain E
core
= 0.2222 eV and E
shell
= 0.3655 eV
for the ground state energies of the decoupled core and shell and we dene
:= E
shell
E
core
= 143.3 meV. In the Hubbard model, the orthogonal
states
core
and
shell
are coupled by the tunneling matrix element t
c
. The
two lowest-lying states and energies are in lowest order in the tunneling

N [
core
(t
c
/)
shell
] , E

E
core
t
2
c
/,

+
N [
shell
+ (t
c
/)
core
] , E
+
E
shell
+t
2
c
/,
(4.3)
where N = /
_

2
+ t
2
c
. In the above states, the tunneling probabil-
ity between the core and the shell is given by p
t
= (t
c
/)
2
. Given a
calculated wavefunction in the coupled structure, we can alternatively de-
termine p
t
from the relative weight of the wavefunction in the core and
in the shell. From the states 1S
e
and 2S
e
discussed above we extract an
average tunneling probability p
t
= 0.046, and thus derive [t
c
[ = 30.7 meV,
E

0.2157 eV, and E


+
0.3721 eV. These energies agree well with
E
1Se
= 0.2188 eV and E
2Se
= 0.3717 eV obtained from the coupled core-
shell model. The corresponding wavefunctions are shown in Fig. 4.12.
There is good agreement between calculations of the wavefunctions and
energies when the whole core-shell potential is used at once, and when the
core and shell are calculated separately and then coupled using this Hub-
91







0 2 4 6 0 2 4 6
0.8
0.4
0
r

(
r
)
r (nm) r (nm)
1S
e
2S
e

+
r
1
r
2
r
3
(a) (b)
Figure 4.12: Coupled wavefunctions determined using the Hubbard model. Core (a)
and shell (b) electron wavefunctions calculated in the coupled potential (black) and the
and in the separate core and shell potentials and then coupled in the Hubbard model
(red).
bard model. This does not address the overall validity of the k p model,
since it used in both calculations. However, the self-consistency does lend
credence to the use of the Hubbard model here.
The exchange interaction of an electron located in the core and an
electron located in the shell can be estimated in the Hubbard model as
follows. We rst estimate the direct Coulomb interaction of two electrons
in the present structure, assuming a uniform dielectric constant = 9. We
obtain the value U
core
78 meV when both electrons are in the core ground
state
core
. For the cases when both electrons are in the shell ground state,

shell
, or one is in the state
core
and the other one in
shell
, we obtain
the direct Coulomb interaction energies U
shell
38 meV and U
core,shell

92
35 meV, respectively. Because the states
+
and

are strongly localized


in the present structure, these Coulomb energies are a good approximation
for the corresponding interaction energies. Hence, the net on-site Coulomb
repulsions are much smaller than for the structure under study, and
the exchange interaction can be approximated by t
2
c
/ 1 meV. For
comparison, we obtain a value on the same order of magnitude, 1.8 meV,
for the Coulomb exchange integral for the case when the two electrons
occupy the states 1S
e
and 2S
e
.
On the basis of these calculations, we see that QDQWs with small
might provide an interesting basis to study interacting spins in coupled
nanostructures. can be tuned all the way through zero by changing
the diameter of the core and the thickness of the shell [51]. An attempt
was made to see eects of two-electron interactions in the present samples.
To do this, a second pump pulse was added to the experiment. The idea
was that the rst pump would excite an electron into the core, and the
second pump would excite another electron into the core or the shell. If
one electron ended up in the core, and one in the shell, their spins would
evolve under the mutual exchange interaction which would be visible in the
TRFR measurement. Both pumps were modulated at dierent frequencies
using two PEMs, and the rst lock-in was set to the dierence or sum
93
frequency. However, it turned out that there are a number of unwanted
background signals that obscured the desired eects, if they were present.
To date, spin-spin coupling in these structures has not been observed.
In conclusion, we have fabricated and measured spin dynamics in cou-
pled core-shell nanocrystal QDQWs. We have observed evidence in the
time-resolved PL, and transient absorption measurements that these struc-
tures contain two metastable excited states. One state is localized mainly
in the core, and the other in the shell. Calculations conrm that these two
coupled states exist in the nanocrystal. The PL, TA, and TRFR measure-
ments all indicate that by varying the energy of the pump, spins can be
selectively excited into the core state alone, or into both the core and the
shell. Furthermore, as we saw in Chapter 3, the electron g-factor and the
energy dependence of the Faraday rotation depend on the conning poten-
tial in the nanocrystal. Therefore, using time-resolved Faraday rotation
spectroscopy, we can selectively read out the spin in either the core or the
shell by the diering precession frequencies and spectral dependences.
4.8 Nanocrystal QDs in an optical cavity
The experiments discussed above have all been performed on large en-
sembles (> 10
8
) of QDs. This ensemble averaging prohibits a number of
94
interesting experiments, such as observing quantum jumps between spin
states, entangling two spins, or even measuring the transverse spin lifetime
without inhomogeneous dephasing. Additionally, virtually all proposals
for spin-based quantum information processing require the ability to work
with single electron spins.
Of course, going from a measurement of > 10
8
spins down to one spin
will involve a large decrease in signal. In order to address this issue, we
have explored the fabrication of integrated optical cavities to enhance the
Faraday (or Kerr) rotation measurement of spins in quantum dots [59].
A Fabry-Perot optical cavity consists of two parallel mirrors. Light
in the cavity reects back and forth between the mirrors. The cavity
supports optical modes at wavelengths such that the round-trip path length
between the mirrors is a whole number of wavelengths. When this condition
is satised, then subsequent reections interfere constructively and the
intensity of the light builds inside the cavity. This enhancement of the
light intensity has been used before to enhance the Faraday eect in other
systems [60, 61, 62, 63]. In this work, alternating dielectric layers of TiO
2
and SiO
2
were deposited on a glass substrate which form a distributed
Bragg reector (DBR). Due to constructive and destructive interference,
a DBR reects light of wavelength , when alternating materials with
95
dierent indices of refraction are layered with layer thickness equal to /4
(for the in the dielectric material). 11 layers of these materials were
deposited on the substrate using e-beam evaporation, and were capped by
a nal SiO
2
layer to serve as the bottom mirror of the cavity. This surface
was chemically functionalized to bind CdSe nanocrystal QDs. The sample
was then immersed in a toluene solution containing 6.6-nm-diameter CdSe
QDs [64]. The chemical functionalization and QD treatment was repeated
until the desired optical density of the QD lm was reached. Then a
wedge shaped SiO
2
layer is deposited on top of the QDs, to complete the
cavity. The wedge shape causes the thickness of the cavity, and therefore,
the position of the cavity resonance, to vary across the sample. Then, a
second, identical DBR was deposited on top. For comparison, half of the
sample was masked during the deposition of the top and bottom DBR
layers (though not for the SiO
2
layers making up the cavity itself). Thus
half of the sample consists of CdSe QDs embedded just in SiO
2
, and the
other half is the same, but also embedded between two DBRs. Figure 4.13
shows a schematic of the sample, and also a cross-sectional transmission
electron micrograph (TEM) of the structure.
Two samples were fabricated, with dierent densities of QDs Sample
A with higher density, and Sample B with lower density. Both samples
96
TiO
2
SiO
2
bottom DBR
top DBR
QDs
control region
cavity
(a)
1 m
20 nm
QDs
bottom DBR
top DBR
cavity
Figure 4.13: Diagram of nanocrystals embedded in an optical cavity. Top: Schematic
of SiO
2
/TiO
2
distributed Bragg reectors (DBRs) forming an optical cavity around a
layer of CdSe nanocrystal QDs. Top layers are transparent to show the interior. Bottom:
cross-section electron micrograph of the sample structure, showing the alternating layers
of the DBRs. A close-up of the center of the cavity shows individual QDs.
show the same absorption spectrum, with the amplitude scaled by the QD
density. As seen by the sharp peaks in the transmission spectra shown
97








%

T
r
a
n
s
m
i
s
s
i
o
n80
100
60
40
20
0
Sample A
550 600 650 700 750 800
%

T
r
a
n
s
m
i
s
s
i
o
n
80
60
40
20
0
550 600 650 700 750 800
(nm) (nm)
Sample B
10
20
0
30
30
0
60
630 640 620 630 640
Q = 110 Q = 165
100
Figure 4.14: Optical transmission spectra of the cavity/QD samples at dierent po-
sitions on the sample. The peak around 600-650 nm is due to the cavity resonance,
which tracks the cavity thickness. The insets show one of the resonance peaks from
each sample, with a t that is used to determine the Q-factor.
in Fig. 4.14, the resonance of the cavity can be tuned from about 600 to
640 nm in both samples by translating across the sample. The quality
factor (Q-factor) of the cavities, which is related to the number of reec-
tions made by the light within the cavity, was found to vary from about
80 to 180 at dierent positions on the sample. (The sample with lower QD
density has higher a Q-factor, since less light is lost from QD absorption.)
By comparing the TRFR signal at the peak of the cavity resonance, with
the same signal measured without the cavity, we can see that the cavity
signicantly enhances the FR signal. Figure 4.15 a and b shows TRFR
traces from the two samples at room temperature, with a transverse mag-
98
F
R

(
m
r
a
d
.
)








0 200 400 600
Delay Time (ps)
Delay Time (ps)
0 200 400 600 100 150 200
Q-Factor
1.5
0
1.0
0.5
0.3
-0.3
0
0.05
-0.05
0
10
15
30
25
20
Sample B
Sample A
B = 0.47 T T = 298 K
probe
= 630 nm
Control B
Control A
(a)
(b) (c)
F
R

(
m
r
a
d
.
)
E
n
h
a
n
c
e
m
e
n
t
Figure 4.15: Cavity-enhanced time-resolved Faraday rotation. (a) Time-resolved FR
on both samples shown on the cavity resonance (black) and on the control region with
no cavity (green and blue). (b) Zoom-in of the green and blue data in (a). (c) The
enhancement of the FR signal due to the cavity as a function of Q-factor.
netic eld of 0.47 T, with and without the cavity. This enhancement is
expected, since the polarization of the light is rotated further with every
pass through the QD layer.
The enhancement factor caused by the cavity is plotted in Fig. 4.15 c as
a function of the cavity Q-factor. As expected, the enhancement is roughly
linear, with a maximum enhancement of a factor of about 28 for a Q-factor
of 180.
99
From an estimate of the QD density and the probe laser focused spot
size, we estimate that in the lower density sample, we are probing about
910
8
QDs. Given the signal-to-noise ratio, and if we were to use a smaller
spot size, we estimate a detection limit of 310
4
QDs for this sample. This
is still a ways away from a single QD, but as we will see in the next chapter,
this problem can be overcome.
100
Chapter 5
Non-destructive Measurement
of a Single Electron Spin
5.1 Motivation and Background
The prospect of quantum computation in conventional material systems
has spurred much research into the physics of carrier spins in semicon-
ductor quantum dots (QDs) [47]. An important element necessary for
spin-based quantum computing is the read-out of the qubit spin state.
Previously demonstrated schemes for single spin read-out in a quantum
dot (examples illustrated in Fig. 5.1) include optical measurements, such
as photoluminescence (PL) polarization [65, 66] or polarization-dependent
absorption [67, 68, 69]. For example, in reference [65], spins are optically
excited into GaAs interface QDs (discussed below), using above-bandgap
101
circularly
polarized PL
spin-dependent
absorption
l
e
a
d
l
e
a
d
d
o
t
electrical readout
(a) (b) (c)
Figure 5.1: Previously demonstrated schemes for single spin detection. (a) Exciton
spin readout via the circular polarization of the PL. (b) Single electron spin detection us-
ing polarization-dependent absorption. (c) Electrical readout based on spin-dependent
tunneling probability out of the QD.
circularly polarized light. The PL from a single QD is collected with a
microscope objective. Due to the selection rules for interband optical tran-
sitions, the circular polarization of the resulting luminescence yields infor-
mation about the spin of the electron and/or hole in the QD. Alternatively,
in reference [69] for example, a narrow-linewidth laser is used to probe the
optical absorption of a single QD. Due to the Pauli exclusion principle, if
an electron is in the quantum dot, the optical transition to that state is
blocked, resulting in a polarization-dependent change in the optical absorp-
tion. If the light is focused tightly on the QD, this single QD absorption
can be surprisingly large.
Single spins can also be read out electrically by measuring the spin-
dependent probability for an electron to tunnel out of the dot [70]. In
102
these experiments, QDs are formed using electrical gates to deplete regions
of a two-dimensional electron gas. The QDs formed within these depleted
regions are easy to contact electrically by leaving a small gap in the depleted
area. Additional gates can be used to shift the energy levels in the QD, and
to control the tunnel coupling to the electrical leads. With a suciently
large magnetic eld and low temperature, the spin states of the QD energy
levels can be split with an energy greater than k
B
T. In this case the QD
energy levels can be shifted so that one spin level is above the Fermi energy
in one of the leads, and the other spin level below the Fermi energy. In this
situation, the higher energy spin state can tunnel out of the QD, resulting
in a measurable current. This current constitutes readout of the electron
spin.
However, these previous single spin readout methods are destructive,
in that they either remove the spin from the dot, or drive transitions in the
system with a resonant optical eld. In other words, these measurements
take the system out of the 2-state, spin-up/spin-down Hilbert space. To
avoid confusion, it is worth dening the term destructive. This term is
used (for example, in Ref. [71] and [72]) as a necessary, but not sucient
condition for a quantum nondemolition (QND) measurement. A QND
measurement must satisfy two conditions: 1. The measurement must not
103
take the state out of the Hilbert space of the system being measured (non-
destructive measurement), and 2. After the measurement, the system must
be in an eigenstate with probability given by the quantum state before
the measurement (projective measurement). Another way of stating this
second condition is that the Hamiltonian of the system must commute with
the Hamiltonian of the measurement, or that back-action of the probe
on the system must be avoided. There has been one previous claim of
non-destructive measurement in the singlet-triplet basis in an electrically-
gated double QD system [73]. However, in this work, one of the electrons
is removed from the QD and then later replaced with an electron in the
same spin state. This seems to satisfy the second condition, not the rst,
making theirs a unique denition of the term non-destructive.
In contrast, we describe measurements of a single electron spin using
Kerr rotation (KR) in which the spin state is probed non-resonantly, thus
minimally disturbing the system. This eective spin-photon interaction
has been used with great success to measure spins non-destructively (and
in fact to perform QND measurements) on ensembles of atoms in a gas.
Specically, Faraday rotation of a probe laser has been used to observe
quantum eects such as measurement-induced decoherence, spin squeez-
ing [74, 75], and teleportation of quantum information [76]. Also, a number
104
of proposed quantum information protocols use Faraday or Kerr rotation
of single spins to generate spin-photon entanglement [77] and optically-
mediated spin-spin entanglement [78, 79].
In the last section, we saw that we were unable to measure less than
about 3 10
4
QDs using Faraday rotation. So is it reasonable to hope for
single QD sensitivity? First of all, we should more carefully choose the
type of QD we are looking at. As discussed in Section 2.4, the magni-
tude of Faraday rotation depends on the momentum matrix elements con-
necting the initial and nal (valence and conduction band) states. Using
the commutation relation for the position operator and the Hamiltonian,
[r, H] p, these matrix elements can be expressed in terms of the position
operator that is the dipole matrix elements:
c[ p[v) c[rH Hr[v) = (E
v
E
c
)c[r[v). (5.1)
From these matrix elements, it can be seen that the Faraday or Kerr eect
will be larger for QDs with a larger spatial extent of the electron and hole
wavefunctions. The nanocrystal QDs in Chapters 3 and 4 are about 5 nm
in all directions. In contrast, electrons in interface uctuation quantum
dots are conned to a region on the order of 100 100 5 nm. As a
result, transitions in nanocrystal QDs typically have dipole matrix elements
with modulus squared of several Debye [80], whereas interface QDs have
105





1 QD
(10 QDs)/10
(10
4
QDs)/10
4
F
a
r
a
d
a
y

r
o
t
a
t
i
o
n

(
n
o
r
m
a
l
i
z
e
d
)
P
r
o
b
a
b
i
l
i
t
y

d
i
s
t
r
i
b
u
t
i
o
n
Energy (a.u.)
Figure 5.2: Single versus ensemble Kerr rotation spectrum. The calculated KR spectra
are normalized by the number of QDs in each case. The black curve shows a single odd-
Lorentzian KR spectrum, as from a single QD. The blue curve shows the sum of 10 such
KR spectra with energies distributed randomly according to the distribution shown in
gray. The red curve shows the same for 10
4
QDs. The normalization by the number of
QDs highlights the fact that the signal is not N times larger for N times as many QDs.
dipole strengths of 50-100 Debye [81]. Therefore, by using these interface
uctuation QDs, we gain a signicant factor in the signal.
It is also important to realize that an ensemble of N QDs will not yield
N times the Faraday rotation signal of one QD. Each QD containing a
spin contributes an odd-Lorentzian feature to the total Faraday rotation
signal centered at the optical transition energy and with the linewidth
of the transition, . Since an ensemble of QDs will be spread out over
a region of energy E, the maximum signal from an ensemble of QDs
106
will be reduced by a factor of roughly E/. Figure 5.2 illustrates this
point. (Note that in the gure, the curves are normalized by the number
of QDs. The actual, non-normalized signal does increase as the number of
QDs increases, but sub-linearly). This leads to another gain in the single
QD signal over the ensemble measurement. Finally, if we can improve the
measurement further, for example, by reducing the noise in the lasers or in
the detection electronics, we may have a hope of measuring a single spin
in a single QD.
5.2 Sample structure
A schematic of the nal sample structure is shown in Fig. 5.3, and details
of the sample fabrication are given in Appendix A. The sample is grown
by molecular beam epitaxy and consists of a single 4.2-nm GaAs QW in
the center of a planar Al
0.3
Ga
0.7
As -cavity. A 2-min. growth interruption
at each QW interface allows large ( 100 nm diameter [81]) monolayer
thickness uctuations to develop that act as QDs [82, 83]. An STM image
of this type of uctuation, along with a cartoon of interface QDs are shown
in Fig. 5.4.
The QD layer is centered within an optical microcavity with a reso-
nance chosen to enhance the interaction of the optical eld with the QD.
107

x28
i
n
p
i
Ohmic
contact
80nm Ti, 1m
aperture
QDs
Al
0.3
Ga
0.7
As AlAs GaAs
doping
top DBR
bottom DBR
focused
lasers
Figure 5.3: Schematic of the sample for single spin measurements. The sample consists
of a layer of GaAs interface uctuation QDs surrounded by distributed Bragg reectors
(DBRs) forming an optical cavity. The structure is doped and gated so as to control
the charging of the QDs and the QW.
See Section 4.8 for more discussion of cavity-enhanced Faraday rotation.
The front and back cavity mirrors are distributed Bragg reectors (DBRs)
composed of ve and 28 pairs of AlAs/Al
0.3
Ga
0.7
As /4 layers, respectively.
This asymmetrical design allows light to be injected into and emitted from
the cavity on the same side. The cavity was designed using a vertical
cavity simulator (called Vertical) to have a resonance aligned the QD
108
AlGaAs
AlGaAs
QDs
GaAs
(a)
(b)
Figure 5.4: (a) STM image of interface uctuations at a GaAs surface. Steps in color
indicate atomic steps in thickness. From D. Gammon et al., Phys. Rev. Lett. 76, 3005
(1996). (b) QDs form at the potential minima caused by localized islands of increased
QW thickness.
emission. However, due to some error in the growth, the cavity resonance
was somewhat lower in energy, though still with some overlap with the
QD emission. The reectivity of the sample at 10 K (Fig. 5.5 c) shows a
cavity resonance centered at 763.6 nm (1.624 eV) with a Q-factor of 120.
The Q of the cavity could be increased by adding more layers to the front
DBR, but increasing the distance from the QDs to the surface of the sam-
ple makes isolation of single QDs more dicult, and also complicates the
sample gating (discussed below and in Appendix A). Based on previous
measurements with similar cavities [59, 62], we expect the KR at the peak
of the cavity resonance to be enhanced by a factor of 15.
109



n
(
1
0
1
6
c
m
-
3
)
0
5
10
V
b
(V)
0 1
(b)






