You are on page 1of 273

UNIVERSITY of CALIFORNIA

Santa Barbara

Quantum Optics with Quantum Dots in Microcavities

A dissertation submitted in partial satisfaction of the


requirements for the degree of
Doctor of Philosophy
in
Physics
by
Matthew T. Rakher

Committee in charge:
Professor Dirk Bouwmeester, Chair
Professor David Awschalom
Professor Andreas Ludwig
December 2008

3342039

2009

3342039

The dissertation of Matthew T. Rakher is approved:

Professor David Awschalom

Professor Andreas Ludwig

Professor Dirk Bouwmeester, Chair

October 2008

Quantum Optics with Quantum Dots in Microcavities

Copyright 2008
by
Matthew T. Rakher

iii

Dedicated to Caitlin, Pearl, and the rest of my family.

iv

Acknowledgements
There are many people without the support of whom this dissertation would
not be possible. Firstly, my advisor Dirk Bouwmeester has been very supportive
of my research at UCSB and he has provided countless insights into our work together. His need to understand absolutely every part of the experiment, although
at times made my life difficult, has been tremendously useful and is something I
hope to make part of my own future research. I also greatly appreciate the freedom he gave me as a graduate student to direct the research. In addition to Dirk,
there are a few other faculty members at UCSB that have contributed greatly to
my work. This includes Professor Pierre Petroff who provided an encyclopedic
knowledge of device and materials physics as well as equipment. Also, Prof. Evelyn Hu and Prof. Larry Coldren were fantastic collaborators and the sources of
many ideas.
Before I met any of them, my research interests were profoundly influenced by
my undergraduate research advisor, Paul Kwiat. He has been enormously helpful
throughout my physics career. I am endlessly indebted to him for allowing my
to join his group as an undergraduate at Illinois and showing my how to be a
successful research scientist.
The beginning of my time at UCSB was spent learning everything I could about

quantum dots from my favorite German postdoc, Stefan Strauf. His experience
and expertise were invaluable to my education. Together, we must have measured
at least a thousand photonic crystal cavities looking for the ever-elusive strong
coupling. Instead, we ended up finding success by exploring other interesting
things to measure. Above all, Stefan taught me to keep my eyes-and-ears open
to what I am not looking for in an experiment. Along with Stefan, much of those
first two years were spent working with Kevin Hennessy of the Hu group and
Antonio Badolato of the Petroff group. I was able to learn a great deal from both
of them. A bit later I had the pleasure of working with Yong-Seok Choi of the
Hu group on the g (2) ( ) laser measurements, and our collaboration was a success
both in the sense of publications and my education. Yong pushed me to work out
the theory and in doing so I was able to learn how much about quantum optics I
didnt yet know.
A person well-worthy of my gratitude is Nick Stoltz. His hours spent toiling in
the MBE lab and cleanroom enabled all of the measurements in the second half
of this dissertation. The month after I returned from Leiden, September of 2007,
was the single-most productive month of all of my time at UCSB. Undoubtedly,
this was mostly due to the fact that Nick needed to defend that month and we
needed some data! Now, he is in business school at UCLA and will undoubtedly
out-earn me by at least an order of magnitude.

vi

Later in my Ph.D., I spent a lot of time working with Susanna Thon and
Hyochul Kim. Like Nick, they spent many nights (and very early mornings)
working in the cleanroom and in the MBE lab. Their ability to learn outside
of the scope of the standard physics graduate student arena has greatly changed
the capabilities of the Bouwmeester Quantum Optics Group and I am very much
indebted to them for their hard work.
Aside from the people I worked with directly, the other members of the QuOpt
Group at UCSB also contributed significantly to my graduate education through
their friendship and their knowledge. This includes the Oxford contingent of Will
Marshall, William Irvine, and Gabriel Durkin. I appreciated every moment we
had in the office together. In addition to them, it has been a pleasure working
with Dustin Kleckner, Juan Hodelin, Hagai Eisenberg, Michiel de Dood, Eduardo
Fonseca, Jan Gudat, Hubert Krenner, Bart-Jan Pors, Evan Jeffrey (for the second
time!), Jenna Hagemeier and Nima Dinyari. Also, meeting the University of
Leiden Quantum Optics group in the Summer of 2007 was a fantastic experience,
and my discussions with Martin Van Exter were both stimulating and fruitful.
I would also like to acknowledge the friendship and support of my group of
friends who did their undergrad in physics with me at Illinois. Our coalition
comprised of Geert Vrijsen, Mark Murphey, Steve Zoerb, Charles (Chip) Alexander Platt III, Jeffrey Condis, and I enjoyed several years of working together on

vii

our physics homework (and other assorted insanity) and most of what I learned
in those days was due in large part to them.
Finally, I would like to thank my family. Back home in Illinois, my mother
Rhonda, father Mark, and brothers Ryan and Ben have always supported me in
my pursuit of a Physics Ph.D. even though in the back of their minds they think
Im a little strange. Most importantly, my wife Caitlin has been an endless source
of support through her love and kindness. Caitlin and our pug Pearl have made
our home the best place to get away from the lab and have some fun.

viii

Curriculum Vit
Matthew T. Rakher
Education
2008
2005
2003

Ph.D. in Physics, University of California, Santa Barbara


M.A. in Physics, University of California, Santa Barbara
B.S. in Physics, Summa Cum Laude, University of Illinois at
Urbana-Champaign, Urbana, Illinois

Publications
Electrically Injected Single Photons from a Quantum Dot in a High Quality Micropillar Cavity, M. T. Rakher, N. G. Stoltz, L. A. Coldren, P. M. Petroff and
D. Bouwmeester, in preparation.
Cavity QED with Charged Quantum Dots, M. T. Rakher, N. G. Stoltz, L. A.
Coldren, P. M. Petroff and D. Bouwmeester, submitted for publication.
Polarization-switchable single photon source using the Stark effect, M. T. Rakher,
N. G. Stoltz, L. A. Coldren, P. M. Petroff and D. Bouwmeester, Applied Physics
Letters 93, 091118 (2008).
High-frequency single-photon source with polarization control, Stefan Strauf,
Nick G. Stoltz, Matthew T. Rakher, Larry A. Coldren, Pierre M. Petroff and
Dirk Bouwmeester, Nature Photonics 1, 704 (2007).
Evolution of the onset of coherence in a family of photonic crystal nanolasers,
Y.-S. Choi, M. T. Rakher, K. Hennessy, S. Strauf, A. Badolato, P. M. Petroff, and
D. Bouwmeester, E.L. Hu, Applied Physics Letters 91, 031108 (2007).
Self-Tuned Quantum Dot Gain in Photonic Crystal Lasers, S. Strauf, K. Hennessy, M. T. Rakher, Y.-S. Choi, A. Badolato, L. C. Andreani, E. L. Hu, P. M.
Petroff, and D. Bouwmeester, Physical Review Letters 96, 127404 (2006).
Frequency control of photonic crystal resonators by monolayer deposition, S.
Strauf, M. T. Rakher, I. Carmeli, K. Hennessy, C. Meier, A. Badolato, M. J. A.

ix

de Dood, P. M. Petroff, E. L. Hu, E. G. Gwinn, and D. Bouwmeester, Applied


Physics Letters 88, 043116 (2006).
Counterfactual quantum computation through quantum interrogation, Onur
Hosten, Matthew T. Rakher, Julio T. Barreiro, Nicholas A. Peters, and Paul G.
Kwiat, Nature (London) 439, 949 (2006).
High quality-factor optical microcavities using oxide apertured micropillars,
N. G. Stoltz, M. Rakher, S. Strauf, A. Badolato, D. D. Lofgreen, P. M. Petroff,
L. A. Coldren, and D. Bouwmeester, Applied Physics Letters 87, 031105 (2005).