E
n
e
r
g
y

(
e
V
)
0
-1
-2
1
z (m)
0 0.5 1
V
b
= 1 V
V
b
= -1 V
(a)
QW
750 775
R
(nm)
1
0.8
(c)


Top DBR
Figure 5.5: Characterization of the single spin sample. (a) Calculation of the conduc-
tion (solid) and valence (dashed) band proles as a function of distance from the sample
surface (z = 0). As the bias voltage is changed from 1 V to -1 V (red and black lines),
the QW is lowered beneath the Fermi energy. (b) Calculated electron density in the
QW. The onset of charging occurs around V
b
= 0.5 V. (c) Reectivity of the sample at
T = 10 K showing a dip caused by the cavity resonance. The width of this dip indicates
a Q-factor of 120.
Additionally, the QDs are embedded in a diode-like structure, allowing
the charging of the QDs and the QW to be controlled with a bias voltage.
The back DBR is Si doped at n = 3 10
17
cm
3
until a depth of 85 nm
110
beneath the QW, which is followed by a 25 nm p-type region (p = 1.5
10
17
cm
3
), with the rest of the structure undoped. A portion of the sample
is etched down to the n-doped layer and an Ohmic Ni/GeAu contact is
deposited. On the unetched region, an 80-nm Ti layer forms a Schottky
contact with 1-m apertures fabricated by electron-beam lithography. This
layer serves as both a front gate and a shadow mask for isolating single
QDs.
The band prole for our structure, calculated with a 1-D self-consistent
Poisson-Schrodinger solver, is shown in Fig. 5.5 a. By applying a bias
across the structure, the conduction band minimum in the QW can be
made to plunge beneath the Fermi level, charging rst the QDs, then the
well itself [84, 85]. This structure is similar to that in Ref. [84] or [85],
with the addition of the thin p-doped layer. The purpose of this layer is
to provide an additional boost to the conduction band energy so that the
QW can be raised above the Fermi level at reasonable bias voltages. The
onset of this charging occurs around 0.5 V (Fig. 5.5 b) according to the
band-structure calculation.
111
5.3 Theory
As discussed several times above, the magneto-optical Kerr eect results
in a rotation of the plane of polarization of linearly polarized light with
energy E upon reection o the sample, and is analogous to the Faraday
eect for transmitted light. For both eects, the rotation angle is deter-
mined by the dierence of the dynamic dielectric response functions for
+
and

circularly polarized light, which are proportional to the interband


momentum matrix elements, P

c,v
=
c
[ p
x
i p
y
[
v
) , where
c
(
v
) is a
conduction (valence) band state [86, 36]. Due to the cavity, both reection
and transmission contribute to the measured polarization rotation. For
simplicity, we refer only to KR. For a single conduction-band energy level
in a QD containing a spin-up electron in a state

, optical transitions to
the spin-up state are Pauli-blocked, and the KR angle is then given by

K
(E) = CE

v
_
[P
+
,v
[
2
[P

,v
[
2
_
E E
0,v
(E E
0,v
)
2
+
2
v
, (5.2)
where C is a constant, and E
0,v
and
v
are the energy and linewidth of
the transition involving the valence band state [
v
), respectively. We focus
on a single transition in the sum in Eq. 5.2 and drop the index v. For
[[ E, where = E E
0
, we note that
K

1
, which
decays slower than the absorption line, (
2
) [36, 81]. Therefore, for
112
a suitable detuning, , KR can be detected while photon absorption is
strongly suppressed. This gives the KR measurement its non-destructive
property.
5.4 Experimental setup
A diagram of the experimental setup is shown in Fig. 5.6. Further details
are given in Appendix B. A cw Ti:Sapphire laser (Ti:Sa1 in the diagram)
to be used as the pump for KR or PL measurements is focused through
a mechanical chopper with frequency f
1
= 4.1 kHz. The intensity of the
laser is attenuated with a neutral density (N.D.) lter. This beam is passed
through a linear polarizer and a liquid crystal variable waveplate (LCVW).
For most measurements, the LCVW is set to quarter-wave to convert the
light to
+
or

circular polarization. A second cw Ti:Sapphire laser


(Ti:Sa2 in the diagram) used as the probe for KR measurements is focused
through a mechanical chopper with frequency f
2
= 21 Hz, and is also atten-
uated with an N.D. lter. This beam is passed through a linear polarizer,
and then made collinear with the pump beam using a beamsplitter. Both
beams reect o of a beamsplitter into a microscope objective. The ob-
jective is mounted on a motorized positioning stage for coarse positioning
in the x- and y-directions, and a piezoelectric stage for ne positioning
113
Ti:Sa 1
(cw pump)
Ti:Sa 2
(cw probe)
choppers
LCVW
polarizer
polarizer
scanning
objective
sample
He flow cryostat (10 K)
electro-
magnet
Fe yoke
longpass
filter
to spectrometer
(for PL)
to diode bridge
(for KR)
optional
mirror
f
1
f
2
V
A-B
lock-in1
f
ref
=f
2
lock-in2
f
ref
=f
1
pre-amp
V
A lock-in3
f
ref
=f
1
for
normalization
signal
Figure 5.6: Setup for single spin detection. Two continuous wave (cw) Ti:Sapphire
lasers provide the pump and probe for KR measurements. The lasers are focused onto
the sample, which sits in a cryostat at 10 K. An electromagnet provides a transverse
magnetic eld. The outgoing light is passed though a longpass lter to remove the pump
light, and is sent to a spectrometer for PL measurements, or to the diode bridge for KR
measurements.
in all three dimensions (scanning in x and y, plus focus). The objective
focuses both beams onto the sample (spot size 1 m), which sits inside
a liquid Helium ow cryostat. The cryostat itself is on a motorized stage
114
for coarse focus control. Additionally, the iron poles of an electromagnet
are positioned above and below the sample cold nger, to apply magnetic
elds up to 0.1 T.
For PL measurements, the probe beam is blocked, and only the pump is
focused onto the sample at T = 10 K to excite electron-hole pairs into the
continuum of states in the QW (E
pump
= 1.654 1.662 eV). The carriers
then relax into the QDs, and the subsequent PL is collected through the
same objective, and sent back through the beamsplitter. The outgoing
light from the beamsplitter is sent through a long-pass optical lter to
block the pump light. The PL is passed through the lter, and focused
through the entrance slit of a 1-m spectrometer. The PL spectrum is then
detected by a liquid-nitrogen-cooled CCD.
For Kerr rotation measurements, the pump and probe are focused on
the sample through the microscope objective onto one of the apertures.
The probe energy can be swept by means of a motor on the birefringent
lter in the laser cavity. The reected light is collected through the objec-
tive and passed through a long-pass optical lter to block the pump beam.
The rotation of the probe polarization is then detected by a balanced pho-
todiode bridge, as described in Section 2.5. The dierence channel of the
diode bridge is sent to a preamplier with 100 gain, and a bandpass lter
115
from 3 kHz to 10 kHz. (Since the pump is modulated by the chopper at
4.1 kHz and the probe beam is modulated at 21 Hz, the signal due to spins
injected by the pump, and measured by the probe will be at 410020 Hz.)
After the preamplier, the signal is sent to the rst lock-in amplier with
reference frequency f
1
and time constant 640 s. The output of this lock-in
is sent to a second lock-in amplier with reference frequency f
2
and a time
constant of one or two seconds. During this measurement time of several
seconds, the pump is repeatedly reinitializing the spin. In this sense, it is a
measurement of a single spin in a QD repeated many times, and averaged
in time. Finally, the pump polarization is switched between
+
and

with the liquid crystal variable waveplate, and a measurement of the rota-
tion angle is taken at each helicity. The dierence between these two values
yields the signal modulated at both the pump and probe frequencies, and
that depends on the sign of the pump helicity. Sweeping the probe energy,
with a measurement at each energy maps out the KR spectrum. Typical
pump and probe intensities are 15 W/cm
2
and 500 W/cm
2
, respectively.
116
5.5 Characterization: PL and Hanle
measurements
We begin by nding a good looking QD. Most of the apertures on the
sample show multiple QDs, but we look for one that has a strongly emitting
QD suciently separated in energy from the other QDs that it can be
individually addressed. The rest of the measurements below will be on the
same QD, unless otherwise noted. We next characterize this QD using PL
measurements, and verify that we can inject a single polarized spin into
the QD by reproducing the results of Ref. [65], in which the spin of a single
electron is measured using PL polarization.
In a typical single dot PL spectrum as a function of the applied bias
(Fig. 5.7), the sharp features (linewidth 100 eV) are characteristic of
single-dot PL [83], demonstrating the presence of only one QD within the
laser focus and within the measured energy window. Above 0.5 V a single
line is observed at 1.6297 eV which is attributed to recombination from
the negatively-charged exciton (trion, or X

) state. Below 0.5 V this line


persists faintly, and a bright line appears 3.6 meV higher in energy due
to the neutral exciton (X
0
) transition. The presence of the X

line at
V
b
< 0.5 V implies that occasionally a single electron is trapped in the
117


X
-





0 0.5 1 1.5 -0.5 -1 -1.5
-0.2
0.2
0
1.625
1.630
1.635
P
L

e
n
e
r
g
y

(
e
V
)
P
L

p
o
l
.
charged uncharged
X
0 X
-
XX
X
-
X
0
(a)
(b)
V
b
(V)
3.6 meV
Figure 5.7: Photoluminescence of a single QD vs. bias voltage, and polarization
thereof. (a) Single QD PL vs. bias voltage. The neutral exciton (X
0
), charged exciton
(X

) and biexciton (XX) lines are identied. (b) The degree of circular polarization
of the X
0
(black) and X

(red) PL lines as a function of bias voltage. The biexciton


PL is unpolarized.
dot, forming an X

when binding to an electron and a hole. In addition,


a faint line at 1.6292 eV is visible from radiative decay of the biexciton
(XX). These assignments of the observed lines are consistent with mea-
surements on similar structures [85, 65], and are further supported by the
linear dependence of the X

and X
0
lines, and the quadratic dependence
of the XX line on the excitation intensity, as shown in Fig. 5.8. Figure 5.9
illustrates these three optical transitions. In this QD we see no evidence
118





X
-
XX
X
0
slope=1.8
slope=1.2
slope=0.9
Pump power (W)
0.1 1 10
10
3
10
4
10
5
P
L

i
n
t
e
n
s
i
t
y

(
c
o
u
n
t
s
)
Figure 5.8: The intensity of the neutral exciton (X
0
), charged exciton (X

), and
biexciton (XX) PL lines as a function of pump power, on a log-log scale. The red lines
are ts to the low power, linear region of the plot. The slope of approximately 1 for
the X
0
and X

curves indicates that they are mainly one-photon processes, while the
slope of approximately 2 for the XX PL indicates that it is a two-photon process.
of a positively charged exciton.
With circularly polarized excitation, spin polarized electrons and heavy
holes can be pumped into the QD due to the optical selection rules of
the GaAs QW [65, 14] (see also Chapter 2). For the purposes of this
discussion, spin polarization parallel to the optically injected electron spin
polarization will be referred to as spin-up, and the opposite spin as
spin-down. Information about the spin polarization in the QD can be
gained from the polarization of the PL [14] (see Chapter 2). The circular
polarization of the PL is determined by switching the helicity of the pump
from
+
to

and measuring the intensity of the


+
-polarized PL, (I
+
119
E
X0
empty dot - exciton
E
X-
E
XX
single electron - trion exciton - biexciton
Figure 5.9: Lowest energy optical transitions in a quantum dot. In all three transitions,
an electron and hole are created (destroyed). (a) In the neutral exciton (X
0
) transition,
the QD begins (ends) empty. (b) In the charged exciton (X

) transition, the QD begins


(ends) with a single electron. (c) In the biexciton (XX) transition, the QD begins (ends)
with an electron and a hole.
and I

, respectively). The polarization of the PL is measured using a


second liquid crystal variable waveplate set to 1/4-wave followed by a
linear polarizer in the PL collection path. The polarization is then dened
as P = (I
+
I

)/(I
+
+ I

) and is shown for the X


0
and X

lines in
Fig. 5.7, in agreement with earlier results [85, 65].
The X
0
PL shows a small, but positive polarization over the entire
range of bias where X
0
PL is present. This reects the polarization of the
injected electrons and holes. The magnitude of the circular polarization
is most likely reduced from the polarization of the injected carriers due to
the anisotropy of the electron-hole exchange interaction in the QD. This
eect arises from elongation of the QDs along the [110] crystal axis [87],
causing electrons and holes to relax into states emitting linearly polarized
120
E
X-
, neg. pol.
(a)
(b)
Figure 5.10: Mechanisms for single spin initialization. (a) A single, optically injected,
spin-up electron relaxes into the QD. (b) The QD initially contains a single electron. An
optically injected electron and hole relax into the QD, forming the X

state. This state


rapidly ( 100 ps) decays by emitting (on average) a negatively circularly-polarized
photon, leaving behind a spin-up electron in the QD.
light.
The polarization of the X

line is determined by the hole spin, as the


two electrons in the trion form a spin-singlet state. In the uncharged regime
(V
b
< 0.5 V), the negative polarization of the X

PL indicates that the


heavy hole undergoes a spin-ip before recombination in most cases. Hole
spin-ips may occur either during energy relaxation in the QW [88] or by
121
an exchange-mediated electron-hole spin-ip [89]. Regardless of the hole
spin-ip process, after the recombination of the X

, the electron left in the


QD is polarized in the spin-up direction. In this way, as shown in Fig. 5.10,
both optical injection and trion recombination serve to pump lone spin-up
electrons into the QD.
When the dot is initially charged near V
b
= 0.5 V, the now dominant
X

line remains negatively polarized, resulting in continued pumping of


the spin-up state. As the electron density in the QW increases with higher
applied bias, the X

polarization becomes positive, as has been previously


observed [85, 65]. In this regime, electrically injected electrons swamp the
optically excited electrons, and the PL only depends on the spin of the
hole.
In a transverse applied magnetic eld, the electron spins precess, de-
polarizing the PL. This is the Hanle eect, described in Chapter 2. The
hole spins do not precess [90] because the heavy and light hole states are
split (by 20 meV in our sample [91]), leading to an eective heavy-hole
g-factor of zero in the plane of the QW. Hanle measurements on this dot
are summarized in Fig. 5.11. In the charged regime, at V
b
= 0.9 V, no
depolarization of the X

PL is observed, as expected for polarization due


to the hole spin. The case is markedly dierent at V
b
= 0.8 V, in the
122





X
0
, -0.8V
X
-
, 0.9V
X
-
, -0.8V
0 1 2 -1 -2
0
-0.2
0.2
0.4
Magnetic Field (kG)
P
L

p
o
l
a
r
i
z
a
t
i
o
n
T = 10 K
Figure 5.11: Single quantum dot Hanle measurements. At V
b
= 0.8 V, the charged
exciton (X

) PL shows a sharp Hanle peak (red) indicating a spin lifetime T

2
10 ns.
The exciton (X
0
) PL shows a much broader Hanle curve (black) due to the short
recombination time of the exciton state. At V
b
= 0.9 V, the X

PL polarization (blue)
shows no dependence on magnetic eld. The green curves are Lorentzian ts.
uncharged regime. Here, the (negatively-polarized) X

line is depolarized
with a half-width, B
1/2
= 80 G. With an estimated electron g-factor of
g
e
= 0.2 [65], B
1/2
= 80 G corresponds to a time-averaged transverse spin
lifetime T

2
= h/B
1/2
g
e

B
= 7 ns, where
B
is the Bohr magneton, and
h is the Planck constant. This sharp Hanle peak has been previously at-
tributed to the electron spin in the QD, prior to X

formation [65]. The


X
0
line shows a much broader peak (B
1/2
= 4.1 kG), with a small nar-
row component at low eld. The broad component is consistent with the
radiative lifetime of the exciton ( 50 ps) [67]. The narrow component
123
has a half-width, B
1/2
= 95 G, similar to the X

width. In fact, this nar-


row peak is expected if a lone electron in the dot can bind and recombine
with a subsequently injected hole. Similar features have been observed in
ensemble Hanle measurements in GaAs QWs [92].
To summarize these PL results, in the uncharged regime spin-polarized
excitons or electrons can be pumped into the dot. Both optical injection
and trion recombination serve to pump spin-up electrons. At high bias in
the charged regime (V
b
= 0.9 V) the PL polarization is due to the hole
spin, obscuring any information about the electron spin polarization. So
far, we have reproduced the results of Ref. [65]. To go further, a more
direct probe of the spin polarization is required.
5.6 Single spin Kerr rotation
The data in the top panel of Fig. 5.12 show the KR signal as a function
of probe energy for
+
and

pump helicity. Here, the applied bias is


V
b
= 0.2 V and the QD is in the uncharged regime. The PL at this
bias is also shown, with the X

and X
0
energies indicated by the dotted
lines. These energies coincide spectrally with two sharp features observed
in the KR data, which we will refer to as

and
0
, respectively. In the
bottom two panels of Fig. 5.12 the sum and dierence of the
+
and

124










1.62 1.63 1.64
Probe energy (eV)
K
e
r
r

R
o
t
a
t
i
o
n

(

r
a
d
)
-100
0
100
-200
-100
0
-100
0
100
X
-
X
0

+
+
-

+
-
-
V
b
=0.2 V
B=0, T=10 K
(a)

-
(b)
(c)
Figure 5.12: Single quantum dot Kerr rotation. (a) KR data as a function of probe
energy with
+
(blue) and

(red) polarized pump at V


b
= 0.2 V, in zero magnetic
eld. The PL at this bias is also shown (black), and the X

and X
0
energies are
indicated by the dotted lines. (b) The sum of the two curves shown in (a), representing
spin-independent eects, such as the spike (labeled
0
) at the X
0
energy. (c) The
dierence of the two curves shown in (a), representing the spin-dependent signal. The
feature

at the X

energy is attributed to single spin detection.


data is shown. The feature
0
at the X
0
energy clearly does not depend
on the sign of the injected spin and is similar to features seen in single
dot absorption measurements [82]. We attribute this peak to polarization-
dependent absorption in the QD. We focus here on the (
+

) data,
125
which represents KR due to the optically oriented spin polarization. The
feature

at the X

energy only appears in the dierence data, indicating


that it is due to the injected spin polarization, shown in Fig. 5.13 at four
dierent bias voltages. For all voltages, the

feature is centered at the


X

transition energy, indicated by the blue triangles. We can t these


data to Eq. 5.2 including only a single transition in the sum, on top of a
broad background (red lines, Fig. 5.13). From the free parameters in these
ts we determine the transition energy E
0
, amplitude A (dened as half
the dierence of the local maximum and minimum near E
0
), and width
of the

KR feature.
Fig. 5.15 shows E
0
compared to the energy of the X

PL line as a
function of the applied bias. Also shown for comparison is the energy of
the XX PL line. The two energies agree well and show the same quantum-
conned Stark shift. Only at the highest bias, where signicant broadening
sets in, do we observe a small anti-Stokes shift between E
0
and the X

PL
energy. This may be caused by interactions with electrons in the QW. (The
term anti-Stokes shift usually refers to a blueshift of emission, relative
to the absorption. Here we see a blueshift of emission, relative to the Kerr
rotation.) For a single electron spin in the QD ground state, the lowest
energy optical transition contributing in Eq. 5.2 is the X

transition (see
126












1.628 1.632
0
-10
-50
-100
0
0
-50
0
-50
K
e
r
r

R
o
t
a
t
i
o
n

(

r
a
d
)
Probe energy (eV)
fits
V
b
=1.1V
V
b
=0.7V
V
b
=0.2V
V
b
= -1V
Figure 5.13: Single spin KR feature as a function of bias voltage. The

KR
feature is present over a large range of bias voltage, though it broadens and decreases in
amplitude as the charging increases. Red lines are ts to an odd-Lorentzian plus a broad
(Gaussian) background. The blue triangles indicate the energy of the X

transition,
determined from PL measurements.
Fig. 5.9). Thus the

KR feature is due to the measurement of a single


electron spin in the QD. We have repeated this measurement on several
other QDs and observed the same

feature, also at the X

PL energy
127










-100
-50
0
50
-100
-50
0
-20
-10
10
0
K
R

(

r
a
d
)
X
-
PL
X
-
PL
X
-
PL
1.628 1.630 1.626 1.628 1.636 1.640
Probe energy (eV) Probe energy (eV) Probe energy (eV)
(a) (b) (c)
Figure 5.14: Single spin detection in other quantum dots. (a) KR and PL spectrum
from the QD weve been looking at so far. (b) KR and PL spectrum from the same
cavity sample as (a), but closer to the cavity resonance. (c) KR and PL spectrum from
the control sample without a cavity.
(shown in Fig. 5.14). The large, broad KR background is likely due to
transitions involving excited electron and hole states both in the QD and
in the QW, which are typically a few meV above the lowest transition [83].
If present, a KR feature due to the X
0
spin should appear centered
at the XX transition energy. The signal-to-noise in our measurement is
not high enough to conclusively identify such a feature. Despite the large
amplitude of the X
0
PL compared to the X

PL in the uncharged bias


regime ( 10 : 1), the short radiative lifetime of the X
0
state results in
a low steady-state X
0
population, and therefore low KR signal. To put
in some numbers, the exciton radiative lifetime in these QDs is typically
about 100 ps. On the other hand, the measured single spin lifetime (see
below) is around 10 ns. Even if a spin polarized exciton is pumped into the
128





X
-
PL
XXPL

-
KR
1.630
1.629
E
n
e
r
g
y

(
e
V
)
-1 0 1
V
b
(V)
Figure 5.15: Energy of single spin Kerr rotation feature compared to transition en-
ergies. The black points indicate the energy of the

feature as determined from the


ts as a function of bias voltage. The blue and red points show the energy of the X

and XX transitions determined from PL measurements.


QD 10 times more than a single electron spin, the time-averaged exciton
KR signal is 10 times lower than the single electron signal. Additionally
the anisotropic electron-hole exchange interaction mentioned above, might
also play a role in reducing the spin polarization of the exciton.
As mentioned above, we might expect to see an enhancement by a
factor of about 15 of the KR signal at the cavity resonance. However, the
QD we have been measuring has its X

transition at E
X
= 1.6297 eV,
whereas the cavity resonance is at E
cav
= 1.624 eV. The FWHM of the
cavity resonance is about 13 meV, so there is still some overlap of the cavity
resonance with the KR feature, though the enhancement may be less than
a factor of 15. We can attempt to quantify the eect of the cavity by
129
measuring other QDs with dierent detuning from the cavity resonance.
Unfortunately, we were only able to nd one good QD with energy lower
than the QD weve been looking at so far. However, it did have larger
signal. We have also fabricated a control sample in which the DBR layers
are replaced with only Al
0.3
Ga
0.7
As (same overall structure, but with no
cavity). We can still observe single electron KR in this sample, though the
signal is signicantly smaller. Figure 5.14 shows single spin KR spectra
from two QDs measured in the cavity sample, and one QD in the control
sample. The cavity appears to be enhancing the signal, though many more
QDs should be measured to be sure. One open question is that, though
the signal is lower in the control sample, it appears that the noise in the
control measurement is also lower.
An eect that may contribute to this is loss in the cavity as the light
spreads out away from the focus spot. The microscope objective focuses
the light in a cone yielding an angle of incidence up to 45

. As a result,
light in the cavity will spread out as it reects multiple times, away from
the focus spot. Of course, only the portion of light that is reemitted from
the cavity from the focus spot will be re-collected by the objective and
detected. Thus the signal will still be enhanced by the cavity, but the
total amount of light making it out will be reduced. In fact, we measure
130
a 10-fold decrease in the intensity of the probe light from going into the
objective to coming out. This reduction in the detected probe light could
certainly increase the noise in the measurement. Further work is needed
to determine what is going on here. One possibility would be to etch the
planar cavity into pillar structures to create a three dimensional cavity.
By applying a transverse magnetic eld B, we can monitor the depo-
larization of the single electron spin through the KR signal. In contrast to
the Hanle measurements described above, the KR probes the spin in the
QD directly and non-destructively, as opposed to being inferred from the
spin-dependent formation of the X

. The KR as a function of B is shown


for three dierent bias voltages (Fig. 5.16). At V
b
= 0.2 V, in the uncharged
regime, a narrow peak is observed with a half-width B
1/2
= 52 G, consis-
tent with the X

Hanle width measured in this regime. At V


b
= 0.7 V,
where the dot has charged, but the PL remains negatively polarized, we
measure a somewhat wider KR depolarization curve, with B
1/2
= 150 G.
When the QW is charged further, the spin lifetime decreases as shown at
V
b
= 1.1 V, with B
1/2
= 1.4 kG. Assuming an eective electron g-factor
of 0.2 [65], these half-widths correspond to transverse spin lifetimes of 11,
3.3, and 0.8 ns, respectively.
The electron spin depolarization curves measured at probe energies
131









0 1 -1
Magnetic field (kG)
0
1.0
0.5
0
1.0
0.5
0
1.0
0.5
K
e
r
r

R
o
t
a
t
i
o
n

(
n
o
r
m
a
l
i
z
e
d
)
V
b
=0.2 V
V
b
=0.7 V
V
b
=1.1 V
-0.3
(meV)
-2.7
5.0
T = 10 K
Figure 5.16: Single spin Kerr rotation Hanle measurements. KR measured at various
bias voltages and detunings as a function of magnetic eld. The typical Hanle depo-
larization curves conrm that we are measuring spin polarization. Clearly the curve
broadens at large charging, indicating a shorter spin lifetime.
detuned from the X

transition by an energy, are shown in the top


two panels of Fig. 5.16 for = 0.3 meV (at the maximum of the

feature), = 2.7 meV (in the low energy tail), and = +5.0 meV (on
the broad, high energy feature). The curves have been normalized by their
132
peak values, which vary with probe energy, but they show quite similar
lineshapes for a given bias. It is somewhat unclear why this is the case.
If the signal away from the

feature is due to electrons in the QW, one


might expect them to show dierent behavior in this Hanle measurement.
One possibility is that the same electron in the QD is being probed at
higher energy from transitions involving higher energy (QW) hole states.
It is also possible that there are dierences between the curves that are too
small to be identied given the signal-to-noise. The dynamics of the single
electron spin and the background signal will be discussed in more detail in
Chapter 6.
Fig. 5.17 a shows g
e
T

2
= h/B
1/2

B
as a function of the applied bias,
measured at a probe energy, E = 1.6288 eV, near the X

transition.
The dashed line indicates the onset of QD charging. The spin lifetime
is largest in the uncharged regime. Here, g
e
T

2
3 ns is consistent with
previous measurements [65] in which the spin dephasing is attributed to
the random, uctuating hyperne eld [13, 12]. (However, in Chapter 6,
doubt will be cast on this attribution.) As the dot and well are charged,
the electron spin lifetime decreases dramatically. This can be caused by
the increasingly rapid capture of a second electron in the dot, which forms
a spin-zero singlet state. If, for example, the bias voltage is set to add
133







-1 0 1
V
b
(V)
1
0.1
g
e
T
2
*
(
n
s
)
0
60
30
K
R

l
i
n
e
w
i
d
t
h
,

(
m
e
V
)
0
0.4
0.2
c
h
a
r
g
e
d
u
n
c
h
a
r
g
e
d
K
R

a
m
p
l
i
t
u
d
e
,
A
(

r
a
d
)
(a)
(b)
(c)
Figure 5.17: Analysis of single spin measurements as a function of bias voltage. (a)
The product of the g-factor and spin lifetime on a log scale as a function of bias voltage.
The dashed line indicates the onset of QD and QW charging. The red triangle shows the
value obtained from the PL Hanle measurements. (b) and (c) Amplitude and linewidth
of the

feature a as a function of bias voltage.


two electrons to the QD in equilibrium, then a pump-injected hole can be
trapped in the QD, and lead to trion emission. After this occurs, a single
spin will be in the dot. This spin will have a very short lifetime though,
since another electron will rapidly return into the QD. Note that this spin
134
polarization is not visible in the PL Hanle measurement described above
or in Ref. [65]. Also, as discussed below, spin ips with electrons in the
QW are likely to be a relevant mechanism in this regime.
The amplitude A of the

KR signal is shown as a function of bias volt-


age (Fig. 5.17 b). The amplitude decreases in the charged regime, reecting
the lower spin lifetime. We have argued above that spin-up electrons are
pumped into the QD in the uncharged regime. Therefore the constant sign
of the KR over the entire range of bias indicates spin-up polarization in the
charged regime as well. Contrary to this observed polarization, the posi-
tively polarized X

PL leaves a spin-down electron in the QD. However,


this electron interacts with the bath of electrons in the QW, which is, on
average, optically oriented in the spin-up direction. The predominant spin
in the QW may be transferred to the electron in the dot via a higher order
tunneling process [93] (essentially, simultaneous tunneling of an electron
out of and into the QD). The nite spin-up polarization measured up to
a large bias suggests that these electron-electron spin ips dominate over
the X

-mediated spin pumping in the charged regime.


As the bias increases above V
b
= 0.5 V, the width of the

KR feature,
grows by a factor of 6, shown in Fig. 5.17 c. A similar increase in linewidth
is seen in the X

PL in the charged regime. This provides further evidence


135
for an increased coupling of the QD to other electronic states as the charg-
ing increases. There are a number of unanswered questions regarding the
observed phenomena at large bias (large charging). The broadening of the
transition, the sign of the KR signal, and the shift of the

energy from
the PL X

energy might indicate that there is some interesting physics to


be studied in this regime.
At the beginning of this chapter, the KR measurement was described
as non-destructive because the Kerr rotation can be measured at a de-
tuning where the optical absorption is small. Of course, there is still some
nite probability of exciting carriers in the QD. Nevertheless, the KR mea-
surement itself is inherently non-destructive in that the mechanism does
not involve the excitation of additional electrons or holes. The fact that
the measurement works is proof that any unwanted destructive eects do
not swamp the non-destructive KR measurement. The ability to do a QND
measurement is of interest in regards to probing quantum eects in a solid
state system. In principle, Kerr rotation or Faraday rotation can satisfy
both of the two requirements for a QND measurement described above.
In order to demonstrate a QND measurement, both the non-destructive
nature and the back-action-evading nature must be sucient that the mea-
surement can be performed in a time less than the time for either of these
136
Figure 5.18: Schematic of photon-mediated coupling between two QDs via Faraday
rotation. From M. Leuenberger, Phys. Rev. B 73, 075312 (2006).
conditions to be violated. There is some hope that this limit could be
achieved. Long T
1
times have been observed for spins in optically accessi-
ble QDs (> 20 ms) [94]. In our measurements, we have been able to see
signal with time constants as low as 5 ms. Additionally, estimates of the
dominant back-action mechanisms indicate that this condition may be met
as well [95].
By probing a single electron in a QD through KR non-resonantly, we
demonstrate a direct measurement of the electron spin with minimal per-
turbation to the system. As a rst application, this method reveals infor-
mation about spin dynamics in single QDs, and constitutes a pathway
towards quantum non-demolition measurements and optically-mediated
entanglement of single spins in the solid state. An example of one such
scheme is shown in Fig. 5.18. The essential point of such proposals is
137
that if an electron spin is in a superposition of up and down, then
the Faraday eect results in a photon in a superposition of polarization
rotation CW and CCW. In this way, the quantum information has
been transferred from a spin to a photon. Much better optical cavities
would probably be needed to implement this scheme. (Fortunately, much
better cavities are certainly possible.) This scheme may also prove useful
for non-destructive measurements in a variety of solid-state qubits, such as
electrically-gated [70] or chemically-synthesized [59] QDs.
138
Chapter 6
Coherent Dynamics of a Single
Spin
6.1 Motivation
The ability to sequentially initialize, manipulate, and read out the state of
a qubit, such as an electron spin in a quantum dot (QD), is a requirement
in virtually any scheme for quantum information processing [96, 47, 97].
However, previous optical measurements of a single electron spin (including
those discussed in the previous chapter) have focused on time-averaged
detection, with the spin being initialized and read out continuously [98,
95, 65, 66, 69]. Here, we modify the measurement scheme of Chapter 5
to directly observe the coherent evolution of an electron spin in a single
QD, using time-resolved Kerr rotation (KR) spectroscopy. This all-optical,
139
non-destructive technique allows us to monitor the precession of the spin in
a superposition of Zeeman-split sublevels with nanosecond time resolution.
The data show an exponential decay of the spin polarization with time,
and directly reveal the g-factor and spin lifetime of the electron in the QD.
Furthermore, the observed spin dynamics provide a sensitive probe of the
local nuclear spin environment.
In Chapter 5, we describe a method for the time-averaged detection of
a single spin in a QD. By scanning the energy of a probe laser around the
lowest energy optical transition in a singly-charged QD (the X

transition),
a single electron spin produces a feature in the KR spectrum with the odd-
Lorentzian lineshape given by Eq. 5.2, centered at the energy of the X

transition, E
X
.
In the present work, as in Chapter 5, the electron is conned to a sin-
gle QD formed by monolayer uctuations at the interfaces of a gallium
arsenide (GaAs) quantum well (QW). The QD is embedded within a diode
structure, allowing controllable charging of the dot with a bias voltage [84].
Also, the QD is centered within an integrated optical cavity to enhance the
small, single spin KR signal [98]. See Fig. 5.3 for a schematic of the sam-
ple structure. With circularly polarized excitation, spin-polarized electrons
and holes are pumped into the QW, according to the selection rules gov-
140
erning interband transitions in GaAs [14]. One or more electrons and/or
holes then relax into the QD. By measuring the subsequent single QD pho-
toluminescence (PL), we determine the equilibrium charge state of the QD
as well as the energies of various interband optical transitions as a function
of bias voltage [98, 85]. The measurements described below are performed
at a bias voltage where the QD is nominally uncharged, and the optical ex-
citation injects one or more electrons or holes. This bias voltage is chosen
to maximize the KR signal (see Chapter 5). In this regime, the QD may
contain a single spin-polarized electron through the capture of an optically
injected electron, or spin-dependent X

decay [98]. Knowing the transition


energy E
X
from the PL measurements, we use the spectroscopic depen-
dence of the Kerr eect to isolate the dynamics of the single electron spin
from that of multiparticle complexes, such as charged or neutral excitons.
That is, we know that the KR feature at the X

energy is due to a single


electron spin in the QD, as opposed to, say, the exciton spin.
6.2 Experimental scheme
In previous work, only the steady-state spin polarization was measured,
concealing information about the evolution of the spin state in time. Here,
we use time delayed pump and probe pulses, shown schematically in Fig. 6.1,
141
Ti:Sa 2
(cw probe)
chopper
LCVW
polarizer
polarizer
scanning
objective
sample
He flow cryostat (10 K)
electro-
magnet
Fe yoke
longpass
filter
to spectrometer
(for PL)
to diode bridge
(for KR)
optional
mirror
f
1
f
2
V
A-B
lock-in1
f
ref
=f
2
lock-in2
f
ref
=f
1
pre-amp
V
A lock-in3
f
ref
=f
1
for
normalization
signal
~150 fs
monochromator
~ ps
Ti:Sa 1
(pulsed pump)
sync
EOM
pulse generator
5 . 1 ns
chopper
t
Figure 6.1: Setup for time-resolved single spin measurements. As compared to the
setup for continuous KR measurement, the pump laser is now mode-locked and passed
through a monochromator to narrow and select the energy of the pump pulses. Also,
the probe laser is passed though an electro-optic modulator (EOM) synchronized with
the pump laser, with electronically adjustable delay.
142
<200ps
1.5ns