Honors and Awards


2003-2006
2005
2003
2003
2003
2001-2003
1999-2003
1999-2003
2002
1999
1999

UCSB Doctoral Scholar Fellow


Lindau Nobel Laureates Meeting U.S. Representative
UCSB Physics Broida Fellowship
NSF Graduate Fellowship Honorable Mention
Ernest M. Lyman Award for Outstanding Graduating Senior
in Physics
Outstanding Teaching Assistant
Chancellors Scholar
James Scholar
Lorella M. Jones Physics Fellowship
Excellence in Physics Fellowship
Crystal Lake Central High School Valedictorian

Abstract

Quantum Optics with Quantum Dots in Microcavities


by
Matthew T. Rakher

This dissertation describes several quantum optics experiments that rely on


the coupling between an atomic-like system and the confined optical modes of a
cavity as described by cavity quantum electrodynamics (QED). The novelty of
these experiments is that they are performed in the solid-state and as such are
extremely interesting for applications of quantum information. These results have
been obtained through a collaborative effort between the research groups of D.
Bouwmeester in Physics, P. M. Petroff in Materials and ECE, L. A. Coldren in
Materials and ECE, and E. L. Hu in ECE. The first set of these experiments
explores this coupling in photonic crystal defect cavities. The first of these experiments shows how very few quantum dots can act as a sufficient gain medium
to generate extremely low-threshold lasing. This surprising result, which arises
due to the non-atomic-like nature of the quantum dots, is verified by a measurement of the photon statistical transition of the cavity mode. This is done for a

xi

series of such devices to elucidate the differences between macroscopic lasers and
nanolasers. Next, a short experiment is discussed which uses the adsorption of
material inside the cryostat to spectrally tune the resonance of a photonic crystal
cavity. This section of the dissertation concludes with an experiment demonstrating an all-optical scheme to precisely determine the spacial location of a single
quantum dot. Then, using this location, a high-quality photonic crystal cavity is
fabricated, and strong coupling between the quantum and cavity is realized. The
second set of experiments employs a novel, electrically-gated, oxide-apertured micropillar cavity to demonstrate a bright source of optically-generated single photons as well as electrically-generated single photons. Furthermore, the intra-cavity
electric field generated by the gating of the structures enabled the demonstration
of cavity QED with charged quantum dots, which has important ramifications
for solid-state quantum information schemes that use the spin of an electron (or
hole) for manipulation. Finally, the intra-cavity field is used to achieve spectral
resonance between a quantum dot and a cavity mode by the Stark effect. This
effect, in combination with a slightly elliptical micropillar, is used to demonstrate
a polarization-switchable single photon source.

xii

Contents
List of Figures

xvii

1 Introduction
1.1 Cavity Quantum Electrodynamics . . . . . . . . . . . . .
1.1.1 The Jaynes-Cummings Hamiltonian . . . . . . . .
1.1.2 Dynamics in a Ideal Cavity . . . . . . . . . . . .
1.1.3 Dynamics in a Real Cavity . . . . . . . . . . . . .
1.1.4 Strong Coupling . . . . . . . . . . . . . . . . . . .
1.1.5 Weak Coupling . . . . . . . . . . . . . . . . . . .
1.1.6 Interaction with a weak probe field: Input-Output
1.2 Self-Assembled Quantum Dots . . . . . . . . . . . . . . .
1.2.1 InAs Quantum Dot Growth . . . . . . . . . . . .
1.2.2 Optical Properties: Theory . . . . . . . . . . . . .
1.2.3 Microphotoluminescence Spectroscopy . . . . . .
1.2.4 Non-classical states of Light . . . . . . . . . . . .
1.2.5 Quantum Dots in Diode Structures . . . . . . . .

1
. . . . .
4
. . . . .
5
. . . . .
8
. . . . . 12
. . . . . 16
. . . . . 17
Formalism 22
. . . . . 28
. . . . . 29
. . . . . 30
. . . . . 40
. . . . . 43
. . . . . 46

2 Self-Tuned Gain in Photonic Crystal Nanolasers


2.1 Introduction . . . . . . . . . . . . . . . . . . . . .
2.2 Photonic Crystal Cavity Design . . . . . . . . . .
2.3 Experimental Lasing Results . . . . . . . . . . . .
2.4 Discussion . . . . . . . . . . . . . . . . . . . . . .
2.5 Self-Tuned Gain . . . . . . . . . . . . . . . . . . .
2.6 Summary and Conclusion . . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

53
53
55
57
62
64
67

3 Validation of Lasing in PC Nanolasers


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69
69

xiii

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

3.2
3.3
3.4
3.5
3.6

Line Defect PC Cavity Modes . . .


Lasing Transition of a Conventional
Lasing Transition of an L3 Laser .
Comparison of different lasers . .
Conclusion and Outlook . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

70
73
73
77
81

4 Frequency control of photonic crystal membrane


4.1 Introduction . . . . . . . . . . . . . . . . . . . . .
4.2 Drift Tuning of a Photonic Crystal Cavity Mode .
4.3 Mode Tuning by monolayer deposition . . . . . .
4.4 Summary and Conclusion . . . . . . . . . . . . .

resonators
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.

.
.
.
.

83
83
84
91
93

5 Optical Positioning of a Single Quantum Dot


5.1 Introduction . . . . . . . . . . . . . . . . . . .
5.2 Experimental Setup . . . . . . . . . . . . . . .
5.3 Sample Design and Experimental Procedure .
5.4 Data Analysis . . . . . . . . . . . . . . . . . .
5.5 Photonic Crystal Cavity Fabrication . . . . .
5.6 Experimental Results . . . . . . . . . . . . . .
5.7 Discussion and Conclusion . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

94
94
96
97
101
101
103
108

.
.
.
.
.

110
110
111
115
117
120

6 Oxide-Apertured Micropillars
6.1 Introduction . . . . . . . . . . . . .
6.2 Micropillar Cavity Design . . . . .
6.3 Model of Micropillar Cavity Modes
6.4 Experimental Results . . . . . . . .
6.5 Discussion and Outlook . . . . . .

. . . .
Laser
. . . .
. . . .
. . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

7 High-Frequency Single-Photon-Source with Polarization Control122


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
7.2 Electrical Contacts . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7.3 Single Photon Results: X0 and X . . . . . . . . . . . . . . . . . 129
7.4 Polarization Control . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.5 Summary and Conclusion . . . . . . . . . . . . . . . . . . . . . . 137
7.6 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.6.1 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . 139
7.6.2 Optical Characterization . . . . . . . . . . . . . . . . . . . 140
7.6.3 Data Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 140
7.6.4 X /X0 Enhancement Factor . . . . . . . . . . . . . . . . . 141
7.7 Rate Equation Model . . . . . . . . . . . . . . . . . . . . . . . . . 142

xiv

8 Electrically-Injected Single-Photon-Source
8.1 Introduction . . . . . . . . . . . . . . . . .
8.2 Micropillar Design for Electrical Injection .
8.3 Experimental Demonstration and Results .
8.4 Comparison to Optical Excitation . . . . .
8.5 Conclusion and Outlook . . . . . . . . . .
9 Cavity QED with Charged Quantum Dots
9.1 Introduction . . . . . . . . . . . . . . . . .
9.2 Sample Design . . . . . . . . . . . . . . . .
9.3 Intra-cavity Charging . . . . . . . . . . . .
9.4 Coherent Cavity Reflection Spectroscopy .
9.5 Intra-cavity Stark Tuning . . . . . . . . .
9.6 Summary and Conclusion . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

10 Polarization-switchable Single Photon Source


fect
10.1 Introduction . . . . . . . . . . . . . . . . . . .
10.2 Sample Design and Fabrication . . . . . . . .
10.3 Polarization Switching . . . . . . . . . . . . .
10.4 Theoretical Model and Discussion . . . . . . .
10.5 Conclusion and Outlook . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

147
147
148
151
154
156

.
.
.
.
.
.