Time (ns)
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)Pump
Probe
1.0
0.5
12
0
0 8 4
Figure 6.2: Temporal prole of the pump and probe pulses. The pump and probe
pulses are measured using a fast photodiode and oscilloscope. The bandwidth of the
measurement is limited to 2 GHz by the photodiode bandwidth.
to map out the coherent dynamics of the spin in the QD. The pump and
probe pulses, as detected by a fast photodiode, are shown in Fig. 6.2.
The setup used for time-resolved single spin measurements is similar to
the setup for the continuous measurement, but with a few important modi-
cations. Instead of the continuous pump laser, a mode-locked Ti:Sapphire
laser provides pump pulses with energy E
pump
= 1.653 eV, and duration
150 fs at a repetition period T
r
= 13.1 ns. The bandwidth of the spec-
trally broad pump pulses is narrowed to 1 meV by passing the pump
beam through a monochromator. The probe pulses are derived from the
same wavelength tunable CW Ti:Sapphire laser as in Chapter 5. However,
the probe laser is now passed through an electro-optic modulator (EOM),
allowing for electrical control of the probe pulse duration from CW down
to 1.5 ns. This technique yields short pulses while maintaining the narrow
143
linewidth and wavelength tunability of the probe laser. Also, it allows us
to adjust the pulse duration so as to maintain enough average power to
achieve good signal-to-noise, while keeping the instantaneous power low
enough to avoid unwanted non-linear eects. This is an important point.
As we will see in the next chapter, high intensity pulses near the E
X

energy induce a shift of the X

energy away from the laser energy. If the


probe pulses induce an appreciable shift of this type, the KR measurement
will not be possible.
The EOM is driven by an electrical pulse generator triggered by the
pump laser, allowing for electrical control of the time delay between the
pump and the probe pulses with picosecond precision. Additionally, in
measurements with pump-probe delay t > 13 ns the pump beam has also
been passed through an electro-optic pulse picker to increase the repetition
period of the pulse train to T
r
= 26.2 ns. Typical time-averaged pump and
probe intensities incident on the sample are 20 W/cm
2
and 200 W/cm
2
,
respectively. Note that this means that the instantaneous intensities are
larger by a factor of the repetition period over the pulse duration. More
details of the setup are given in Appendix B.
The microscope objective, sample, cryostat, and KR measurement ap-
paratus is the same as in the continuous single spin measurement in Chap-
144
ter 5. As in the time-resolved measurements described in Chapters 3 and
4, though the signal is averaged for several seconds (the spin is reinitial-
ized and probed millions of times), the stroboscopic pump/probe technique
allows measurement with high time resolution.
For a xed delay between the pump and the probe, the KR angle,

K
, is measured as a function of probe energy. At each point, the pump
excitation is switched between
+
and

polarization at a rate 1/t


switch
,
and the spin-dependent signal is obtained from the dierence in
K
at the
two helicities. The resulting KR spectrum is t to Eq. 5.2 plus a constant
vertical oset, y
0
. The amplitude,
0
, of the odd-Lorentzian is proportional
to the projection of the spin in the QD along the measurement axis. The
origin of the vertical oset, y
0
, is unknown but might be due to the broad
KR feature from free electron spins in the QW, as discussed below (also see
the discussion in Chapter 5). By repeating this measurement at various
pump-probe delays, the evolution of the spin state can be mapped out.
6.3 Theoretical model
When a magnetic eld is applied along the z-axis, transverse to the in-
jected spin (known as the Voigt geometry), the spin can be described in
the basis along the eld giving eigenstates [ ) and [ ), with eigenvalues
145
S
z
= h/2. The pump pulse initializes the spin at time t = 0 into the
superposition [(t = 0)) = ([) i[))/

2 , for

polarized excitation.
If isolated from its environment, the spin state then coherently evolves ac-
cording to [(t)) = (e
it/2
[) ie
it/2
[))/

2 , where h = g
B
B
z
is the
Zeeman splitting. When the probe arrives at time t = t, the spin state
is projected onto the y-axis, resulting in an average measured spin polar-
ization of S
y
(t)) = ( h/2) cos( t) . This picture has not included the
various environmental eects that cause spin decoherence and dephasing,
inevitably leading to a reduction of the measured spin polarization with
time.
In the simplest case, the evolution of the measured KR amplitude can
be described by an exponentially decaying cosine,

K
(t) = A (t) exp
_
t
T

2
_
cos( t), (6.1)
where A is the overall amplitude, (t) is the Heaviside step function, and
T

2
is the eective transverse spin lifetime (though this measurement elim-
inates ensemble averaging, the observed spin lifetime may be reduced from
the transverse spin lifetime, T
2
, by inhomogeneities that vary in time). To
model our data, we sum over the contributions from each pump pulse sep-
arated by the repetition period T
r
, and convolve (denoted ) with the
146

-
S
z
< 0
S
z
> 0

-
S
z
= 0 S
z
= 0
B
z
no misalignment misalignment
B
z
Figure 6.3: Illustration of spin misalignment leading to dynamic nuclear polarization.
Left: with the pump incident perpendicular to the magnetic eld, there is no component
of the injected spin along the magnetic eld. Right: A misalignment of the pump from
the perpendicular may yield a component of the spin along the magnetic eld, with
opposite sign for opposite pump helicity.
measured probe pulse shape, p(t),

0
(t) = p
_

n
(t nT
r
)
_
. (6.2)
This measurement technique is also sensitive to small nuclear spin po-
larizations (see Section 2.1 for discussion of nuclear spin polarization). Ide-
ally there should be no induced steady-state nuclear polarization in this
experimental geometry. Since the magnetic eld is applied perpendicular
to the direction of the spin, nuclear spins that are polarized by the elec-
tron spins precess around the applied eld, resulting in zero steady-state
polarization. However, for any misalignment of the pump laser from the
perpendicular, there is a projection of the spin along the magnetic eld, and
right (left) circularly polarized light induces a small dynamic nuclear polar-
ization (DNP) parallel (antiparallel) to the applied magnetic eld [14, 99].
147
This is illustrated in Fig. 6.3. Due to the hyperne interaction this acts on
the electron spin as an eective magnetic eld, increasing (decreasing) the
total magnetic eld, resulting in a dierent precession frequency for right
and left circularly polarized pump excitation. Since each data point is the
dierence of the KR signal with
+
and

polarized excitation, a small


deviation from perpendicular between the magnetic eld and the electron
spin yields a measured KR signal
(t) = A (t) exp
_
t
T

2
_
[cos((+)t) +cos(()t)], (6.3)
where = g
B
B
nuc
/ h is the frequency shift due to the steady-state eective
nuclear eld, B
nuc
.
6.4 Time-resolved single spin
measurements
The single spin KR amplitude as a function of delay, measured with a
3-ns duration probe pulse and a magnetic eld B = 491 G, is shown in
Fig. 6.4 a, exhibiting the expected oscillations due to the coherent evolution
described above. Fig. 6.4 b-f show a sequence of KR spectra at several
delays, and the ts from which the data in Fig. 6.4 a are obtained. In the
inset of Fig. 6.4 a the oset y
0
is shown, which oscillates with the same
148
K
R

(

r
a
d
)
-40
-30
-35


Probe energy (eV)
1.629 1.630
-25
t=0.3ns




Probe energy (eV)
1.629 1.630
K
R

(

r
a
d
)
-5
5
0
t=2.3ns
K
R

(

r
a
d
)



Probe energy (eV)
1.629 1.630
0
15
10
5
t=3.3ns



Probe energy (eV)
1.629 1.630
K
R

(

r
a
d
)
-5
10
0
5
t=5.2 ns



Probe energy (eV)
1.629 1.630
K
R

(

r
a
d
)
-10
0
-5
t=6.8 ns




0
(

r
a
d
)
20
10
0
-10
-20
Pump-probe delay (ns)
0 10 15 20 5
Delay (ns)
y
0
(

r
a
d
)
0 10
40
20
0
20
(a)
T = 10 K
B = 491 G
T
2
* = 8.4 3.5ns
0

x
S
z B
S
(b) (c) (d)
(e) (f)
Figure 6.4: Coherent precession of a single spin. (a) As the pump-probe delay is
varied, coherent precession is seen as oscillations in the measured projection of the spin.
The red curve is a t to the model, and the dotted line shows the same curve but without
the probe pulse convolution. The inset shows the oset, y
0
as a function of delay. The
error bars are the standard error in the ts to the KR spectra. (b-f) KR spectra at
increasing pump-probe delay showing amplitude oscillations in time. The red curves
are the ts from which the solid points in part (a) were obtained.
149
frequency as the single spin KR but decays with a shorter lifetime. This
behavior may be consistent with that of free electron spins in the QW,
previously investigated in time-averaged measurements [92]. Due to the
small connement energy of these QDs (several meV) relative to the QW,
one does not expect a signicant shift in the g-factor between the QDs and
the QW.
The solid line in Fig. 6.4 a is a t to Eq. 6.2, yielding = 0.98
0.02 GHz and T

2
= 8.4 3.5 ns. The dashed line shows Eq. 6.2 without
the probe pulse convolution, plotted with the same parameters for com-
parison. This data is taken with t
switch
= 1 s, and for now, the eects of
nuclear polarization are ignored. The eects of nuclear polarization will be
addressed below.
In Fig. 6.5 a the precession of the spin is shown at three dierent mag-
netic elds. As expected, the precession frequency increases with increasing
eld. The solid lines in Fig. 6.5 a are ts to Eq. 6.2, and the frequency
obtained from such ts is shown in Fig. 6.5 b as a function of magnetic eld.
A linear t to these data yields an electron g-factor of [g[ = 0.170.02, con-
sistent with the range of g-factors for these quantum dots found in previous
ensemble or time-averaged measurements [65, 100]. At zero magnetic eld,
as shown in Fig. 6.6, the spin lifetime is found to be T

2
= 10.9 0.5 ns.
150


0
(
a
.

u
.
)
T = 10 K
Pump-probe delay (ns)
B = 1195 G
B = 929 G
B = 491 G
0 2 4 6 8 -2 10 12
0
0
0
(a)



1000 0 500
0
1.0
2.0

(
G
H
z
)
(b)
|g| = 0.17 0.02
Magnetic field (G)
Figure 6.5: Single spin precession as a function of magnetic eld. (a) Spin precession
at three magnetic elds. The error bars are the standard error in the ts to the KR
spectra. (b) Precession frequency, , as a function of magnetic eld. The error bars
represent the standard deviation from repeated measurements. The linear t (red)
yields a g-factor of 0.17 0.02.
This value agrees with previous time-averaged [65] and ensemble [92, 100]
measurements where the relevant decay mechanism is often suggested to be
dephasing due to slow uctuations in the nuclear spin polarization. How-
ever, these polarization uctuations are not expected to result in a single
151
exponential decay of the electron spin [13, 12]. Roughly speaking, the
random distribution of nuclear polarizations yields a Gaussian-like decay
envelope. Using these non-Markovian models to t our data (see Fig. 6.6)
results in an increase of
2
to 24.3 from 3.8 for the t with a single expo-
nential decay, suggesting that other decay mechanisms than nuclear spin
uctuations might also be relevant in this case. In these QDs, the elec-
tronic level spacing of 1 meV [87] is of the same order as k
B
T for this
temperature range. Therefore, thermally-activated or phonon-mediated
processes [101, 102, 103, 104] which yield an exponential decay, might be
signicant in this regime.
In order to investigate the eects of nuclear polarization on the elec-
tron spin, we have varied the rate 1/t
switch
at which the pump helicity is
switched. Fig. 6.7 a and b shows the single spin KR signal as a function of
time with t
switch
= 1 s (as in Fig. 6.4), and t
switch
= 10 s, respectively, with
otherwise identical conditions. With larger t
switch
, the nuclear polarization
has time to build, as is evidenced in Fig. 6.7 b by the visible beating, as ex-
pected from Eq. 6.3. A t of the data in Fig. 6.7 a (t
switch
= 1 s) to Eq. 6.3
convolved with the probe pulse yields = 0 0.04 GHz and T

2
= 8.4 ns,
whereas for the data in Fig. 6.7 b (t
switch
= 10 s) = 0.14 0.02 GHz
and T

2
= 8.3 ns with a
2
of 2.3. For comparison, a single-frequency t
152



Pump-probe delay (ns)
B = 0 G
0 10 20
40
80
0
(a)

0
(

r
a
d
)
0 10 20
Pump-probe delay (ns)
(b)
Figure 6.6: Single spin dynamics at zero magnetic eld. (a) Decay of the spin polar-
ization with B = 0. The red line is a single exponential t convolved with the probe
pulse, giving a reduced
2
= 3.8. The error bars are the standard error from the ts
to the KR spectra. (b) The same data as (a) with a t to a model of nuclear spin
dephasing (blue), yielding an obviously poorer t, and a reduced
2
= 24.3. The a t
to the same model multiplied by an additional exponential decay factor, to model both
nuclear dephasing and other decoherence mechanisms, is shown in red. This t is also
signicantly worse than that in (a).
of the data in Fig. 6.7 to Eq. 6.2 yields a slightly larger
2
of 2.8, and
a signicantly shorter T

2
of 5.7 ns. It is unlikely that t
switch
would have
such an eect on the spin lifetime, and moreover, nuclear polarization is
the only eect in this system known to act on the spin with timescales
on the order of seconds. The inset in Fig. 6.7 shows the KR amplitude
as a function of t
switch
for a xed pump-probe delay. A t of the data to
Eq. 6.3 with saturating exponentially in t
switch
reveals a DNP saturation
time of 1.5 s similar to what has previously been found in these QDs [105].
153







0 5 10 15 20
-20
10
0
-10
10
-20
20
0
-10
Pump-probe delay (ns)

0
(

r
a
d
)
B = 491 G, T = 10 K
(a)
(b)
t
switch
= 1 s
t
switch
= 10 s

0
(

r
a
d
)
Switch time (s)
0
5 10
0
10
20
20
t = 6.7 ns
Figure 6.7: The eects of nuclear polarization on single spin precession. (a) Spin
precession with B = 491 G, and t
switch
= 1 s. The red curve is a t to the model
including DNP, yielding = 0 0.04 GHz and T

2
= 8.3 ns. (b) The same as (a), but
with t
switch
= 10 s. The t (red) now yields = 0.14 0.02 GHz and T

2
= 8.3 ns. The
inset shows the spin signal at a xed delay, t = 6.7 ns as a function of t
switch
. The
red curve is a t to the two-frequency model with saturating exponentially in t
switch.
(Specically, (t
switch
) =
0
[1 exp(t
switch
/t
n
)], where
0
is the maxi-
mum frequency shift, and t
n
is the DNP saturation time.) The nuclear
polarization seen in Fig. 6.7 corresponds to an eective magnetic eld of
100 G, or an electron spin splitting of 100 neV. For comparison, when the
magnetic eld and the electron spin are parallel, hyperne spin splittings
154
1000 times larger have been observed [105]. Given the typical size of
these QDs [87], the electron interacts with 10
5
nuclear spins. Since the
hyperne splitting here is smaller by a factor of 1000 than the maximum
observed splitting, the data shown in Fig. 6.7 represent the detection of at
most 10
5
/10
3
= 100 polarized nuclear spins.
These measurements constitute a noninvasive optical probe of the co-
herent evolution of a single electron spin state with nanosecond temporal
resolution, which is a key ingredient for many spin/photon-based quan-
tum information proposals [77, 79]. Furthermore, this technique provides
a sensitive probe of the dynamics of the spin, revealing information about
the spin coherence time and g-factor. Future work may exploit this ability
to further explore the relevant decoherence mechanisms and the electron-
nuclear spin interactions, and to observe the coherent manipulation of sin-
gle spins in real time.
155
Chapter 7
Ultrafast Manipulation of
Single Spin Coherence
7.1 Motivation and Background
Using ultrafast optical pulses to coherently manipulate the spin state of
an electron is a key ingredient in many proposals for solid-state quantum
information processing [106, 107, 108, 109, 110, 111]. Though electrical
control of single spins has been achieved [112], the nanosecond timescales
required for such manipulation limits the number of operations that can
be performed within the spin coherence time. In that work, single elec-
trons were conned to a gate-dened 2DEG QD, and the spin control was
achieved via spin resonance induced by a stripline deposited on the sample.
The speed of such a spin rotation is limited by the maximum attainable AC
156
S
y
= -1/2
S
y
= +1/2
J
y
= -3/2 J
y
= +3/2

-
Effective magnetic field, B
Stark
E
X-
E
X-
-
tipping
pulse
(a) (b)
Figure 7.1: Illustration of the optical Stark eect. (a) The relevant QD transitions
illustrated in the basis along the y-axis (the growth direction). The single electron
ground state is coupled optically to the negatively charged exciton (trion) state. For
a given circular polarization, the selection rules allow a non-zero matrix element only
for one such transition. (b) When an o-resonant circularly polarized optical eld is
applied, one spin state is shifted due to the OSE. This results in a spin splitting (eective
magnetic eld) for the single electron.
magnetic eld. In contrast, spin control via picosecond-scale optical pulses
yields an improvement of several orders of magnitude in the manipulation
time. In this chapter, we experimentally demonstrate such a scheme for a
single electron spin in a QD, monitoring the coherent evolution of the spin
state using time-resolved Kerr rotation spectroscopy. The spin is subjected
to an intense, o-resonant laser pulse, which induces a rotation of the spin
through angles up to radians on picosecond timescales.
The optical (or ac) Stark eect (OSE) was rst studied in atomic sys-
tems in the 1970s [113, 114, 115] and subsequently explored in bulk semi-
157
conductors and in quantum wells [116, 117, 118]). In recent years, the
OSE has been used to observe ensemble spin manipulation in a quantum
well [119], and to control orbital coherence in a QD [120]. Additionally,
other optical manipulation schemes have been explored on ensembles of
spins [121, 122]. Using time-dependent perturbation theory (similar to the
treatment in Section 2.4), it is found that an optical eld with intensity
I
tip
, detuned from an electronic transition by an energy , induces a shift
in the transition energy
E
D
2
I
tip