158
158
160
163
166
169
171

using the Stark ef173


. . . . . . . . . . . 173
. . . . . . . . . . . 174
. . . . . . . . . . . 177
. . . . . . . . . . . 181
. . . . . . . . . . . 184

11 Conclusion

186

A Quantum Dot Fine Structure

189

B g (2) ( ): A Semiclassical-Langevin Approach

194

C g (2) (0): A Quantum Density Matrix Approach

204

D Reconvolution

218

E Using the Cryostat


E.1 Sample Unloading . . . . . . . . . . . . . . . . . . . . . . . . . . .
E.2 Sample Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E.3 Cool-Down Procedure . . . . . . . . . . . . . . . . . . . . . . . .

227
227
229
233

xv

F List of Samples
237
F.1 Photonic Crystal Samples . . . . . . . . . . . . . . . . . . . . . . 238
F.2 Micropillar Samples . . . . . . . . . . . . . . . . . . . . . . . . . . 239

xvi

List of Figures
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
1.9
1.10
1.11
1.12
1.13
1.14

Quantum Level Diagram in Cavity QED . . . . . . .


Rabi Oscillations in an Ideal Cavity . . . . . . . . . .
Spectral Anti-crossing under Loss . . . . . . . . . . .
The Purcell Effect . . . . . . . . . . . . . . . . . . . .
Reflection Spectra with and without a Quantum Dot
Reflection Spectra for Different Detunings . . . . . .
AFM image of Quantum Dots . . . . . . . . . . . . .
Quantum Dot Shell Structure . . . . . . . . . . . . .
Biexciton Cascade and Trion Levels . . . . . . . . . .
PL Experimental Setup . . . . . . . . . . . . . .
Single QD PL Spectrum . . . . . . . . . . . . . .
g (2) ( ) for a Single QD . . . . . . . . . . . . . . . . .
Band Structure of a QD embedded in a pin diode . .
Spectra of a QD in a pin Structure . . . . . . . . . .

2.1
2.2
2.3
2.4

SEM, simulation, and spectra of an L3 photonic crystal


Lasing transition of an L3 cavity . . . . . . . . . . . .
Cavity mode linewidth and intensity . . . . . . . . . .
Self-tuned gain in PC lasers . . . . . . . . . . . . . . .

3.1
3.2
3.3
3.4

Line defect photonic crystal cavity modes . . . . . . .


Lasing transition of a conventional laser . . . . . . .
Lasing transition of an L3 laser . . . . . . . . . . . .
Photon statistical comparison of PC and conventional
sitions . . . . . . . . . . . . . . . . . . . . . . . . . .

4.1
4.2

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

10
12
18
21
25
27
31
33
37
42
44
47
48
50

cavity
. . . .
. . . .
. . . .

.
.
.
.

.
.
.
.

56
58
60
65

. . . . . . .
. . . . . . .
. . . . . . .
lasing tran. . . . . . .

72
74
76

Spectral redshift of a cavity mode due to molecular adsorption . .


Drift tuning and cycling . . . . . . . . . . . . . . . . . . . . . . .

86
87

xvii

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

79

4.3
4.4

Temperature dependence of cavity modes and quantum dots . . .


Comparison of cavity mode redshift for different cavities . . . . .

5.1
5.2
5.3
5.4
5.5
5.6

Experimental Setup for Optical Positioning . . . . . . . . . . .


SEM image of the Optical Positioning Sample . . . . . . . . .
Optical Positioning Data . . . . . . . . . . . . . . . . . . . . .
SEM Image of a Positioned Photonic Crystal . . . . . . . . . .
Strong Coupling between a PC cavity mode and a QD . . . .
Spectral Anti-Crossing for a QD coupled to a PC cavity mode

6.1
6.2
6.3

Micropillar device schematic, SEM, and mode simulation . . . . . 114


Micropillar mode spectra . . . . . . . . . . . . . . . . . . . . . . . 117
Fabry-Perot measurement of micropillar mode . . . . . . . . . . . 119

7.1
7.2
7.3
7.4
7.5
7.6
7.7

Single Photon Source Device Design . . . . .


Effect of gate voltage on SPS performance .
Single QD Characterizations . . . . . . . . .
Impact of QD charging on SPS performance
Demonstration of polarization control . . . .
Level diagrams for X0 and X . . . . . . . .
Results of Rate Equation Model . . . . . . .

8.1
8.2
8.3

Device schematic for electrically-injected single photons . . . . . . 150


Electroluminescence . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Electrically injected single photons . . . . . . . . . . . . . . . . . 155

9.1
9.2
9.3
9.4

Schematic and image of charging structure


Intra-cavity QD charging . . . . . . . . . .
Resonant reflection spectroscopy . . . . . .
Stark effect tuning . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

161
165
168
170

10.1
10.2
10.3
10.4

Linearly polarized, non-degenerate cavity modes .


Stark shift tuning for polarization switching . . .
Fast polarization switching . . . . . . . . . . . . .
Average polarization visibility vs. drive frequency

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

176
179
180
183

.
.
.
.

.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.

89
90
98
99
102
104
106
107

126
128
131
133
136
143
145

A.1 X 0 Fine Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . 192


B.1 Calculated L-L curves for various lasers . . . . . . . . . . . . . 196
B.2 Relaxation Oscillations in High PC Lasers . . . . . . . . . . . . 200
B.3 g (2) ( ) for High PC Lasers . . . . . . . . . . . . . . . . . . . . . 201

xviii

B.4 Theoretical Photon Statistics Transition of a High Laser . . . . 202


C.1
C.2
C.3
C.4

Level Scheme for Density Matrix Formulation


Photon number distribution (n) . . . . . . .
hni as a function of rp . . . . . . . . . . . . .
g (2) (0) as a function of rp . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

206
213
215
217

D.1 TCPC Raw Data and Instrument Response . . . . . . . . . . . . 220


D.2 2 of TCPC data . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
D.3 Fit using Reconvolution . . . . . . . . . . . . . . . . . . . . . . . 226
E.1 Sample Mounting . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
E.2 Properly Mounted Sample . . . . . . . . . . . . . . . . . . . . . . 232
F.1 Micropillar Cavity Layout . . . . . . . . . . . . . . . . . . . . . . 238

xix

Chapter 1
Introduction
The primary significance of the experiments in this dissertation is that they
combine two seemingly disparate fields, the first of which, cavity quantum electrodynamics, is ground in atomic physics and quantum optics while that latter,
semiconductor quantum dots, is based off of condensed matter physics and materials science. When combined they enable similar experiments to what has been
done in atomic physics albeit with new obstacles and challenges as well as new
physics. The potential impact of encountering, overcoming, and understanding
these challenges lies in the potential for future technologies and devices, specifically
in the area of quantum information. While experiments on atoms and trapped
ions are of great fundamental interest, the inherent difficulties in using them make
their potential technological impact somewhat cloudy. Most experiments that use

1 Introduction
them act more as elegant proof-of-principle demonstrations of quantum information protocols rather than cornerstones on which to build scalable devices [15].
That is not to say that these experiments are not without merit, but rather that
quantum information as a field hasnt advanced much beyond initial demonstrations using small system sizes. On the other hand, condensed matter physics
and materials science have contributed immeasurably to information technology
in its present state. State-of-the-art nanofabrication techniques as well as very
high quality material growth have been the driving forces behind most major advances in information technology. That being said, the physics of these systems
at the nanoscale is not nearly as clean and elegant as in atomic physics systems,
and therein lies the challenge. It is this challenge that the work described herein
takes steps to address using the precision measurement techniques and theoretical
framework of atomic physics. In this way, the atomic-like nature of a solid-state
system, the semiconductor quantum dot, can be elucidated; while at the same
time, the limitations of treating it as atomic-like can also be determined. Both
of these qualities will play a large part in understanding the results of the experiments in this dissertation.
It is of utmost importance to recognize that the work presented in this dissertation does not result from the effort of an individual, but results from the combined
effort of an interdisciplinary research collaboration. The tasks involved in these