_
/
, (7.1)
where D is the dipole moment of the transition, and and are the
permittivity and permeability of the material [117]. Figure 7.1 shows the
relevant energy levels for the QD system considered here. The ground
state consists of a single electron in the lowest conduction band level, spin-
split by a small magnetic eld, B
z
. The lowest energy interband transition
is to the trion state consisting of two electrons in a singlet state and a
heavy hole. Due to the optical selection rules (see Chapter 2), the dipole
strength of this transition in the basis along the y-direction from the spin-
up (-down) ground state is zero for
+
(

) polarized light, as indicated


in the diagram. Therefore, for circularly polarized light, the OSE shifts
just one of the spin sublevels and produces a spin splitting in the ground
158
state which can be represented as an eective magnetic eld, B
Stark
, along
the light propagation direction. By using ultrafast laser pulses with high
instantaneous intensity to provide the Stark shift, large splittings can be
obtained to perform coherent manipulation of the spin within the duration
of the optical pulse (here, B
Stark
10 T). Note that this phenomenon can
also be described in terms of a stimulated Raman transition [108, 111], or
as an avoided crossing between excitons and photons [123].
7.2 Experimental scheme
As in Chapters 5 and 6, the sample consists of a layer of charge-tunable
GaAs interface QDs embedded in an optical cavity (see Fig. 5.3). A
schematic of the experimental setup is shown in Fig. 7.2. (Additional de-
tails of the setup are given in Appendix B.) This is similar to the setup of
the last two chapters, but again, with some signicant changes. In this case,
three synchronized, independently tunable optical pulse trains are focused
onto the sample: the pump, the probe, and the tipping pulse (TP). The
pump and tipping pulse are both derived (by means of a beamsplitter) from
the mode-locked Ti:sapphire laser generating a train of 150-fs-duration
pulses at a repetition rate of 76 MHz. The pump is circularly polarized,
and tuned to an energy E = 1.646 eV (FWHM 1 meV), thereby injecting
159
Ti:Sa 2
(cw probe)
chopper
LCVW
polarizer
polarizer
scanning
objective
sample
He flow cryostat (10 K)
electro-
magnet
Fe yoke
bandpass
filter
to spectrometer
(for PL)
to diode bridge
(for KR)
optional
mirror
f
1
f
2
V
A-B
lock-in1
f
ref
=f
2
lock-in2
f
ref
=f
1
pre-amp
V
A lock-in3
f
ref
=f
1
for
normalization
signal
~150 fs
~ ps
Ti:Sa 1
(pulsed pump/TP)
sync
EOM
pulse generator
5 . 1 ns
chopper
monochromators
~ 25 ps
polarizer
/4 plate
d
e
l
a
y
t
probe
t
tip
Figure 7.2: Setup for Stark tipping measurements. As compared to the setup for time-
resolved single spin measurements, we now split the pump laser to obtain the tipping
pulse (TP). The TP is passed through a monochromator, and circularly polarized. Also
the pump is now sent through a delay line to control the pump-TP delay.
160




K
R

(

r
a
d
)
Energy (eV)
20
-60
-140
1.624
1.628
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
XX
X
-
TP

0
Figure 7.3: Energy scales and relevant optical spectra for Stark tipping measurements.
The solid red line shows the charged exciton (X

) and biexciton (XX) PL lines from


the QD. The black circles show a single electron KR spectrum, with the odd-Lorentzian
t (solid black line) from which the KR amplitude
0
is obtained. The dashed red line
shows the spectrum of the tipping pulse at a detuning of = 4.4 meV.
spin-polarized electrons and holes into the continuum of states above the
QD [14]. One or more of these electrons or holes can then relax into the
QD. The circularly polarized TP (duration 30 ps, FWHM = 0.2 meV)
is tuned to an energy below the lowest QD transition (see Fig. 7.3) and is
used to induce the Stark shift. The relative time delay between the pump
pulse and the TP is controlled by a mechanical delay line in the pump
path.
As in Chapter 6, the probe pulse is generated by passing a narrow
linewidth continuous-wave laser through an electro-optic modulator syn-
chronized with the pump/TP laser. The resulting 1.5-ns-duration pulses
probe the spin in the QD through the magneto-optical Kerr eect [124].
161
This eect arises from the real part of the dielectric response function of
the QD, and results in a spin-dependent rotation of the polarization of the
linearly polarized probe upon reection o of the sample (see Section 2.4).
As the energy of this probe light is scanned across the QD transition energy,
E
X
, an odd-Lorentzian lineshape ( x/[1+x
2
]) centered at the transition
energy is seen in the Kerr rotation (KR) spectrum [98, 95]. By tting such
a curve (as shown in Fig. 7.3) we can extract the amplitude of this feature,

0
, which is proportional to the projection of the spin polarization in the
QD along the light propagation direction.
As before, the sample is mounted on the cold nger of a He ow cryo-
stat at the focus of a microscope objective, at a temperature T = 10 K.
The pump, probe and TP are focused and spatially overlapped on the sam-
ple with a spot size of 1 m. The temporal prole of the probe pulse
is measured using a 2 GHz bandwidth photodiode, and the TP prole is
measured using a streak camera and found to have a FWHM of 25 ps. Po-
larized spins are initially injected into higher energy states of the QW. A
single spin can be trapped in the QD either by relaxation of a single elec-
tron into the QD, or by spin-dependent formation and recombination of a
negatively charged trion. Because of the spectral energy sensitivity of our
measurement, we only probe the QD when it is singly charged - if the dot
162
contains, for example, an exciton, there is no contribution to the signal at
the X

energy. The reected light is collected through the same objective


and the rotation of the probe polarization is detected by a balanced photo-
diode bridge. Typical time-averaged pump and probe intensities incident
on the sample are 20 W/cm
2
and 200 W/cm
2
, respectively. The instan-
taneous probe intensity is thus 1700 W/cm
2
, which is suciently small
to prevent the probe from inducing an appreciable (non-spin-dependent)
optical Stark shift. For example, when the probe is detuned from the X

transition by 50 eV (approximately equal to the linewidth of the transi-


tion), the energy shift due to the probe is expected to be 20 eV. (This
is calculated from the measured values below.) Since the probe is linearly
polarized, this shift aects both spin states equally. This small shift may
alter the lineshape of the KR spectrum, but since the shift is less than the
linewidth of the transition, the odd-Lorentzian feature in the KR spectrum
is still visible.
In a typical measurement, the pump pulse arrives at t = 0 along the y-
axis (growth direction), and in some cases, a single spin-polarized electron
will relax into the quantum dot. For pump helicity

, this electron is (up


to a global phase) initially in the state [(t = 0)) = ([) i[))/

2, where
up and down are chosen as the basis along the external magnetic eld
163
B
z
. The spin then begins to coherently precess at the Larmor frequency
= g
B
B
z
/ h: [(t)) = (exp(it/2)[) i exp(it/2)[))/

2, where g
is the eective electron g-factor,
B
is the Bohr magneton, and h is the
reduced Planck constant. At time t = t
tip
, the TP arrives and generates
an additional spin splitting along the y-axis for the duration of the pulse.
During this time, the spin precesses about the total eective eld (which
is typically dominated by B
Stark
), and then continues to precess about the
static applied eld. The probe then measures the resulting projection of
the spin in the QD, S
y
) at t = t
probe
. This sequence is repeated at the
repetition frequency of the laser (76 MHz), and the signal is averaged for
several seconds for noise reduction. As described in Chapter 5, the pump
and probe are modulated using mechanical choppers, allowing for lock-
in detection to measure only spins that are injected by the pump. Also,
the pump is switched between
+
and

, with a measurement made at


each helicity. The spin signal is then taken as the dierence between these
values, eliminating any spurious signal from spins not generated by the
pump (e.g. phonon-assisted absorption from the TP [125]), or non-spin-
dependent rotation of the probe polarization.
To map out the coherent dynamics of the spin in the QD, KR spectra
are measured as a function of pump-probe delay. Figure 7.4 a shows a
164







E
probe
(eV)

K
(

r
a
d
)

K
(

r
a
d
)
0
10
-10
0
10
-10
D
e
l
a
y

(
n
s
)
2
0
4
1.6295 1.6300
E
probe
(eV)

K
(

r
a
d
)

K
(

r
a
d
)
0
-20
0
-20
1.6295 1.6300


D
e
l
a
y

(
n
s
)
2
0
4
(a)
(b)
t
probe
=0ns
t
probe
=2ns
t
probe
=0ns
t
probe
=2ns
No TP TP at t
tip
= 1.3 ns

K
(
a
.
u
.
)
1
-1







(i)
(ii)
(iv)
(iii)
(i)
(ii) (iv)
(iii)
TP
Figure 7.4: Single spin Kerr rotation vs. delay and probe energy with and without
the tipping pulse. (a) and (b) KR angle,
K
as a function of probe energy and pump-
probe delay. White (black) represents positive (negative) KR. In (a) no tipping pulse is
applied. In (b) the tipping pulse is applied at t
tip
= 1.3 ns, with intensity and detuning
set to cause a rotation. (i)-(iv) show linecuts from (a) and (b) illustrating the eect
of the TP. Specically, a reversal of the sign of the KR signal after the TP. The dashed
blue line indicates the X

energy.
plot of the KR angle,
K
as a function of probe energy and t
probe
with
an applied eld B
z
= 715 G and no TP. Horizontal line-cuts display the
dispersive lineshape centered at the transition energy E
X
, as shown in
Fig. 7.3. As t
probe
is swept along the vertical axis, the precession of the
165
spin can be seen as the oscillations in
K
. When the TP is applied at
t
tip
= 1.3 ns, as in Fig. 7.4 b, there is a signicant change in the KR
spectra. For t < t
tip
, the KR signal is essentially the same as in Fig. 7.4 a,
but for t > t
tip
the sign of the signal is reversed. This can be clearly seen in
the line-cuts shown in Fig. 7.4 i-iv. Line-cuts (i) and (iii) are both before
the TP and show the same behavior, whereas line-cut (iv) has the opposite
sign of line-cut (ii) as a result of the TP.
As described in Chapter 5, the quality factor of the cavity (Q = 120)
implies that the FWHM of the cavity resonance is about 13 meV. There-
fore, there is signicant overlap of both the probe and the tipping pulse
(TP) with the cavity resonance. As the linearly polarized probe eld builds
inside the cavity, the Kerr eect is enhanced, leading to an increase in the
KR signal. From similar measurements of Kerr or Faraday rotation in a
cavity [62, 59], we expect enhancement by a factor of 15 at the peak of
the resonance. We might expect a similar enhancement of the OSE as the
TP resonates in the cavity. However, as will be discussed further below,
the magnitude of the eect actually appears to be lower than expected
from the theory, possibly due to the spreading of the light in the planar
cavity.
166

B
(a) (b) (c)
Figure 7.5: Sequence of rotations in the Stark tipping model. (a) Before the TP,
the spin is initially slightly misaligned from the y-axis, and then precesses about the
magnetic eld, B
z
. (b) At t = t
tip
, the spin is instantaneously rotated about the y-
axis through an angle
tip
. (c) After the TP, the spin continues to precess about the
magnetic eld.
7.3 Theoretical model
It is convenient to understand the observed spin dynamics in the Bloch
sphere picture, described in Section 2.1. Here, the spin state is represented
as a vector (S
x
, S
y
, S
z
), where (0, 0, S
z
) represents the eigenstates [) and
[), and vectors with nonzero S
x
and S
y
represent coherent superpositions
of [) and [). In this picture, the dynamics of the spin can be calculated by
applying the appropriate sequence of rotation matrices to the initial state.
Figure 7.5 illustrates the sequence of rotations described by the model.
167
The initial spin state at t = 0 is taken to be

S
0
=
_
_
_
_
_
_
_
_
_
_
0
S
0,y
S
0,z
_
_
_
_
_
_
_
_
_
_
, (7.2)
where the initial component in the z-direction, S
0,z
is assumed to be small,
due to misalignment of the pump beam from normal incidence. Before
the tipping pulse arrives, the spin freely precesses around the applied eld.
Thus for t < t
tip
, we apply a rotation about the z-axis at frequency :

S(t) =
_
_
_
_
_
_
_
_
_
_
cos t sin t 0
sin t cos t 0
0 0 1
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
0
S
0,y
S
0,z
_
_
_
_
_
_
_
_
_
_
=
_
_
_
_
_
_
_
_
_
_
S
0,y
sin t
S
0,y
cos t
S
0,z
_
_
_
_
_
_
_
_
_
_
, t < t
tip
. (7.3)
At t = t
tip
we assume that the tipping pulse rotates the spin through an
angle
tip
about the y-axis. Since the duration of the TP is much less than

1
, the tipping is assumed to occur instantaneously.

S(t
tip
) =
_
_
_
_
_
_
_
_
_
_
cos 0 sin
0 1 0
sin 0 cos
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
S
0,y
sin t
S
0,y
cos t
S
0,z
_
_
_
_
_
_
_
_
_
_
, t < t
tip
. (7.4)
168
At t > t
tip
, this state

S(t
tip
) then precesses freely about the z-axis at the
Larmor frequency:

S(t > t
tip
) =
_
_
_
_
_
_
_
_
_
_
cos t

sin t

0
sin t

cos t

0
0 0 1
_
_
_
_
_
_
_
_
_
_

S(t
tip
) =
_
_
_
_
_
_
_
_
_
_
S
0,y
(cos
tip
sin t
tip
cos t

+ cos t
tip
sin t

) S
0,z
sin
tip
cos t

S
0,y
(cos
tip
sin t
tip
sin t

+ cos t
tip
cos t

) S
0,z
sin
tip
sin t

S
0,y
sin
tip
sin t
tip
+S
0,z
cos
tip
_
_
_
_
_
_
_
_
_
_
(7.5)
where t

= t t
tip
.
As can be seen from Eq. 7.5, the TP may result in a signicant non-
precessing component of the spin along the z-axis. Through the hyperne
interaction, electron spins can ip with nuclear spins. The component of
the nuclear spin along the external eld B
z
does not precess, and there-
fore can build over time in the process of dynamic nuclear polarization
(DNP) [14, 105] (see Section 2.1 and Chapter 6). As nuclear polarization
builds along the z-axis, it acts back on the electron spin as an eective mag-
netic eld B
n
. Thus in Eqs. 7.3 and 7.5, must be replaced by

= +
n
.

n
(= g
B
B
n
/ h) is proportional to the steady state nuclear polarization,
which in turn is proportional to S
z
. Thus equating
n
and S
z
with a
169
constant of proportionality , we have

n
= (S
0,y
sin
tip
sin
n
t
tip
+S
0,z
cos
tip
). (7.6)
Solving this equation numerically for
n
and substituting

into Eqs. 7.3


and 7.5, self-consistently yields the coherent spin dynamics of the system
as a function of time. In calculating the dynamic nuclear polarization,
we take the steady-state spin in the z-direction to be proportional to the
third component of Eq. 7.5. In principle, there is also a small component
present before the tipping pulse in Eq. 7.3, but we nd that it does not
substantially aect the results so we have neglected it for simplicity. The
magnitude of the nuclear polarization seen here is consistent with that
observed in Chapter 6 (Ref. [124]). This provides a nice conrmation of
this model, since in Chapter 6 the nuclear polarization was just included
phenomenologically as a shift in the precession frequency, while here the
misalignment is included in the model.
To model the results below, we include the nite spin coherence time,
T

2
, and a phenomenological I
tip
-dependent term to account for imperfect
delity of the spin rotations or other background eects, with a character-
istic scale, I
0
. Finally, we take the dierence between
+
and

pump
170
helicity yielding:

0
(t;

S
0
, , I
tip
, , T

2
, I
0
) =
_
S
y
(t;

S
0
) S
y
(t;

S
0
)
_
exp(t/T

2
) exp(I
tip
/I
0
).
(7.7)
The six parameters in the model are given explicitly in Eq. 7.7: the
initial spin direction

S
0
, the precession frequency , the intensity of the
TP I
tip
, the DNP eciency , the time-averaged transverse spin lifetime
T

2
, and the phenomenological factor I
0
. The term with I
0
is chosen to be
a decaying exponential, but other forms, such as a linear dependence work
just about equally well within the experimental uncertainty.
7.4 Ultrafast optical spin manipulation
Figure 7.6 a shows the time evolution of a single spin in a transverse mag-
netic eld, with no TP applied. Each data point is determined from the t
to a KR spectrum at a given pump-probe delay, as in Fig. 7.4. If we con-
volve Eqs. 7.3 and 7.5 with the measured prole of the probe pulse (shown
in Fig. 6.2), we can perform a least-squares t to this data and determine
various parameters in the model: , T

2
, and the eective eld from the
nuclear polarization, B
n
. The red curve in Fig. 7.6 a shows the result of
this t, and the dotted line is the corresponding plot of Eqs. 7.3 and 7.5
without the probe pulse convolution. As expected, the spin is initialized
171







Pump-probe delay (ns)
0 2 4 6 -2

0
(
a
.
u
.
)
0
0
0
TP off
t
tip
= 1.3 ns
t
tip
= 2.6 ns
(a)
(b)
(c)
T = 10 K
B
z
= 715 G
fit
model
model
x
z
y
B
S
v +
+ i
i

S
v
TP
TP

S
v
Figure 7.6: Coherent rotation of a single electron spin. (a) Coherent spin precession
with no TP. Error bars indicate the standard error in the ts to the KR spectra. The
red curve is a t to the model convolved with the probe pulse, and the dotted line is the
same, without the probe pulse convolution. The t yields parameters = 1.39 GHz,
T