1 Introduction
kind of experiments include sample design, sample growth, device fabrication, optical characterization, simulation, and data analysis. Within this collaboration,
my primary contributions have been:
1. Giving guidance on the required sample parameters and sample design.
2. Optical testing and optimization of samples.
3. Creating theoretical models to guide and understand the experiments.
4. Originating and performing quantum optics experiments on the samples.
5. Analyzing the data and making conclusions based on them.
Aside from my efforts, these experiments greatly benefitted from the device growth
and fabrication by fellow Ph.D. students Nick Stoltz, Hyochul Kim, and Susanna
Thon.
This dissertation is organized as follows. The introduction is split into two
parts that are each necessary to understand the experiments described in this dissertation. The first part introduces cavity quantum electrodynamics in a general
context and the second describes semiconductor quantum dots. The first half of
the experimental content of this work describes measurements on quantum dots
interacting with the cavity mode of a photonic crystal defect cavity. The latter
half describes measurements on electrically-gated quantum dots interacting with

1 Introduction
the cavity modes of a micropillar cavity. There are a total of six appendices, the
first three of which give detailed theoretical analysis to support the introductory
material and the results of experiments, and the last three are written primarily
as a guide to future students in the Bouwmeester Group.

1.1

Cavity Quantum Electrodynamics

The underlying physics detailed in the experiments found in this dissertation is embodied in a fundamental quantum optics concept, that of cavity quantum electrodynamics (QED). At its simplest, cavity QED is the interaction of an
atomic-like quantum system with a confined electromagnetic mode of a cavity.
Studies of this interaction have led to many breakthroughs in our understanding
of fundamental physics, most notably in the areas of quantum mechanics and the
quantum nature of electromagnetic fields. These types of experiments include the
first demonstration of field quantization in a cavity [6]. For a complete review
of early atomic experiments, see Ref. [7]. The cavity QED paradigm is also an
interesting test bed for properties of single atoms or single atomic-like systems
because cavity confinement effectively increases the coupling between the quantum system and the electromagnetic field. This is especially useful in the optical
regime where there are numerous instruments and techniques available for the

1 Introduction
manipulation and detection of light fields. An external field such as a laser can
be efficiently coupled into a microcavity, and since that microcavity is coupled to
the single quantum system, cavity QED acts a coherent measurement pathway
from the quantum world to the macroscopic world. Furthermore, cavity QED
systems are ideal experimental arenas for quantum information, communication,
and computation schemes [8]. Specifically, schemes that use coupling between
photons and atomic-like systems such as quantum repeaters strongly benefit from
the kind of interaction strength found in cavity QED [9]. Additionally, many
proposals within measurement-based quantum computation [10] make use of
the coupling in cavity QED to efficiently capture single photons from two-level
emitters [11]. Then, joint measurements on the emitted photons can be used to
perform remote operations on the emitters inside the cavities. Thus, cavity QED
is a very interesting field of study within physics which has ramifications for fundamental physics as well as the potential for important contributions to quantum
information science.

1.1.1

The Jaynes-Cummings Hamiltonian

Before describing the solid-state quantum optics experiments found in this


dissertation, it is imperative to introduce the physics behind cavity QED. This is
important not only to understand the results of experiments, but also to determine

1 Introduction
what kind of experiments are possible and how they can be performed. This discussion begins with the fundamental object of cavity QED, the Jaynes-Cummings
Hamiltonian for a two-level system interacting with a single electromagnetic field
mode, which can be expressed as

1
H = ~a z + ~c a a + ~g + a + a ,
2

(1.1)

where ~a = Ea Eb is the difference in energy between upper level | ai and


lower level | bi, c is the cavity resonance frequency and g is the atom-cavity
coupling parameter [12]. The constant term due to the vacuum energy of the field
has been neglected. a (a ) are the cavity field lowering (raising) operators with
with bosonic commutation relation [a, a ] = 1. z , + , and are spin-1/2 Pauli
operators which when expressed in the atomic Hilbert space can be represented
as 22 matrices

1 0
,
z =

0 1

0 1
,
+ =

0 0

0 0
,
=

1 0

(1.2)

[+ , ] = z .

(1.3)

with the following commutation relations


[+ , z ] = 2+ ,

[ , z ] = 2 ,

The action of the interaction term is clearly understood to do the following: the
first term removes one quanta of energy from the field and promotes the atom to

1 Introduction
| ai and the second term adds one quanta of energy to the field and reduces the
atom to | bi.
The parameter g that governs the strength of the interaction depends intimately on the properties of the atomic transition and the volume in which the
cavity mode is confined. This dependence be can understood because g arises
~ and is
from the coupling of the transition dipole moment
~ to the electric field E
defined as [13]

~
~g = h~
Ei .

(1.4)

The strength of the electric field can be related to the spatial extent of the cavity
mode by defining the cavity mode volume Vef f as
Vef f =

1
~ r)|2 ]
M ax[n2 (~r)|E(~

~ r)|2 ,
d3~r n2 (~r)|E(~

(1.5)

where n(~r) is the spatially-dependent index of refraction and M ax[ ] denotes the
maximum value. For clarity, the spatial dependence of the index will be removed
~ r)|2 ] = n2 |Emax |2 ). If the cavity mode is a vacuum mode, then it
(M ax[n2 (~r)|E(~
ef f
must contain total energy 21 ~c . Then by using the definition of the energy stored
in a classical electromagnetic field, |Emax | can be expressed as
s
~c
.
|Emax | =
2
2nef f o Vef f

(1.6)

In the literature, it is common to find the strength of an atomic transition


written in terms of the oscillator strength f and not the bare transition dipole
7

1 Introduction
moment . These two parameters are related by
f=

2mo a 2
~e2

(1.7)

where mo is the free electron mass and e is the charge of an electron [13]. Combining Eq. 1.6 and 1.7 into the definition of g with a c yields the most useful
form for calculations

s
g=

e2 f
.
4mo n2ef f o Vef f

(1.8)

From Eq. 1.8, it is easily understood why cavity QED experiments benefit from
very small cavities and why the development of such microcavities has been very
intensively pursued.

1.1.2

Dynamics in a Ideal Cavity

The dynamics that arise from such a coupling can be easily calculated from

the Schrodinger equation ~ t


| i = H | i. It is convenient to analyze the system

in the basis of states labeled | i, ni where i = {a, b} represents the state of the
atom and n = {0, 1, 2, . . .} represents the number of photons in the cavity, so that
the state | i can be expressed as a column vector
n
| i =

| b, 0i , | a, 0i , | b, 1i , . . . , | a, ni , | b, n + 1i , . . .

oT

(1.9)

1 Introduction
and is shown pictorially on the left side of Fig. 1.1. In this basis H can be written
as

1
2 a

H = ~

2 a

21 a + c

..
.

..

...

...

.
1

2 a

g n+1

+ nc

g n+1

21 a + (n + 1)c

..
.

. . .

0 . . .

0 . . .

. . .
.

0 . . .

0 . . .

..
. ...
(1.10)

It is clear from the form of Eq. 1.10 that the system can be fully diagonalized by
considering only the 22 matrix

Hn+1

1
g n+1
2 a + nc

.
= ~

g n + 1 21 a + (n + 1)c

(1.11)

Diagonalization of Eq. 1.11 results in the following eigenstates


| in+1,

n
1
g n + 1 | a, ni +
=p
Nn+1,

!
o
2
+ g 2 (n + 1) | b, n + 1i ,
4
(1.12)

where = c a and normalization


"

2
Nn+1,

2
+ g 2 (n + 1)
=2
4

2
+ g 2 (n + 1)
4

#
(1.13)

1 Introduction

b,n+1]

y]n+1,+

a,n]

2Wn+1
y]n+1,-

b,2]

.
.
.

.
.
.

b,1]
wc

b,0]

a,1]

y]2,+
y]2,-

a,0]

y]1,+
y]1,-

wa

Figure 1.1: Quantum level diagram for the cavity QED system. The levels shown
on the left are the eigenstates for an uncoupled system while those on the right
are the dressed states which result from coupling. The separation between each
state in each subset is 2~n+1 .
with energies

En+1, =

1
n+
2

~c ~

1
g 2 (n + 1) + 2 .
4

(1.14)

The formation of these new eigenstates of the system is shown schematically


in Fig. 1.1. As in the quantum mechanics of atomic levels in an external magnetic
field, coupling between states leads to the formation of new eigenstates. While

10

1 Introduction
the result of Eq. 1.12 is quite general, it is instructive to consider the resonant
case with = 0, yielding eigenstates

1
| in+1, = | a, ni | b, n + 1i
2

(1.15)

with energies En+1, = (n + 21 )~c ~g n + 1. These states, called dressed


states, are depicted on the right in Fig. 1.1. These eigenstates of the system
are notable because quantum mechanically, it is no longer useful to consider the
two subsystems separately, but rather they must be considered as one quantum
system with the energy shared between the atom and the cavity. Furthermore, if
the system begins its evolution with the atom in the excited state and n photons
in the cavity, then the time-dependent wavefunction of the system will be
n
o

1
1
| (t)i = e(n+ 2 )c eg n+1t | i1,+ + eg n+1t | i1, .
2

(1.16)

The probability that the atom is in state a at time t is given by


Pa (t) = | ha, n|(t)i |2 = cos2 (n+1 t),

(1.17)

where n+1 = g n + 1 is called the Rabi frequency. Thus, the interaction with
the cavity causes the energy, which was initially stored entirely in the atom, to
oscillate from the atom to the cavity and back again at a frequency n+1 as
shown in Fig. 1.2 for n = {0, 1, 4}. If there are no losses, the energy will oscillate
undamped, and for a given initial energy (in terms of the atomic state and the
11

1 Introduction

1.0
0.8
PaHtL

0.6
0.4
0.2
0.0
0.0

0.5

1.0

1.5 2.0
gt

2.5

3.0

Figure 1.2: A plot of Eq. 1.17 for n = 0, 1,


4 in red, green, and blue respectively.
An increase in n effectively increases g by n + 1.
photon number), the wavefunction will continue oscillating between the two states
in that energy subspace.

1.1.3

Dynamics in a Real Cavity

While the physics of the last section is interesting, it represents an ideal case. In
reality, and specifically in the laboratory, the situation is not so simple because the
cavity is coupled to its external electromagnetic environment. This coupling gives
rise to a finite time that a photon can oscillate in the cavity and correspondingly
a finite spectral width. The strength of this coupling is often parameterized by

12

1 Introduction
the quality factor Q of the cavity. This quantity can be expressed in terms of the
electric field amplitude (not energy!) loss rate of the cavity as Q = c /2. The
impact of on cavity QED is to effectively degrade g. This can be understood
heuristically by noting that if g is small compared to , then once the energy is
transferred from the atom to cavity as in Fig. 1.2, it will be lost before it can
re-excite the atom.
Calculating the exact dynamics of the lossy cavity QED system becomes much
harder because will couple the mth photon branch of states to all states with
n < m. However, the dynamics can be calculated analytically in the limit of
very small photon number, which truncates the Hilbert space. In the Heisenbergy
picture, the equation of motion for an operator O is

O = [H, O] + L(O),
~

(1.18)

where H is defined in Eq. 1.1 and L is a loss operator to be defined later. The
operators needed to adequately describe the system are a, a , + , and . The loss
operators can be defined following the seminal work of Gardiner and Collett [14]
where they are taken to be Markovian loss operators arising from an interaction
with a reservoir of harmonic oscillators. Thus, application of Eq. 1.18 to the

13

1 Introduction
operators of interest yields
a = c a g a

(1.19)

= a + gaz + z

(1.20)

a = c a + g+ a

(1.21)

+ = a + ga z + z +

(1.22)

where is the dipole decay rate which gives rise to a homogeneous atomic linewidth
of 2. For the systems considered here, is much smaller than both g and and
can often be neglected. Eq. 1.19-1.22 cannot be solved easily because of the
nonlinear terms az and a z . To linearize the equations, the dynamics will be
calculated in the limit of very small photon number, which means that the atom
is almost always in its ground state (hz i = 1). Thus, z will be approximated
as -1. Furthermore, only Eq. 1.19 and 1.20 need to be solved because Eq. 1.21
and 1.22 are just their Hermitian conjugates. This simplification results in the set
of linear equations
a = c a g a

(1.23)

= a ga .

(1.24)

This set of coupled, linear differential equations can be solved using linear algebra

14

1 Introduction
by rewriting Eq. 1.23-1.24 in matrix form as
v = M v,
where

v=
,

(1.25)

g
c
.
M=

g
a

(1.26)

Of course, for an eigenvector v oscillating at with decay rate , the action of the
time derivative would bring down a factor of , so the resonant frequencies
and their decay rates can be obtained by finding the eigenvalues of M. This
procedure results in the following resonant frequencies and decay rates

1
= (a + c ) ( + )
2
2

r
g2 +

1
[(a c ) + ( )]2 .
4

(1.27)

This represents the main result of this section. Note that in the absence of losses
, 0, Eq. 1.27 is exactly the same as Eq. 1.14, the energies of the eigenstates
found in the lossless case. Thus, Eq. 1.27 represents the quasi-energies of the open
system. Secondly, in the limit of no coupling, Eq. 1.27 reduces to the pure state
of the atom and the cavity respectively, as it should.
There are two regimes of interesting dynamics that result from Eq. 1.27 depending on whether the term under the square root is positive or negative. At
resonance, this term is given as g 2 ( )2 /4, and just as suspected, the dynamics depend on the ratio of the coupling g to the losses , . If this term is positive,
15

1 Introduction
then it adds an additional oscillating term to the dynamics and Rabi oscillations
can occur just like in the lossless case. If it is instead negative, then the whole
term will be imaginary and the decay rates will be modified. These regimes are
called the strong coupling and weak coupling regimes respectively.

1.1.4

Strong Coupling

First, the strong coupling regime will be explored in more detail. Eq. 1.27 is
plotted as a function of detuning = a c in Fig. 1.3 for parameters typical
of a quantum dot interacting with a photonic crystal defect cavity in the strong
coupling regime, like found in Chap. 5. The thick, solid lines denote while the
surrounding dashed lines denote the linewidths of those resonances given by .
The solid lines show a clear anti-crossing, an unmistakable characteristic of the
formation of new eigenstates. In the case of a real cavity, the energy splitting at
the anti-crossing point is
r
E = 2~1 = 2~

g2

1
( )2 ,
4

(1.28)

and compared to the ideal case of Eq. 1.17, the losses effectively reduce the coupling g, as suspected. In addition to the spectral anti-crossing, the linewidths of
the resonances undergo an interesting transition. The solid, red line near the top
of the graph begins at large negative detunings as a broader transition, which is

16

1 Introduction
characteristic of the cavity mode, before narrowing near the anti-crossing point
and then narrowing further until approaching the linewidth of the quantum dot
emission. On the other hand, the solid blue line begins as quantum dot like,
with a very narrow linewidth, and ends up as cavity-like, with a broad linewidth.
The anti-crossing of the resonances as well as the exchange of linewidths are clear
spectral indicators of strong coupling. Experimentally, these resonances (and to a
certain extent the linewidths) can be measured simply by recording the spectrum
of the photoluminescence from the QD and the cavity mode as a function of detuning. In quantum dot cavity QED experiments, this is the primary technique
for measuring strong coupling and this was first observed in 2004 by Ref. [15] and
[16] by spectrally tuning the quantum dot via temperature into resonance with a
microcavity mode.