2
= 5.5 ns, and B
n
= 68 G. (b) and (c) Same conditions as (a) but with the TP applied
at t
tip
= 1.3 ns and t
tip
= 2.6 ns, respectively, with intensity I
tip
= 4.7 10
5
W/cm
2
and detuning = 2.65 meV, to induce a 1.05 rotation. The red curves in (b) and
(c) are from the model, using parameters obtained elsewhere. The gray dashed line
highlights the change in sign of the spin precession in (b) as compared to (a) and (c).
The diagrams at the right illustrate the eect of the TP on the spin dynamics.
172
at t = 0, and then precesses freely about the applied eld.
It should be noted that in tting the time-resolved data shown in
Fig. 7.6 a, we obtain the eective nuclear eld B
n
, which is not explic-
itly a parameter in the model. In fact, B
n
depends on both the mis-
alignment S
0,z
/S
0,y
and the dynamic nuclear polarization (DNP) eciency
. However, the misalignment is determined from the data in Fig. 7.7
(S
0,z
/S
0,y
= 0.11). Using this number, is found to be 1.2, and all of
the parameters in the model are specied.
The data in Fig. 7.6 b and c show the same coherent spin dynamics of
Fig. 7.6 a, but with the TP applied at t = t
tip
. The intensity of the TP
is chosen to induce a 1.05 rotation about the y-axis, which is determined
as discussed below. In Fig. 7.6 b, the TP arrives at t
tip
= 1.3 ns, when
the projection of the spin is mainly along the x-axis. This component of
the spin is thus rotated by the TP through radians. The predicted
spin dynamics as given by Eqs. 7.3 and 7.5 is shown in the dotted red
line, and the same curve convolved with the probe pulse is given by the
solid red line. Note that this curve is not a t all of the parameters
are determined either in the t to Fig. 7.6 a, or as discussed below. Only
the overall amplitude of the curve has been normalized. Here, the spin
is initialized at t = 0, and as before, precesses freely until the arrival of
173
the TP. After the TP, the spin has been ipped and the resulting coherent
dynamics show a reversal in sign. This can be clearly seen by comparing
the sign of the measured signal at the position indicated by the dashed line
in Fig. 7.6.
Figure 7.6 c shows the spin dynamics again with the same parameters,
but with t
tip
= 2.6 ns. The spin at this delay will have only a small
projection in the x-z plane and the TP-induced rotation about the y-axis
will have only a small eect on the spin state. This expectation is borne out
in the data, where the spin dynamics show essentially the same behavior as
in the absence of the TP (Fig. 7.6 a). Again, the model yields qualitatively
the same behavior.
7.5 Further exploration and control
measurements
Further details of this spin manipulation can be investigated by varying the
TP intensity, I
tip
, and the detuning, of the TP from the QD transition
energy for a xed t
probe
and t
tip
as illustrated in Fig. 7.7 b. In Fig. 7.7 a, the
KR signal,
0
, as a function of I
tip
is shown at t
probe
= 2.5 ns with the TP
arriving at t
tip
= 1.3 ns, for three dierent values of . When I
tip
= 0, the
spin precesses undisturbed and yields a negative signal at t
probe
= 2.5 ns as
174






TP intensity (10
5
W/cm
2
)
0 2 4
0
-20
-40
40
20

0
a
t

t
p
r
o
b
e
=

2
.
5

n
s

(

r
a
d
)
= 1.64 meV
= 3.66 meV
= 5.26 meV
T = 10 K
B
z
= 715 G

(a)



t
probe
t
tip
Pump-probe delay (ns)
0 4

0
(
a
.
u
.
)
0
I
tip
= 0
(b)
Figure 7.7: Dependence of Stark tipping on tipping pulse intensity and detuning. (a)
Single spin KR amplitude,
0
as a function of TP intensity, I
tip
, at three detunings from
the X

transition. The tipping pulse arrives at t


tip
= 1.3 ns, and the probe is xed
at t
probe
= 2.5 ns, as illustrated in (b). The gray lines are ts to the model, varying
only one parameter, the strength of the OSE, , and misalignment S
0,z
/S
0,y
= 0.11,
and I
0
= 6.9 10
5
W/cm
2
. The tipping pulse intensity required for a rotation at
= 1.64 meV is indicated by the arrow.
in Fig. 7.6 a. As I
tip
is increased, the spin is coherently rotated through an
increasingly large angle, and the observed signal at t
probe
= 2.5 ns changes
sign and becomes positive, as in Fig. 7.6 b. Furthermore, the strength of
the OSE is expected to decrease linearly with the detuning , as seen in
Eq. 7.1. The gray lines in Fig. 7.7 a are plots of Eq. 7.5 with parameters
taken from the t in Fig. 7.6 a, and
tip
= I
tip
. From these curves, we
additionally obtain the phenomenological factor, I
0
= 6.9 10
5
W/cm
2
.
175



Detuning (meV)
0 2 4 6
I

(
1
0
5
W
/
c
m
2
)
0
5
10
Figure 7.8: Strength of the Stark eect as a function of detuning. The tipping pulse
intensity, I

required for a rotation as a function of detuning from the X

transition.
The red line is a linear t, through the origin.
This same value is used in all of the model curves shown. From this, we
can estimate the delity of a -rotation to be approximately 80%. Here,
the delity is calculated as the actual signal after a -rotation divided by
the dierence between the ideal signal with and without a -rotation. The
only parameter that is changed between the three curves in Fig. 7.7 a is
the strength of the OSE, . The TP intensity required for a -rotation,
I

= /, is shown in Fig. 7.8 as a function of detuning, displaying the


expected linear dependence.
The slope of the line in Fig. 7.8, 1.7 10
5
W/meVcm
2
, is a measure
of the observed strength of the optical Stark eect. If we assume a typical
dipole moment for this QD transition, D = 50 Debye, then from Eq. 7.1,
176
we expect this slope to be 1.3 10
4
W/meVcm
2
. The origin of this factor
of 13 reduction of the optical Stark shift is not clear, but may be due to loss
in the planar cavity as the light spreads away from the QD to ll the cavity
volume, as discussed in Chapter 5. This eect would not signicantly aect
the KR enhancement, since we only measure light that is re-emitted from
the cavity at the same point. For the OSE however, the magnitude of the
eect depends on the intensity of the light at the QD. In fact, we measure
a factor of 10 decrease in the intensity of light from going into to coming
out of the objective. Whether or not this loss is due to the spreading of
the light in the cavity, this is likely related to the smaller-than-expected
OSE. (This factor of 10 is a lower bound, since there is also some outgoing
light from back-reections o of the objective and other surfaces.)
The data in Fig. 7.7 most clearly show the eects of DNP on the ob-
served spin dynamics. In the absence of nuclear polarization, one would
expect the curves in Fig. 7.7 a to be cosinusoidal, crossing zero at an inten-
sity half that required for a rotation. DNP, however, which is maximal
when
tip
/2, distorts this ideal cosine form, as is well-described by
the model. Additionally, an eect of the misalignment of the initial spin
direction out of the x-y plane can be seen in the signal near I
tip
= 0. The
TP rst rotates the spin into the x-y plane, increasing the signal slightly,
177
TP intensity (10
5
W/cm
2
)











0 1 2 3 0 1 2 3
0
-20
-40
20
40

0
a
t

t
p
r
o
b
e
=

2
.
5

n
s

(

r
a
d
)
TP
-
, -I
0
TP
+
, I
0
TP
-
, I
0
=
TP
+
, I
0
=
TP
-
TP
+
(a) (b)
Figure 7.9: Comparison of the Stark eect with tipping pulses of opposite helicity. (b)
Single spin KR amplitude,
0
as a function of TP intensity for both helicities of the TP.
As above, the TP arrives at t
tip
= 1.3 ns and the probe is xed at t
probe
= 2.5 ns. (a)
Model curves of the data in (b), showing qualitative agreement. For the two helicities,
the sign of I
0
is reversed. The dotted lines show the model with I
0
= (perfect delity,
no background eects).
and then rotates it past the x-y plane as I
tip
is increased. This is conrmed
by reversing the helicity of the TP so that the spin rotation is in the same
direction as the misalignment, instead of against it (Fig. 7.9).
In order to conrm that we are manipulating a single electron spin,
we performed several additional control measurements. Even though the
modulation and lock-in detection should prevent the measurement of spin
coherence generated by the TP, we have repeated some of the measure-
ments with the helicity of the TP reversed. This should reverse the direc-
tion of the spin rotation, which, to a rst approximation, should not aect
178
the measured signal. However, the symmetry of the situation is broken
by the initial misalignment S
0,z
/S
0,y
, which yields a small but measurable
dierence, depending on whether the TP rotates the spin in the same or
opposite direction as the misalignment. The black points in Fig. 7.9 b show
the same type of data as in Fig. 7.7 a with TP detuning = 2.65 meV.
The red data is the same, but with the TP helicity reversed. Signicantly,
the observed behavior is qualitatively the same. If the observed eect were
an artifact due to TP-generated spin polarization, one would expect to see
a sign change with the TP helicity. The fact that there is no sign change
is strong evidence for the spin manipulation model described above.
The solid curves in Fig. 7.9 a are from the model, using the same
parameters as in the text. The only change is that for the red curve,
the parameter I
0
(accounting for TP-induced background eects or loss of
delity) was multiplied by (-1). For comparison, the dotted curves show
the prediction of the model with perfect delity (I
0
innite). Apparently,
the sign of I
0
depends on the helicity of the TP. This provides a clue as
to the relevant mechanisms, but more study is required for a complete
understanding. Regardless of the details of the mechanisms underlying
these background eects, the model captures the essential features of the
data, including the eects of the misalignment, eectively shifting the two
179






0 2 4 6 8
0
20
40
-20
Delay (ns)

0
(

r
a
d
)
TP
+
TP
-
TP off



0 2 4 6
Delay (ns)
circularly pol.
pump
linearly pol.
pump
(a) (b)
Figure 7.10: Comparison of measurement with circularly and linearly polarized pump.
(a) The usual time-resolved single spin measurement, but with linearly polarized pump.
Thus non-spin-polarized electrons are injected at t = 0, and in the red and blue data,
the TP is applied at t
tip
= 1.3 ns. (b) For comparison, the single spin dynamics with
circularly polarized pump, and no TP.
curves horizontally, most visible at low TP intensity.
In order to further isolate any unwanted eects of the TP, we performed
another control measurement in which the pump was linearly polarized, so
that electrons were still injected into the QD, but they were not spin po-
larized, shown in Fig. 7.10 a. We then performed the usual pump-probe
delay scan measurement as described above, with the circularly polarized
TP arriving at t
tip
= 1.3 ns. This measurement was performed with both
helicities of the TP, and with the TP blocked. The resulting signal was
barely above the noise oor, though there appeared to be some TP-induced
signal whose sign depended on the TP helicity with a magnitude of several
180
microradians. This is an order of magnitude smaller than the signal ob-
served with the circularly polarized pump (Fig. 7.10 b), and therefore sets
an upper bound on any spurious signal due to TP-generated spins.
Much of the deviation of the data from the model can be explained by
slow drift of experimental parameters during the measurement. In par-
ticular, the observed eects are very sensitive to the focus on the sample,
since the intensities of the pump, probe and TP all vary quadratically with
the focused spot size. Additional deviations may be due to the simplistic
description of the TP-induced background eects used here. For example,
in the case of phonon-assisted transitions to the trion state, one would ex-
pect the type of spin-selective decoherence described in Ref. [122]. Further
measurements of the background eects will be needed to determine their
cause, with the aim of increasing the delity of these single spin rotations.
In conclusion, we have demonstrated the ability to coherently rotate a
single electron spin through angles up to radians on picosecond timescales.
A simple model including interactions with nuclear spins reproduces the
observed electron spin dynamics with a single set of parameters for all of
the measurements. In principle, at most 200 single qubit ips could be per-
formed within the measured T

2
of 6 ns. However, by using shorter tipping
pulses and QDs with longer spin coherence times, this technique could be
181
extended to perform many more operations within the coherence time. A
mode-locked laser producing 100-fs-duration tipping pulses could poten-
tially exceed the threshold ( 10
4
operations) needed for proposed quan-
tum error correction schemes [47]. Additionally, the spin manipulation
demonstrated here may be used to obtain a spin echo [126], possibly ex-
tending the observed spin coherence time. These results represent progress
toward the implementation of scalable quantum information processing in
the solid state.
182
Chapter 8
Conclusion
In this dissertation, results have been presented focusing on measuring
and controlling coherent spin dynamics in semiconductor QDs. Faraday
and Kerr rotation have served as useful probes of the spin polarization.
Chapters 3 and 4 investigated novel layered nanocrystal QD structures.
First, layered nanocrystals were chemically synthesized with a low bandgap
shell, sandwiched between a high bandgap core and outer shell. These
structures were characterized using PL spectroscopy, and the spin dynam-
ics were probed using time-resolved Faraday rotation measurements. The
results revealed that these nanocrystals act analogously to a planar QW,
but in a spherical, nanoscale geometry. Specically, the QD energy levels
and g-factors shift with the shell thickness, in agreement with calculations.
This structure can be taken further by adding two coupled regions in
183
one nanocrystal. With structures consisting of a low bandgap core and
shell, separated by a high bandgap barrier, electrons and holes are conned
to either the core or shell region. As with single shells, PL and Faraday
rotation measurements indicate that the energy levels and g-factors depend
on the core and shell dimensions. Furthermore, with appropriate selection
of the pump and probe energies, the core and shell states can be initialized
and read-out selectively.
These types of QD structures are interesting because of their simplicity
of fabrication, and tremendous exibility in design. However, this simplic-
ity and exibility also leads to complications such as surface defects, and
unknown factors in the environment of the QDs. In Chapters 5, 6 and
7, we switch to MBE-grown interface uctuation QDs which have fewer
tunable properties but are more well-behaved than nanocrystal QDs.
In Chapter 5, Kerr rotation was employed as a non-destructive, optical
probe of single spins in a QD. By measuring the KR spectrum, the expected
signature was identied at the expected energy from a single electron spin
in the QD. Measurements of the spin in a magnetic eld showed the ex-
pected depolarization of the spin, and yielded information about the spin
lifetime and spin decoherence mechanisms.
The single spin KR technique was then extended into the time domain
184
(Chapter 6), using pulsed pump and probe lasers. In this measurement,
we probe the coherent dynamics of the spin in the QD with nanosecond
temporal resolution. This allows us to determine the g-factor and spin
lifetime of the electron, and also gain further insight into the relevant spin
decoherence mechanisms. Additionally, we observe eects of nuclear spin
polarization on the spin dynamics, making this measurement a sensitive
probe of the local nuclear spin environment.
Finally, in Chapter 7, we demonstrate ultrafast coherent manipulation
of the spin in the QD, observing the resulting dynamics using the single
spin KR measurement. Here, a third pulse is applied to the QD, detuned
from any optical transitions to generate an eective magnetic eld via the
optical Stark eect. This large, transient eective magnetic eld allows the
spin to be coherently rotated through angles up to radians, on picosecond
timescales.
These results have touched on initialization, manipulation, and detec-
tion of spin coherence in semiconductors. These are three of the key re-
quirements for applications in both spintronics and quantum information
processing. Together with the exibility of semiconductor technology, these
results provide some hope that we might one day fulll the promise of spin-
based technology. In the words of Martin Luther King, Jr., Hes allowed
185
me to go to the mountain. And Ive looked over, and Ive seen the promised
land! I may not get there with you, but I want you to know tonight that
we as a people will get to the promised land.
186
Appendix A
Sample structure and
processing
The cavity structure for the single spin measurements was designed using
the Vertical simulation software package. Initially, samples were fabricated
with higher-Q cavities, with more AlAs/AlGaAs layers in the top DBR.
However, this made the QD charging and the isolation of single QDs more
dicult. Both of these eects are because with more layers in the top
DBR, there is more distance from the surface of the sample to the QD
layer. Looking at Fig. 5.5 a, with a thicker top DBR, there is a larger
lever arm raising and lowering the QW with respect to the Fermi energy.
Thus large voltages would be necessary to change the charging of the QW.
Also, with a thicker top DBR, the 1 m diameter apertures at the surface
are separated from the QDs by something like 2 m, which doesnt help
187
in terms of isolating single QDs. So we have compromised by making a
lower-Q cavity with a thinner top layer. However, the cavity design has
been made compatible with a three-dimensionally conned pillar design in
which Q-factors as high as 48,000 have been achieved [127]. This provides
an interesting future avenue to pursue.
The doping scheme for the sample was designed using the self-consistent
Schrodinger-Poisson solver, 1d Poisson, made by Gregory Snider. The
structure is based on previous designs [84], but the presence of the cavity
adds an extra complication. The QW must be placed near the back contact
to allow electrons to tunnel in and out. This is ne in the structure in Ref.
[84], but with the large lever arm due to the front DBR, an unrealistic
voltage would need to be applied to pull the QW above the Fermi energy.
To solve this problem, a thin p-doped region is added, as shown in Fig.
5.5 a. This forms a p-n junction, which has the eect of pulling the bands
up in the region of the this junction, positioning the QW minimum near
the Fermi energy.
The growth of the IFQD samples was carried out by N. G. Stoltz. Care
must be taken to calibrate the growth rates accurately. Small deviations
in the thickness of cavity layer shift the resonance away from the desired
wavelength. (A 1% error in the thickness yields a 1% error in the resonance
188
wavelength, so if the target is 750 nm, instead you would get 758 nm - a shift
greater than the width of the resonance.) The samples were grown on 2 or
3 inch diameter semi-insulating GaAs substrates. Due to inhomogeneity
of the beam ux over the wafer, the cavity resonance varied across the
wafer. Using a simple reectivity setup, we could map out the resonance
wavelength in real time to select the piece of the wafer that we wanted to
use.
A piece of the wafer with the desired cavity resonance was then cleaved
(typically 6 8 mm), and processed as follows:
1. Mask half the sample with AZ4110 photoresist.
2. Etch with 80:8:1 H
2
O:H
2
O
2
:H
2
SO
4
down to back n-doped layer (
8 nm/s).
3. Remove resist, and apply AZ4251. Negative expose and develop two
square areas on the etched region, for back contacts.
4. E-beam evaporate Ni/Ge/Au/Ni/Au (5/25/65/20/200 nm) onto
the sample, and lifto (in acetone, usually with some sonication to help it
along).
5. Anneal in forming gas, using the strip annealer at 420

C for 1 minute.
Check that the two back contacts are now Ohmic (linear I-V). If not, anneal
more.
189
6. Apply maN-2403 resist for negative e-beam exposure.
7. Write the desired pattern of apertures using e-beam lithography (I
have used both a now-decommissioned JEOL system, and an FEI Sirion
system for this step). The pattern is typically 5 5 arrays of circular
apertures with diameters ranging from 200 to 2000 nm.
8. Develop in CD-26 for 1 minute, leaving circles of resist where the
apertures will be.
9. Evaporate 80 nm of Ti over the whole sample and lift-o the aper-
tures, resulting in the sample covered entirely in Ti, minus the apertures.
Lift-o is done using 1165 stripper at 80