1.1.5

Weak Coupling

The weak coupling regime is identified by the modification of the atomic spontaneous emission rate. If g , then the term under the square root in
Eq. 1.27 at resonance can be expanded and to linear order in g/ (neglecting )

17

1 Introduction

1.3304

1.3302
1.3300
1.3298
1.3296
-0.0004

-0.0002
0.0000
0.0002
D = Ha-cL

0.0004

Figure 1.3: A plot of Eq. 1.27 as a function of = a c for the following


parameters: {g, , , c } = {100 106 , 50 106 , 10 106 , 1.33}, all in units
of eV. These parameters are typical cavity QED parameters of a quantum dot
interacting with a photonic crystal defect cavity with Q 13000 in the strong
coupling regime. The thick, red and blue solid lines denote respectively while
the surrounding dashed lines denote the linewidths of those resonances given by
.

18

1 Introduction
results in
+ + = a

g2

= c +

(1.29)
g2
.

(1.30)

As before + is the atomic-like resonance and from Eq. 1.29, the cavity modified
decay rate is now
cav = 2

g2
.

(1.31)

In 1946, Edward Purcell predicted an enhanced emission rate from an atom coupled to a cavity using Fermis Golden Rule [17]. In his work, he compared the
rate from an atom in a cavity to the free space decay rate and this ratio is called
the Purcell Factor Fp . This ratio can be derived from Eq. 1.31, the free space
spontaneous decay rate
o =

a3 n2
,
3o ~c3

(1.32)

the definition of f in Eq. 1.7, and = c /2Q to be


3 Q
cav
= 2
Fp =
o
4 Vef f

.
n

(1.33)

Equation 1.33 has provided the mathematical impetus for the development of
cavities with a high Q/Vef f , although it is important to note that in the strong
p
coupling regime, Q/ Vef f is the relevant ratio.

19

1 Introduction
The enhancement of spontaneous emission is often measured as a function of
detuning from the cavity resonance by a time-resolved photon counting experiment. That is, one often measures the atomic emission decay rate on resonance
with the cavity and then compares it to either the free-space decay rate, or the
off-resonance rate. Fig. 1.3 plots the ratio of the cavity-enhanced decay rate to
the free-space decay rate as a function of the detuning for various values of g.
As expected, an increase in g leads to an enhancement of the emission rate on
resonance with the cavity. The other values used in the curve are typical values
for a quantum dot in a micropillar cavity with Q 40000 like found in Chap. 6.
Surprisingly, for these values very close to the strong-coupling regime, the Purcell
enhancement is only about a factor of 5, even though a simple calculation using
Eq. 1.33 yields a value more like 70. This is due to the fact that 6= 0 and that
the value of g is large enough that 1st order perturbation theory is no longer valid.
By measuring the emission rate of quantum emitters in cavities, weak coupling
was first demonstrated in semiconductor microcavities in 2000-2001, see Ref. [18
21].

20

1 Introduction

4.5

g=2.5 eV
g=3.3 eV
g=4.1 eV
g=5. eV
g=5.8 eV
g=6.6 eV
g=7.5 eV

4.0
GcavGo

3.5
3.0
2.5
2.0
1.5
1.0
107

108

109
D HHzL

1010

1011

Figure 1.4: A plot of cav /o = 2+ /o from Eq. 1.27 as a function of = a c


in Hz for the following parameters: {, , c } = {16 106 , 2 106 , 1.33},
all in units of eV. These parameters are typical for a quantum dot interacting with a micropillar cavity with Q 40000. These curves are made for
g = {2.5, 3.3, 4.1, 5, 5.8, 6.6, 7.5} eV and the spontaneous lifetime enhancement
increases with g.

21

1 Introduction

1.1.6

Interaction with a weak probe field: Input-Output


Formalism

A final method of detecting coupling between the emitter and a cavity mode
is accomplished by weakly probing the system with a laser. The Input-Output
formalism developed by Collett and Gardiner [14] provides the best way to properly include the interaction with external measurement modes and/or reservoirs.
This is done by adding input fields a1,in , a2,in , and bin to the equations of motion
Eq. 1.23 and 1.24,
a = c a g a

21 a1,in 22 a2,in ,

= a ga

p
2bin ,

(1.34)
(1.35)

where 1 and 2 are the field loss rates of the first and second mirror respectively
and once again is the total cavity field loss rate. This will include both mirror
losses as well as any internal loss mechanisms such as scattering or absorption.
There are two input electromagnetic modes because the cavity is double-sided.
One mode would be sufficient for a single-sided cavity. bin represents the input
annihilation operator of the atomic reservoir and can be used to simulate the effect
of temperature. The output field a1,out is related to the input field and the cavity
field by
a1,out = a1,in +
22

21 a.

(1.36)

1 Introduction
This system of linear equations can by solved by Fourier Transform, here
defined as
1
a(t) =
2

Z
det a
().

(1.37)

The Fourier transforms of Eq. 1.34 and 1.35 can be expressed in matrix form as
cin = H
v,
where

and

1,in + 22 a
2,in
21 a
,
cin =


2 bin

(1.38)

v =
,

(1.39)

g
(c )

.
H=

g
(a )

(1.40)

Thus, v = H1 cin and a


and
can be written in term of the input fields as
p

1,in + 22 a
2,in ) + g 2bin
a
= a ( 21 a

= c

2bin + g( 21 a
1,in + 22 a
2,in ),

(1.41)
(1.42)

where
a = ( a ),

(1.43)

c = ( c ),

(1.44)

1 = c a + g 2 .

23

(1.45)

1 Introduction
Suppose the system is probed by a weak laser field incident on mirror 1 so that
h
a1,in a
1,in i =

|p |2
,
1

(1.46)

where p is the electric field amplitude of the laser field which is proportional to
the square root of the intensity. If there is no field incident on mirror 2 and both
baths are taken to be at zero temperature, then
h
a2,in a
2,in i = hbinbin i = 0.

(1.47)

The normalized reflection spectrum in output 1 is then given by


R=

h
a1,out a
1,out i
h
a1,in a
1,in i

= |1 21 a |2 .

(1.48)

Similarly, the normalized transmitted spectrum is


T =

2,out i
h
a2,out a
h
a1,in a
1,in i

= 41 2 |a |2 .

(1.49)

For the experiments in this dissertation, only R is easily experimentally accessible.


Fig. 1.5 is a plot of the reflection spectrum, Eq. 1.48, using typical parameters
for a micropillar cavity. The maroon curve shows the spectral response if there is
no quantum dot in the cavity. It is a simple dip with width equal to 2. On the
other hand, the blue curve is the spectrum for a quantum dot interacting with the
cavity with g = 10 eV, a reasonable value for a system like found in Chap. 9.
The response is drastically modified and now has two dips with a narrow peak in
24

1 Introduction

1.0
0.8

0.6
0.4
0.2
0.0

1.29995

1.30000
Energy HeVL

1.30005

Figure 1.5: A plot of Eq. 1.48, the reflection spectrum, for cavity QED parameters
typical of a micropillar cavity like found in Chap. 9 ( = 15 eV) with 1 = /2.
The maroon curve shows the spectral response if there is no quantum dot in the
cavity (g = 0) while the blue curve shows the response with a quantum dot (g = 10
eV, = 2 eV).

25

1 Introduction
the middle. The height of this peak and the separation of the two dips depend
strongly on g. This plotted system is in the strong coupling regime, and this
can be determined from the spectrum by comparing the separation between the
two dips to the width of the dip in the spectrum without the QD. Fig. 1.6 shows
the reflection spectrum for various quantum dot-cavity detunings using the same
parameters as in Fig. 1.5. As the QD is spectrally shifted closer to the cavity
resonance, the dip due to the cavity is pushed towards higher energies, just like
what was found in the anti-crossing spectrum of Sec. 1.1.4. Similarly, for positive
detunings the cavity resonance is pushed towards lower energies. It is important
to note that even for weak coupling, the reflection spectrum is still composed of
two dips with a peak in the middle, but the spectral difference between the dips
is smaller than cavity width.
The experimental benefit of performing this kind of measurement over a simpler anti-crossing measurement like that of Sec. 1.1.4 is in the enhanced spectral
resolution. The resolution of a spectral anti-crossing measurement using photoluminescence is limited by the resolution of the spectrometer. For a 1.25 m
spectrometer with a grating like used in the measurements in this dissertation,
the spectral resolution is 30 eV. This value is on the same scale as expected
vacuum Rabi-splittings and thus doesnt provide a very precise measurement. Additionally, the homogeneous linewidth 2 is typically 4 eV and is much smaller

26

1 Introduction

Figure 1.6: A plot of Eq. 1.48 for various quantum dot-cavity detunings. Shown
are plots for detunings of a c = -37.5, -22.5, -7.5, 7.5, 22.5, and 37.5 eV. All
other cavity QED parameters and plot axes are the same as in Fig. 1.5.

27

1 Introduction
than the minimum resolution. On the other hand, reflection measurements are
limited by the linewidth of the probe laser, which for the laser used in Chap. 9 is
a few MHz ( 0.02 eV). Thus, the resolution is at least three orders of magnitude better, and analysis of reflection spectra can lead to a precise determination
of the cavity QED parameters g, , and . The first measurements reported in
literature using this kind of technique were reported in late 2007 and can be found
in Ref. [22] and [23]. This ends the summary of the physics behind cavity QED,
and the next section introduces the solid-state emitter used in this work.

1.2

Self-Assembled Quantum Dots

The atomic-like system used for all of the experiments in this dissertation is the
self-assembled quantum dot. The main types of quantum dots are electrostaticallydefined quantum dots, colloidal quantum dots, and self-assembled quantum dots.
Electrostatically-defined QDs have been the focus of significant research efforts
in the physics of electronic transport, and some early experiments within this
system include measurements of Coulomb Blockade and the Kondo Effect [24].
On the other hand, self-assembled and colloidal QDs are of interest primarily for
optical physics experiments. While colloidal QDs are of interest for cavity QED,
self-assembled QDs are the focus of this dissertation, and the rest of this section

28

1 Introduction
is devoted to their description.

1.2.1

InAs Quantum Dot Growth

Self-assembled quantum dots are grown using high quality, semiconductor thin
film techniques such as molecular beam epitaxy (MBE) or chemical vapor deposition. The highest quality QDs are grown in the nearly defect free environment
of MBE in the Stranski-Krastanow growth mode [25], and they will be the type
used in this work. MBE growth consists of evaporating a small flux of material
from a heated source flange and depositing it onto a heated substrate. The high
temperature enables the atoms to diffuse over the surface of the substrate, creating an ordered epitaxial layer. Growth of InGaAs QDs begins by depositing a few
layers of GaAs (typically 100 nm) on a GaAs substrate. This initial layer is called
the smoothing or buffer layer. Then, Indium and Arsenide are deposited simultaneously on the smooth GaAs surface. Because InAs has a slightly larger lattice
constant than GaAs (by about 7%), the InAs becomes elastically strained, and
the strength of this strain grows quadratically with the film thickness [26]. The
first 1.60.1 monolayers of InAs form a thin quantum well type structure, called
the wetting layer. After the formation of the wetting layer, the strain becomes
large enough that it is energetically favorable for the InAs to clump into small
islands [27, 28] as shown in the atomic force microscope (AFM) image in Figure

29

1 Introduction
1.7. The location of these islands is random in the plane perpendicular to the
growth direction and the density depends strongly on the growth conditions. The
size of each island also depends on the growth conditions, but they look roughly
like pancakes with a thickness of 2 nm in the growth direction and a lateral
diameter of 25 nm. Finally, the islands are covered with more GaAs, called the
capping layer. Then the partially-covered island (PCI) technique is used to anneal
the QDs. This technique causes an infusion of Ga into the InAs QD and changes
its shape. These effects combine to move the optical transitions into the near
infrared regime ( 950 nm), where Si-based detectors still work with tolerable
efficiencies.

1.2.2

Optical Properties: Theory

Because the band gap of InAs is smaller than that of GaAs, an electron or hole
moving through the material would feel a 3-D confining potential at the location
of the QD. This 3-D confinement creates a zero-dimensional density of states for
electrons similar to that of atoms, which is why these types of QDs are often
referred to as artificial atoms. It is important to note however, that within the
small size of the QD there exist between a thousand and ten-thousand In, Ga,
and As atoms, making the QD truly mesoscopic.
The single particle energy states within the QD are often calculated via an

30

1 Introduction

Figure 1.7: An atomic force microscopy (AFM) image of an uncapped layer of


InAs QDs. The image size is 1 m by 1 m. Courtesy Antonio Badolato.

31

1 Introduction
effective potential formulation. At the simplest level, this effective potential is
taken to be a square well (quantum well) in the growth direction (z) and a harmonic oscillator in the plane (x, y) [29, 30]. The effective potential for the hole is
exactly the same as for the electron except with a different effective mass. The
strong confinement in the z direction due to its much smaller length scale allows
only one confined state for the electron and hole with energies Eze,h . Since the z
direction doesnt lead to anything interesting, the z confinement energies can be
added to the band gap of the quantum dot material Eg to form an effective band
gap energy,
Eg,ef f = Eg + Eze + Ezh .

(1.50)

The value of Eg,ef f is typically 1.25 eV for InAs QDs [31]. The weaker confinement in the lateral direction leads to the QD shell type structure. Following the
work of Ref. [32], the lateral confinement energies for electrons and holes are
e,h
El,m
= ~ e,h (l + m + 1) l, m = {0, 1, 2, . . .},

(1.51)

where l and m are quantum numbers for the motion, though not in the obvious x
and y directions, but rather in terms of degrees of freedom which facilitate calculation of the orbital angular momentum. e,h are the curvatures of the harmonic potentials for electrons and holes and are significantly different due to the difference
of effective masses (~e 30 meV, ~h 4 meV, me 0.063mo , mh 0.51mo ).