C, with a little sonication if it


looks like its not all coming o.
10. Dene the front contact in photoresist over the apertures.
11. Etch in 100:1 H
2
O:HF for about 30 s, until the Ti has disappeared
everywhere but the front contact, and the underlying sample becomes vis-
ible.
12. Remove the resist, and now the sample is done. Check the I-V to
make sure it shows rectifying diode-like behavior.
The sample was mounted on the cold nger of the cryostat using con-
ductive silver paint, and wires were attached to the contacts by pressing
the wires into Indium.
190
Appendix B
Details of single spin detection
and manipulation
Figure B.1 shows again the setup for single spin measurement and manip-
ulation. Note that this diagram is not to scale, and leaves out additional
components such as lens pairs for collimating the beam and changing the
beam diameter, mirrors for additional beam steering, and lters for attenu-
ating the beams. Also, the order of some components has been rearranged
in the diagram for clarity when it doesnt aect the functionality. Here, I
will describe the setup in more detail.
191
Ti:Sa 2
(cw probe)
chopper
LCVW
polarizer
polarizer
scanning
objective
sample
He flow cryostat (10 K)
electro-
magnet
Fe yoke
bandpass
filter
to spectrometer
(for PL)
to diode bridge
(for KR)
optional
mirror
f
1
f
2
V
A-B
lock-in1
f
ref
=f
2
lock-in2
f
ref
=f
1
pre-amp
V
A lock-in3
f
ref
=f
1
for
normalization
signal
~150 fs
~ ps
Ti:Sa 1
(pulsed pump/TP)
sync
EOM
pulse generator
5 . 1 ns
chopper
monochromators
~ 25 ps
polarizer
/4 plate
d
e
l
a
y
t
probe
t
tip
Figure B.1: Setup for single spin measurement and control.
192
B.1 Optical path
The laser used for the pump and tipping pulses is a Coherent Mira Ti:sapphire
laser which is operated in CW for the time-averaged measurements (Chap-
ter 5) and is mode-locked for the time-resolved measurements (Chapters 6
and 7). The probe laser is a Coherent 890 Ti:sapphire ring laser, which out-
puts a tunable, narrow linewidth CW beam. The wavelength of the probe
laser is controlled by means of a stepper motor driving the birefringent
lter in the laser cavity.
For the time-resolved measurements, the probe is converted from CW to
pulses by means of an electro-optic modulator (EOM), purchased from EO-
Space, Inc. Light is coupled in and out of the EOM by means of a single-
mode, polarization-maintaining ber. A typical extinction ratio for the
EOM was around 50:1. The EOM is driven using a pulse generator (Agilent
81110A) with minimum pulse width of 1.5 ns and electrically controllable
delay with picosecond resolution. The pulse generator is triggered by the
output of the photodiode in the pump laser cavity. As was mentioned in
Chapter 6, we cannot use a mode-locked laser for the probe pulse (though
that was our rst idea). It is virtually impossible to get a pulse from a
mode-locked laser to be longer than about 100 ps. In order to maintain the
same total power to the detector, this means that the instantaneous power
193
in the pulses would be signicantly higher than in the present case. As
was discussed in Chapter 7, this would lead to a signicant optical Stark
shift as the probe laser approaches the transition. The eect is that the
QD resonance avoids the probe laser, making the range of small detuning
inaccessible.
The monochromator in the pump path is a Mini-chrom spectrometer
from Optometrics LLC, with manual control of the grating and path length
of 74 mm. Here, we just set the grating to the desired wavelength, and then
never touch it. The delay line in the pump path consists of a retroector
mounted on a sliding cart, which is driven by a stepper motor. This delay
line is used to control the time delay between the pump pulse and tipping
pulse.
The monochromator in the tipping pulse path is an Acton SP2150i,
with a path length of 0.15 m. The grating here is controlled via GPIB.
The three beams are all overlapped and sent through the microscope
objective, a Nikon L-plan SLWD with a numerical aperture of 0.7 and
working distance of 6 mm. The microscope objective is mounted on a
piezo-electric-actuated stage (P-563 from Physik Instrumente) with 300 m
range in x, y, and z, and nanometer positioning accuracy. The piezo stage
is in turn mounted on a PM500 motorized nanomover stage from Newport,
194
for coarse positioning in x, and y (where z is the focus). The cryostat is
mounted on a stepper-motor-driven translation stage which allows coarse
focus positioning, as well as allowing the cryostat to be pulled back for
putting in and taking out the sample.
The temperature of the sample is monitored by a temperature sensor
mounted at the end of the cold nger, near the sample. The bias voltage
is applied to the sample by means of a Yokogawa 7651 DC power supply.
In the outgoing light path, a lter is used to block the pump light,
and in the Stark tipping measurements, the tipping pulse. When the TP
was not present, a long-pass optical lter from Omega Filters was used. In
order to block both the pump and TP, while passing the probe, a bandpass
interference lter was used from CVI optics. The angle of the lter was
adjusted to optimize the attenuation of the pump and TP.
For PL measurements, the light was free-space coupled into an Ac-
ton AM510 1-m spectrometer, and detected using a Princeton Instruments
liquid-nitrogen-cooled CCD. For KR measurements, the probe was sent to
a diode bridge setup as shown in Fig. 2.12. In this case, the angle of
the half-wave plate was adjusted via a stepper motor to keep the diode
bridge balanced (see below for more details). The diode bridge circuit
was home-built and based on the design considerations outlined in Ref.
195
[16]. The diode bridge had outputs for each photodiode, A and B, and
the dierence, A-B. The dierence signal was sent to a Stanford Research
preamplier with gain of 100 and bandpass lter from 3 kHz to 10 kHz.
The signal then went to two lock-in ampliers (EG&G 726X models), used
as described above. Additionally, the A channel alone was sent to another
lock-in amplier with reference frequency at the probe chopper frequency
(f
1
) to normalize the KR signal, and to monitor the reectivity for posi-
tioning purposes (see below for more details).
B.2 Measurement control scheme
The measurement of the KR spectra was controlled by a Labview routine
that was set up as follows:
1. The stepper motor on the ring laser sets the initial probe energy.
2. The diode bridge is automatically balanced (see below).
3. Every n
th
iteration, the automatic positioning routine is run (see
below).
4. The pump helicity is set to
+
(using the liquid crystal variable
waveplate).
5. Wait for a time t
switch
.
6. A value from each of the lock-ins is stored.
196
7. The pump helicity is switched to

.
8. Wait for a time t
switch
.
9. Take another value from each lock-in.
10. Increment the probe energy and if the end of the scan has not been
reached, return to step 2.
11. Save the results.
The above sequence is for taking a single KR spectrum. This procedure
is then incorporated into other routines that take a series of KR spectra,
incrementing some other parameter, such as pump-probe delay, or tipping
pulse intensity. In these longer scans, the routine to optimize the focus and
zero the electro-optic modulator (see below) are run in between successive
KR spectra.
B.3 Odds and ends
There are a few routines that were used to keep the system stable over long
periods of time, so that long delay scans ( 10 hours) could be automated.
A signicant source of error in the single spin measurements is slow drift
of the sample relative to the laser spot. (For a 1 m diameter spot, even
100 nm of drift is a signicant error.) To compensate for this, the aper-
ture is automatically repositioned every so often. This is accomplished by
197
scanning the laser in a horizontal and vertical sweep over the aperture and
measuring the reectivity using the A channel of the diode bridge. The
resulting curves are then t to a Gaussian function to accurately determine
the position of the center of the aperture.
Additionally, drift of the focus position is also a signicant issue. To
compensate for this drift, the same reectivity measurements are performed
as for the positioning routine. Here, however, the measurement is repeated
at a number of dierent focus positions and the width of the t curves
is plotted versus focus position. Having previously determined the width
when the laser is optimally focused, we use a linear t of these values to
refocus the lasers.
For optimal operation of the diode bridge, the amount of light going
into the A and B photodiodes should be as close to equal as possible (for
common-mode rejection). This is accomplished by rotating the half-wave
plate to equalize the two channels. However, as the probe wavelength is
scanned the balanced position of the waveplate changes slightly, possibly
due to birefringence in the optics in the setup. For this reason, the wave-
plate is automatically rotated using a stepper motor to rebalance the diode
bridge. The balance of the bridge is detected by sending the dierence out-
put (after the preamp and lter) to another lock-in amplier at the probe
198
chopper frequency (f
1
). Even though this frequency is ltered out, there
is still a series of spikes in the signal at this frequency, proportional to the
unbalanced-ness of the bridge. The output of this lock-in is then zeroed by
using a simple search algorithm to move to stepper motor on the half-wave
plate.
One more source of long-term drift is in the electro-optic modulator.
The modulator is used by applying a DC bias voltage to completely null the
output of the modulator. Then the pulse generator supplies an additional
voltage to make the modulator transmissive. Over time however, the DC
bias needed to null the signal drifts (possibly due to charging eects in the
modulator). So every so often, the pulse generator is turned o, and the
DC bias voltage is scanned in a range around the present voltage, while
monitoring the A channel of the diode bridge. The resulting curve (which
has a minimum close to zero where the modulator is nulled) is t to a
Gaussian function to determine the new optimum bias voltage.
199
Appendix C
Other theoretical views of
Faraday rotation
Though this dissertation focuses on experimental results, it is useful to
have some feeling for the theory underlying these phenomena. In Sec-
tion 2.4, the origin of the Faraday rotation eect was outlined in terms of
time-dependent perturbation theory for an electron subjected to a classical
radiation vector potential, A. In this appendix, I will describe two other
ways of thinking about Faraday rotation. There will be no rigorous calcu-
lations, just a sketch of the theoretical description to hopefully give some
additional intuition regarding these phenomena.
One way of seeing the origin of the Faraday rotation eect is to con-
sider the (classical) coecient of absorption and index of refraction of a
material [15].
200
a
b
s
o
r
p
t
i
o
n
i
n
d
e
x

o
f

r
e
f
r
a
c
t
i
o
n
photon energy

Figure C.1: Top: Spin-dependent absorption for left and right circularly polarized
light. Bottom: Corresponding indices of refraction, and the resulting Faraday rotation
spectrum.
The real and imaginary parts of an analytic function () =
1
() +
i
2
() are related by the Kramers-Kronig relations:

1
() =
2

T
_

0
d

2
(

2
() =
2

T
_

0
d


1
(

2
,
(C.1)
where T indicates the principal value.
201
The dielectric response function is such a function, whose real and imag-
inary parts yield the index of refraction and absorption spectra. The ab-
sorption spectrum near the band-edge of a semiconductor, shown schemat-
ically in the top panel of Fig. C.1, yields a peak in the index of refraction
at the energy of the band edge. In the gure, the band-edge energies are
shown to be dierent for oppositely polarized light, which can be caused by
Pauli blocking from a population of occupied electron states of a particu-
lar spin. The indices of refraction corresponding to the absorption spectra
(via Eq. C.1) are shown in the bottom panel of Fig. C.1. As discussed in
Section 2.4, the Faraday rotation angle is given by the dierence in the
index of refraction for right and left circularly polarized light, shown in
black. Note that this curve is similar to the one in Fig. 2.10 b.
An alternative way of looking at Faraday rotation is in a picture where
both the electron and the photons are considered to be quantized [128].
Here, the eect can be thought of as forward scattering of the photons
o of the electron. In the second-quantized picture, photons are described
in terms of creation and annihilation operators: a

xk
and a
xk
, respectively,
where k labels the photon momentum and x indicates linear polarization
along x. The creation operator for a circularly polarized photon would
then be a

,k
= a

xk
ia

yk
for

polarization.
202
The interaction between the electron and the photon is given by the
Hamiltonian H
int
= p A/m, as in Section 2.4. The interaction Hamil-
tonian can be written down for the case of a dipole transition with energy
and dipole moment D (for example, a heavy hole to conduction band
transition), and simplied by only considering the o-resonant parts. The
form of this Hamiltonian can be guessed from the symmetry considerations
discussed above. Namely, a photon of a given helicity will only interact
with a transition to one spin state. Including only non-resonant, forward
scattering the interaction Hamiltonian is given by
H
S
= 2

kk

+,k
a
+,k
[ ) [ +a

,k
a
,k
[)[
=
z

kk

+,k
a
+,k
a

,k
a
,k
,
(C.2)
where the sum is over all photon momenta, and the prefactor can be cal-
culated to be = D
2
h/2V . Here is the detuning of the photon energy

k
away from the transition energy, = h(
k
). (There is a constant
oset between the two lines in Eq. C.2.)
Now we can consider the eect of this Hamiltonian as a perturbation
on an initial, coherent photon state linearly polarized along x
[
0
) = e
a

xk
0
[0), (C.3)
where [0) is the vacuum, and is a complex number related to the am-
plitude and phase of the light. The perturbation to this state from the
203
Hamiltonian above can be calculated using time-dependent perturbation
theory. However, the result can be qualitatively anticipated by rewriting
the perturbation in terms of linearly polarized photons:
H
S
= 2i
z

kk

xk
a
yk
+a

yk
a
xk
. (C.4)
If this Hamiltonian acts on photons polarized along x, there are some
y photons mixed in, and vice versa. That is to say, the angle of the polar-
ization is rotated.
The perturbed state can be calculated quantitatively, and the Faraday
rotation angle can be calculated from the expectation value for the electric
eld of the light. Assuming the rotation angle is small, the Faraday rotation
resulting from a spin-up (down) electron is given by

F
= E
y
)/E
x
) = C
D
2
hc
_
h

_
, (C.5)
where C is a factor that depends on the volume in which the electron and
photon interact. This reproduces the odd-Lorentzian form of Eq. 2.19 in
the limit of no transition broadening.
204
References
[1] Chang, S. B. R. & Kirschvink, J. L. Magnetofossils, the magnetiza-
tion of sediments, and the evolution of magnetite biomineralization.
Annual Review of Earth and Planetary Sciences 17, 169195 (1989).
[2] Bazylinski, D. A., Frankel, R. B. & Konhauser, K. O. Modes of
biomineralization of magnetite by microbes. Geomicrobiology Journal
24, 465 475 (2007).
[3] Kimm, C. S. Der Philosoph von Teufelstal. Asia Minor 4, 108
(1927).
[4] Shu-hua, L. Origine de la boussole II. Aimant et boussole. Isis 45,
175196 (1954).
[5] Sakurai, J. J. Modern Quantum Mechanics (Addison-Wesley, 1994).
205
[6] Myers, E. B., Ralph, D. C., Katine, J. A., Louie, R. N. & Buhrman,
R. A. Current-induced switching of domains in magnetic multilayer
devices. Science 285, 867870 (1999).
[7] Datta, S. & Das, B. Electronic analog of the electro-optic modulator.
Appl. Phys. Lett. 56, 665667 (1990).
[8] Mails, T. E. The Mystic Warriors of the Plains: The Culture, Arts,
Crafts and Religion of the Plains Indians (Marlowe and Company,
2002).
[9] Yu, P. Y. & Cardona, M. Fundamentals of Semiconductors (Springer,
2001).
[10] Kikkawa, J. M. & Awschalom, D. D. Resonant spin amplication in
n-type GaAs. Phys. Rev. Lett. 80, 4313 (1998).
[11] Paget, D., Lampel, G., Sapoval, B. & Safarov, V. I. Low eld
electron-nuclear spin coupling in gallium arsenide under optical
pumping conditions. Phys. Rev. B 15, 5780 (1977).
[12] Merkulov, I. A., Efros, A. L. & Rosen, M. Electron spin relaxation
by nuclei in semiconductor quantum dots. Phys. Rev. B 65 (2002).
206
[13] Khaetskii, A. V., Loss, D. & Glazman, L. Electron spin decoherence
in quantum dots due to interaction with nuclei. Phys. Rev. Lett. 88
(2002).
[14] Meier, F. & Zakharchenya, B. P. (eds.) Optical Orientation: Modern
Problems in Condensed Matter (North Holland, Amsterdam, 1984).
[15] Boswarva, I. M., Howard, R. E. & Lidiard, A. B. Faraday eect in
semiconductors. Proceedings of the Royal Society of London Series
a-Mathematical and Physical Sciences 269, 125 (1962).
[16] Kato, Y. K. Electrical Manipulation of Electron Spin Coherence in
Nonmagnetic Semiconductors. Ph.D. thesis, University of California
Santa Barbara (2005).
[17] Hines, M. A. & Guyot-Sionnest, P. Synthesis and characterization
of strongly luminescing ZnS-capped CdSe nanocrystals. Journal of
Physical Chemistry 100, 468471 (1996).
[18] Dabbousi, B. O. et al. (CdSe)ZnS core-shell quantum dots: Synthesis
and characterization of a size series of highly luminescent nanocrys-
tallites. Journal of Physical Chemistry B 101, 94639475 (1997).
207
[19] Li, J. J. et al. Large-scale synthesis of nearly monodisperse CdSe/CdS
core/shell nanocrystals using air-stable reagents via successive ion
layer adsorption and reaction. Journal of the American Chemical
Society 125, 1256712575 (2003).
[20] Kortan, A. R. et al. Nucleation and growth of CdSe on ZnS quantum
crystallite seeds, and vice versa, in inverse micelle media. Journal of
the American Chemical Society 112, 13271332 (1990).
[21] Mews, A., Eychm uller, A., Giersig, M., Schooss, D. & Weller, H.
Preparation, characterization, and photophysics of the quantum-dot
quantum-well system CdS/HgS/CdS. Journal of Physical Chemistry
98, 934941 (1994).
[22] Mews, A., Kadavanich, A. V., Banin, U. & Alivisatos, A. P. Struc-
tural and spectroscopic investigations of CdS/HgS/CdS quantum-dot
quantum wells. Phys. Rev. B 53, 1324213245 (1996).
[23] Tian, Y. C., Newton, T., Kotov, N. A., Guldi, D. M. & Fendler, J. H.
Coupled composite CdS-CdSe and core-shell types of (CdS)CdSe and
(CdSe)CdS nanoparticles. Journal of Physical Chemistry 100, 8927
8939 (1996).
208
[24] Little, R. B., El-Sayed, M. A., Bryant, G. W. & Burke, S. Formation
of quantum-dot quantum-well heteronanostructures with large lattice
mismatch: ZnS/CdS/ZnS. Journal of Chemical Physics 114, 1813
1822 (2001).
[25] Battaglia, D., Li, J. J., Wang, Y. J. & Peng, X. G. Colloidal two-
dimensional systems: CdSe quantum shells and wells. Angewandte
Chemie-International Edition 42, 50355039 (2003).
[26] Schooss, D., Mews, A., Eychm uller, A. & Weller, H. Quantum-dot
quantum-well CdS/HgS/CdS - Theory and experiment. Phys. Rev.
B 49, 1707217078 (1994).
[27] Jaskolski, W. & Bryant, G. W. Multiband theory of quantum-dot
quantum wells: Dim excitons, bright excitons, and charge separation
in heteronanostructures. Phys. Rev. B 57, R4237 (1998).
[28] Bryant, G. W. & Jaskolski, W. Tight-binding theory of quantum-dot
quantum wells: Single-particle eects and near-band-edge structure.
Phys. Rev. B 67 (2003).
[29] Pokatilov, E. P., Fonoberov, V. A., Fomin, V. M. & Devreese, J. T.
Development of an eight-band theory for quantum dot heterostruc-
tures. Phys. Rev. B 64 (2001).
209
[30] Braun, M., Burda, C., Mohamed, M. & El-Sayed, M. Femtosecond
time-resolved electron-hole dynamics and radiative transitions in the
double-layer quantum well of the CdS/(HgS)
2
/CdS quantum-dot-
quantum-well nanoparticle. Phys. Rev. B 64, 035317 (2001).
[31] Gupta, J. A., Awschalom, D. D., Efros, A. L. & Rodina, A. V. Spin
dynamics in semiconductor nanocrystals. Phys. Rev. B 66 (2002).
[32] Rodina, A. V., Efros, A. L. & Alekseev, A. Y. Eect of the surface
on the electron quantum size levels and electron g factor in spherical
semiconductor nanocrystals. Phys. Rev. B 67 (2003).
[33] Schrier, J. & Whaley, K. B. Tight-binding g-factor calculations of
CdSe nanostructures. Phys. Rev. B 67 (2003).
[34] Chen, P. C. & Whaley, K. B. Magneto-optical response of CdSe
nanostructures. Phys. Rev. B 70 (2004).
[35] Stern, N. P. et al. Spin dynamics in electrochemically charged CdSe
quantum dots. Phys. Rev. B 72, 1613034 (2005).
[36] Meier, F. & Awschalom, D. D. Faraday rotation spectroscopy of
quantum-dot quantum wells. Phys. Rev. B 71, 2053159 (2005).
210
[37] Wei, S. H., Zhang, S. B. & Zunger, A. First-principles calculation of
band osets, optical bowings, and defects in CdS, CdSe, CdTe, and
their alloys. Journal of Applied Physics 87, 13041311 (2000).
[38] Xia, J. B. Electronic-structures of zero-dimensional quantum wells.
Phys. Rev. B 40, 85008507 (1989).
[39] Efros, A. L. Luminescence polarization of CdSe microcrystals. Phys.
Rev. B 46, 74487458 (1992).
[40] Rossler, U. (ed.) Semiconductors, II-VI and I-VII Compounds: Semi-
magnetic Compounds, vol. 41 (Springer, Heidelberg, 1999).
[41] Hugonnard-Bruy`ere, S., Buss, C., Vouilloz, F., Frey, R. & Flytzanis,
C. Faraday-rotation spectra of semimagnetic semiconductors. Phys.
Rev. B 50, 22002207 (1994).
[42] Linder, N. & Sham, L. J. Theory of the coherent spin dynamics
in magnetic semiconductor quantum wells. Physica E 2, 412416
(1998).
[43] Sham, L. J. Theory of spin coherence in semiconductor heterostruc-
tures. Journal of Magnetism and Magnetic Materials 200, 219230
(1999).
211
[44] Meier, F., Cerletti, V., Gywat, O., Loss, D. & Awschalom, D. D.
Molecular spintronics: Coherent spin transfer in coupled quantum
dots. Phys. Rev. B 69 (2004).
[45] Ekimov, A. I., Onushchenko, A. A., Plyukhin, A. G. & Efros, A. L.
Size quantization of excitons and determination of the parameters of
their energy spectrum in CuCl. Soviet Physics JETP 61, 891 (1985).
[46] Wolf, S. A. et al. Spintronics: A spin-based electronics vision for the
future. Science 294, 14881495 (2001).
[47] Awschalom, D. D., Loss, D. & Samarth, N. (eds.) Semiconductor
Spintronics and Quantum Computation. NanoScience and Technol-
ogy (Springer-Verlag, Berlin, 2002).
[48] Ouyang, M. & Awschalom, D. D. Coherent spin transfer between
molecularly bridged quantum dots. Science 301, 10741078 (2003).
[49] Eychm uller, A., Mews, A. & Weller, H. A quantum-dot quantum-well
- CdS/HgS/CdS. Chemical Physics Letters 208, 5962 (1993).
[50] Lifshitz, E. et al. Optically detected magnetic resonance study of
CdS/HgS/CdS quantum dot quantum wells. Journal of Physical
Chemistry B 103, 68706875 (1999).
212
[51] Battaglia, D., Blackman, B. & Peng, X. Coupled and decoupled dual
quantum systems in one semiconductor nanocrystal. Journal of the
American Chemical Society 127, 1088910897 (2005).
[52] Gupta, J. A. Coherence and Manipulation of Spin States in Semicon-
ductor Nanostructures. Ph.D. thesis, University of California Santa
Barbara (2002).
[53] Gywat, O., Meier, F., Loss, D. & Awschalom, D. D. Dynamics of
coupled qubits interacting with an o-resonant cavity. Phys. Rev. B
73, 125326 (2006).
[54] Berezovsky, J. et al. Spin dynamics and level structure of quantum-
dot quantum wells. Phys. Rev. B 71, 0813094 (2005).
[55] Kiselev, A. A., Ivchenko, E. L. & Rossler, U. Electron g factor in
one- and zero-dimensional semiconductor nanostructures. Phys. Rev.
B 58, 1635316359 (1998).
[56] Klimov, V. I., McBranch, D. W., Leatherdale, C. A. & Bawendi,
M. G. Electron and hole relaxation pathways in semiconductor quan-
tum dots. Phys. Rev. B 60, 1374013749 (1999).
213
[57] Crooker, S. A., Barrick, T., Hollingsworth, J. A. & Klimov, V. I.
Multiple temperature regimes of radiative decay in CdSe nanocrystal
quantum dots: Intrinsic limits to the dark-exciton lifetime. Appl.
Phys. Lett. 82, 27932795 (2003).
[58] Schrier, J. & Wang, L. W. Electronic structure of nanocrystal
quantum-dot quantum wells. Phys. Rev. B 73 (2006).
[59] Li, Y. Q. et al. Cavity enhanced Faraday rotation of semiconductor
quantum dots. Appl. Phys. Lett. 88 (2006).
[60] Takeda, E. et al. Faraday eect enhancement in Co-ferrite layer in-
corporated into one-dimensional photonic crystal working as a Fabry-
Perot resonator. Journal of Applied Physics 87, 67826784 (2000).
[61] Gourdon, C. et al. Magneto-optical imaging with diluted magnetic
semiconductor quantum wells. Appl. Phys. Lett. 82, 230232 (2003).
[62] Salis, G. & Moser, M. Faraday-rotation spectrum of electron spins
in microcavity-embedded GaAs quantum wells. Phys. Rev. B 72,
115325 (2005).
[63] Kavokin, A. V. et al. Resonant Faraday rotation in a semiconductor
microcavity. Phys. Rev. B 56, 10871090 (1997).
214
[64] Seferos, D. S., Banach, D. A., Alcantar, N. A., Israelachvili, J. N.
& Bazan, G. C. alpha,omega-bis(thioacetyl)oligophenylenevinylene
chromophores from thioanisol precursors. Journal of Organic Chem-
istry 69, 11101119 (2004).
[65] Bracker, A. S. et al. Optical pumping of the electronic and nuclear
spin of single charge-tunable quantum dots. Phys. Rev. Lett. 94
(2005).
[66] Ebbens, A. et al. Optical orientation and control of spin memory in
individual InGaAs quantum dots. Phys. Rev. B 72 (2005).
[67] Stievater, T. H. et al. Measurement of relaxation between polar-
ization eigenstates in single quantum dots. Appl. Phys. Lett. 81,
42514253 (2002).
[68] Li, X. Q., Wu, Y. W., Steel, D. G., Gammon, D. & Sham, L. J.
Raman coherence beats from the entangled state involving polarized
excitons in single quantum dots. Phys. Rev. B 70 (2004).
[69] Hogele, A. et al. Spin-selective optical absorption of singly charged
excitons in a quantum dot. Appl. Phys. Lett. 86 (2005).
215
[70] Elzerman, J. M. et al. Single-shot read-out of an individual electron
spin in a quantum dot. Nature 430, 431435 (2004).
[71] Walls, D. F. & Milburn, G. J. Quantum Optics (Springer, 1994).
[72] Grangier, P., Levenson, J. A. & Poizat, J.-P. Quantum non-
demolition measurements in optics. Nature 396, 537542 (1998).
[73] Meunier, T. et al. Nondestructive measurement of electron spins in
a quantum dot. Phys. Rev. B 74, 1953035 (2006).
[74] Geremia, J. M., Stockton, J. K. & Mabuchi, H. Real-time quantum
feedback control of atomic spin-squeezing. Science 304, 270273
(2004).
[75] Kuzmich, A., Mandel, L. & Bigelow, N. P. Generation of spin squeez-
ing via continuous quantum nondemolition measurement. Phys. Rev.
Lett. 85, 15941597 (2000).
[76] Julsgaard, B., Kozhekin, A. & Polzik, E. S. Experimental long-
lived entanglement of two macroscopic objects. Nature 413, 400403
(2001).
[77] Meier, F. & Awschalom, D. D. Spin-photon dynamics of quantum
dots in two-mode cavities. Phys. Rev. B 70 (2004).
216
[78] Leuenberger, M. N., Flatte, M. E. & Awschalom, D. D. Teleporta-
tion of electronic many-qubit states encoded in the electron spin of
quantum dots via single photons. Phys. Rev. Lett. 94 (2005).
[79] Leuenberger, M. N. Fault-tolerant quantum computing with coded
spins using the conditional Faraday rotation in quantum dots. Phys.
Rev. B 73 (2006).
[80] Zorman, B. & Friesner, R. A. Transitions in ZnS and CdSe quantum
dots and wave-function symmetry. Journal of Chemical Physics 118,
59375946 (2003).
[81] Guest, J. R. et al. Measurement of optical absorption by a single
quantum dot exciton. Phys. Rev. B 65, 241310 (2002).
[82] Zrenner, A. et al. Quantum dots formed by interface uctuations in
AlAs/GaAs coupled-quantum-well structures. Phys. Rev. Lett. 72,
33823385 (1994).
[83] Gammon, D., Snow, E. S., Shanabrook, B. V., Katzer, D. S. & Park,
D. Homogeneous linewidths in the optical spectrum of a single gal-
lium arsenide quantum dot. Science 273, 8790 (1996).
217
[84] Warburton, R. J. et al. Optical emission from a charge-tunable quan-
tum ring. Nature 405, 926929 (2000).
[85] Bracker, A. S. et al. Binding energies of positive and negative trions:
From quantum wells to quantum dots. Phys. Rev. B 72 (2005).
[86] Kupriyanov, D. V. & Sokolov, I. M. Optical-detection of magnetic-
resonance by classical and squeezed light. Quantum Optics 4, 5570
(1992).
[87] Gammon, D., Snow, E. S., Shanabrook, B. V., Katzer, D. S. & Park,
D. Fine structure splitting in the optical spectra of single GaAs
quantum dots. Phys. Rev. Lett. 76, 30053008 (1996).
[88] Dzhioev, R. I. et al. Optical orientation of donor-bound excitons in
nanosized InP/InGaP islands. Physics of the Solid State 40, 1587
1593 (1998).
[89] Cortez, S. et al. Optically driven spin memory in n-doped InAs-GaAs
quantum dots. Phys. Rev. Lett. 89 (2002).
[90] Tischler, J. G., Bracker, A. S., Gammon, D. & Park, D. Fine struc-
ture of trions and excitons in single GaAs quantum dots. Phys. Rev.
B 66, 081310 (2002).
218
[91] El Khali, Y., Gil, B., Mathieu, H., Fukunaga, T. & Nakashima, H.
Dependence of the light-hole heavy-hole splitting on layer thickness
and substrate orientation in GaAs-(GaAl)As single quantum wells.
Phys. Rev. B 39, 1353313536 (1989).
[92] Dzhioev, R. I. et al. Optical orientation and the Hanle eect of neu-
tral and negatively charged excitons in GaAs/Al
x
Ga
1x
As quantum
wells. Phys. Rev. B 66 (2002).
[93] Lehmann, J. & Loss, D. Cotunneling current through quantum dots
with phonon-assisted spin-ip processes. Phys. Rev. B 73 (2006).
[94] Kroutvar, M. et al. Optically programmable electron spin memory
using semiconductor quantum dots. Nature 432, 8184 (2004).
[95] Atat ure, M., Dreiser, J., Badolato, A. & Imamoglu, A. Observation
of Faraday rotation from a single conned spin. Nature Phys. 3, 101
(2007).
[96] Loss, D. & DiVincenzo, D. P. Quantum computation with quantum
dots. Phys. Rev. A 57, 120 (1998).
219
[97] Hanson, R., Kouwenhoven, L. P., Petta, J. R., Tarucha, S. & Van-
dersypen, L. M. K. Spins in few-electron quantum dots. Reviews of
Modern Physics 79, 121749 (2007).
[98] Berezovsky, J. et al. Nondestructive optical measurements of a single
electron spin in a quantum dot. Science 314, 19161920 (2006).
[99] Salis, G., Awschalom, D. D., Ohno, Y. & Ohno, H. Origin of en-
hanced dynamic nuclear polarization and all-optical nuclear magnetic
resonance in GaAs quantum wells. Phys. Rev. B 64, 195304 (2001).
[100] Gurudev-Dutt, M. V. et al. Stimulated and spontaneous optical
generation of electron spin coherence in charged GaAs quantum dots.
Phys. Rev. Lett. 94, 2274034 (2005).
[101] Semenov, Y. G. & Kim, K. W. Phonon-mediated electron-spin phase
diusion in a quantum dot. Phys. Rev. Lett. 92, 0266014 (2004).
[102] Golovach, V. N., Khaetskii, A. & Loss, D. Phonon-induced decay
of the electron spin in quantum dots. Phys. Rev. Lett. 93, 0166014
(2004).
220
[103] Khaetskii, A. V. & Nazarov, Y. V. Spin-ip transitions between
Zeeman sublevels in semiconductor quantum dots. Phys. Rev. B 64,
125316 (2001).
[104] Erlingsson, S. I., Nazarov, Y. V. & Falko, V. I. Nucleus-mediated
spin-ip transitions in GaAs quantum dots. Phys. Rev. B 64, 195306
(2001).
[105] Gammon, D. et al. Electron and nuclear spin interactions in the
optical spectra of single GaAs quantum dots. Phys. Rev. Lett. 86,
51765179 (2001).
[106] Imamoglu, A. et al. Quantum information processing using quantum
dot spins and cavity QED. Phys. Rev. Lett. 83, 42044207 (1999).
[107] Combescot, M. & Betbeder-Matibet, O. Theory of spin precession
monitored by laser pulse. Solid State Commun. 132, 129134 (2004).
[108] Chen, P., Piermarocchi, C., Sham, L. J., Gammon, D. & Steel, D. G.
Theory of quantum optical control of a single spin in a quantum dot.
Phys. Rev. B 69, 075320 (2004).
221
[109] Pryor, C. E. & Flatte, M. E. Predicted ultrafast single-qubit opera-
tions in semiconductor quantum dots. Appl. Phys. Lett. 88, 233108
(2006).
[110] Economou, S. E., Sham, L. J., Wu, Y. & Steel, D. G. Proposal for
optical U(1) rotations of electron spin trapped in a quantum dot.
Phys. Rev. B 74, 205415 (2006).
[111] Clark, S. M., Fu, K.-M. C., Ladd, T. D. & Yamamoto, Y. Quantum
computers based on electron spins controlled by ultrafast o-resonant
single optical pulses. Phys. Rev. Lett. 99, 040501 (2007).
[112] Koppens, F. H. L. et al. Driven coherent oscillations of a single
electron spin in a quantum dot. Nature 442, 766771 (2006).
[113] Cohen-Tannoudji, C. & Dupont-Roc, J. Experimental study of Zee-
man light shifts in weak magnetic elds. Phys. Rev. A 5, 968984
(1972).
[114] Cohen-Tannoudji, C. & Reynaud, S. Dressed-atom description of
resonance uorescence and absorption spectra of a multi-level atom
in an intense laser beam. J. Phys. B 10, 345363 (1977).
222
[115] Suter, D., Klepel, H. & Mlynek, J. Time-resolved two-dimensional
spectroscopy of optically driven atomic sublevel coherences. Phys.
Rev. Lett. 67, 20012004 (1991).
[116] Combescot, M. & Combescot, R. Excitonic Stark shift: A coupling
to semivirtual biexcitons. Phys. Rev. Lett. 61, 117120 (1988).
[117] Jore, M., Hulin, D., Migus, A. & Combescot, M. Laser-induced
exciton splitting. Phys. Rev. Lett. 62, 7477 (1989).
[118] Papageorgiou, G. et al. Spectral dependence of the optical Stark
eect in ZnSe-based quantum wells. Phys. Rev. B 69, 085311 (2004).
[119] Gupta, J. A., Knobel, R., Samarth, N. & Awschalom, D. D. Ultrafast
manipulation of electron spin coherence. Science 292, 24582461
(2001).
[120] Unold, T., Mueller, K., Lienau, C., Elaesser, T. & Wieck, A. D.
Optical Stark eect in a quantum dot: Ultrafast control of single
exciton polarizations. Phys. Rev. Lett. 92, 157401 (2004).
[121] Dutt, M. V. G. et al. Ultrafast optical control of electron spin co-
herence in charged GaAs quantum dots. Phys. Rev. B 74, 125306
(2006).
223
[122] Wu, Y. et al. Selective optical control of electron spin coherence in
singly charged GaAs-Al
0.3
Ga
0.7
As quantum dots. Phys. Rev. Lett.
99, 097402 (2007).
[123] Klingshirn, C. Semiconductor Optics (Springer, 2006), 3rd edn.
[124] Mikkelsen, M. H., Berezovsky, J., Stoltz, N. G., Coldren, L. A. &
Awschalom, D. D. Optically detected coherent spin dynamics of a
single electron in a quantum dot. Nature Phys. 3, 770773 (2007).
[125] Lehmen, A. V., Zucker, J. E., Heritage, J. P. & Chemla, D. S. Phonon
sideband of quasi-two-dimensional excitons in GaAs quantum wells.
Phys. Rev. B 35, 6479(R) (1987).
[126] Rosatzin, M., Suter, D. & Mlynek, J. Light-shift-induced spin echoes
in a J= 1/2 atomic ground state. Phys. Rev. A 42, 1839(R)1841
(1990).
[127] Stoltz, N. G. et al. High-quality factor optical microcavities using
oxide apertured micropillars. Appl. Phys. Lett. 87 (2005).
[128] Martin, I. Personal communication (2006).
224

Вам также может понравиться