32

1 Introduction

d
p
s

2,0

1,1

CB

0,2

1,0 0,1
0,0

Eg eff
s
p
d

0,0
1,0 0,1
2,0 1,1 0,2

VB

Figure 1.8: The band diagram of quantum dot energy levels in the x y plane.
Each confined state is labeled by the quantum numbers (l, m) as well as the shell
n. In addition, the allowed interband optical transitions s s and p p are shown
as dotted lines.
Since the potential is harmonic, the spacings between levels are equal as shown in
Fig. 1.8.
In the parlance of atomic physics, this gives rise to shells labeled as n = l + m
with n = 0, 1, 2 corresponding to the s, p, and d shell respectively as shown in
Fig 1.8. For the weak excitations considered in this dissertation, these three shells
are the only relevant shells, and all higher order levels can be neglected. In this

33

1 Introduction
basis, the orbital angular momentum is
Le,h
l,m = ~(l m),

(1.52)

where (+) is for electrons and (-) is for holes. Since the system is two-dimensional,
the orbital degeneracy of the nth shell is easily seen to be n + 1. For example, the
p-shell is composed of two energetically degenerate states | 1, 0i and | 0, 1i with
angular momenta +1 and 1 in units of ~ whereas the d-shell is composed of
| 2, 0i, | 1, 1i, and | 0, 2i with angular momenta +2, 0, and 2 respectively. These
states are shown schematically in Fig. 1.8. The number of particles in each orbital
state is determined by the Pauli Exclusion Principle, which will depend on the
allowed spins. In strained zinc-blende structures like InAs/GaAs quantum dots or
quantum wells, the light-hole band (Jz = 1/2) is energetically split off from the
heavy-hole band1 (Jz = 3/2) and can be neglected for low-energy physics near
the point at low temperature [33]. Thus, a single-band approximation for the
valence band is reasonable and holes are taken to have spin 3/2. Electrons in the
conduction band have Sz = 1/2, so the number of particles for both electrons
and holes that are allowed in a single orbital level is 2, leaving a total degeneracy
of the nth shell to be 2(n + 1).
The dominant allowed interband optical transitions are between states with
the same (l, m) [30]. A further requirement is that the total spin Mz = Sz + Jz
1

The size of the splitting is proportional to L2


z , where Lz is the height of the QD along z.

34

1 Introduction
must be 1, creating a right ( + ) or left ( ) circularly polarized photon [30].
Thus, a spin +1/2 (-1/2) electron can only recombine with a spin -3/2 (+3/2)
hole. The s s and p p transitions are shown in Figure 1.8. To calculate
the energies of photons emitted by such transitions, this simple model must be
expanded to include direct Coulomb interactions between carriers. It has been
shown that the energy scale of these interactions is small enough that most of
the physics can be captured in a simple Hartree perturbation [29, 30]. Using this
more complete model, the energies of states as well as recombination energies can
be calculated in a simple way.
The lowest optical excitation is a single exciton X 0 composed of one electron
and one hole. Since there are two spin possibilities for each carrier, there are
four degenerate X 0 states. There are two optically-bright states corresponding to
Mz = 1 as well as two levels with Mz = 2 that are dark. Even though these
states are dark, they can play a significant role in the optical decay dynamics of
the bright X 0 states and this is discussed in detail in Chap. 7. In the harmonic
potential model, X 0 will have an energy (relative to Eg,ef f )
EX 0 = ~e + ~h V eh ,

(1.53)

where V eh = (e2 /4) e h | 1r |e h is the direct Coulomb term and acts to


reduce the total energy. Thus, V eh can be interpreted as the exciton binding

35

1 Introduction
energy and is roughly 20 meV for the type of QDs considered here. Because X 0
decays to a vacuum ground state, a photon emitted by the recombination of a

single exciton will have energy EX


0 = EX 0 . In the laboratory, photons emitted

by an X 0 recombination are measured to be between 900 nm (1.38 eV) and 970


nm (1.28 eV).
The next lowest optical excitation that is charge neutral is the biexciton XX 0 .
This state is composed of two electrons with opposite spin in the s-shell and two
holes with opposite spin in their s-shell. Because both shells are filled there is
only one XX 0 , but since it decays to X 0 this decay can occur two ways as shown
in Fig. 1.9a. The energy of XX 0 is
EXX 0 = 2~(e + h ) 4V eh + V ee + V hh ,

(1.54)

and the recombination energy to X 0 is

e
h
eh
EXX
+ V ee + V hh .
0 = EXX 0 EX 0 = ~( + ) 3V

(1.55)

It is useful to consider the this decay relative to the X 0 emission,

eh
EXX 0 = EXX
+ V ee + V hh .
0 EX 0 = 2V

(1.56)

In the InAs/GaAs material system, the large hole mass causes its wavefunction to
be more localized than that of the electron. As a result, V hh > V eh > V ee can be
taken to be generally true . Furthermore, Ref. [29] predicts that V hh +V ee 2V eh ,
36

1 Introduction

a.
XX

b.
|

Xo |

E2

|
s
s

E1
|

|vac

Figure 1.9: a). A level diagram of the biexciton cascade. The levels are labeled
as XX 0 or X 0 as well as with the spin states of the electrons (, ) and holes
(, ). Each transition is either right ( + ) or left ( ) circularly polarized. The
cascaded emission leads to photons that are maximally polarization entangled (a
Bell state) if E1 = E2 . b). A level diagram for the negatively charged trion labeled
with electron and hole spins as well as the polarization of each transition.
so that the biexciton transition should be within 1 meV or so of the exciton
transition. In practice, this is usually true, although the biexciton transition is
almost always found to be to the low energy side of the exciton transition for the
QDs in the experiments in this work.
The cascaded decay of XX 0 X 0 | vaci is of great interest to quantum
information because the polarization of the photon from the biexciton decay is
correlated to that of the exciton decay as shown in Fig. 1.9a. If the biexcitons
recombination energy were exactly equal to that of the exciton (E1 = E2 in
Fig. 1.9a or EXX 0 = 0), then the full quantum state of the two photons would

37

1 Introduction
be
i
1 h
| i = + 1 2 + 1 + 2 ,
2

(1.57)

where the numbering denotes the time ordering of the decay. This state is very
special because it is a maximally entangled Bell state [34, 35]. In practice, observing this state is difficult experimentally because the QD is often slightly elliptical2 .
This breaks the energetic degeneracy of the two bright X 0 states by an energy 2b
and causes the emission to become linearly polarized [36]. This effect is explained
in detail in App. A. Therefore, the two main obstacles to achieving entangled
photon pairs from the biexciton cascade are achieving EXX 0 = 0 and b = 0.
Two groups have found ways around these obstacles, and entangled photons were
demonstrated in 2006 in the experiments of Ref. [37] and [38].
Using the simple harmonic model, it is also easy to describe the trion transitions of a quantum dot. A trion is a singly charged excitation and comes in two
varieties. The negatively charged trion X is composed of two electrons with opposite spin in the s-shell and one hole in the s-shell as shown in Fig. 1.9b, whereas
the positively charged trion X + is composed of two holes with opposite spin and
one electron, all in their s-shells. It is important to note that trions do not have
optically dark transitions because, using the negatively charged trion as an example, the two electrons must have opposite spin and therefore a bright transition
2

InAs QDs tend to be elongated along the [110] crystal axis.

38

1 Introduction
always exists for any hole spin. Using the same analysis as for the biexciton, the
relative transition energies of the trions can be expressed as
EX = V eh + V ee ,

(1.58)

EX + = V eh + V hh .

(1.59)

Using the rule V hh > V eh > V ee again, the X transition is expected to be lower
in energy than X 0 while the X + transition should be higher in energy. In fact, this
is exactly what is experimentally observed and will be shown in the next section.
These states are of great interest for quantum information applications as well
as fundamental physics because they decay to a ground state which is composed
of a single trapped spin, and information about that spin can be obtained through
the polarization of the emitted photon, as shown schematically in Fig. 1.9b.
These states can be created with a small probability by optical excitation or by
electrically-gating the QD, which will be discussed in detail in Sec. 1.2.5. Many
interesting experiments have been performed using these states including electron
[39] and hole [40] spin cooling as well as time-resolved measurements of the electron spin relaxation [41, 42]. While the model detailed above is relatively simple,
it captures most of the physics behind the optical properties of QDs. In recent
years, several groups have used numerical theoretical models which include the
full exchange interaction and motional correlations to predict the fine structure

39

1 Introduction
and optical properties of QDs with as many as seven carriers down to the eV
energy scale. These predictions have been shown to match up extremely well with
experiments on single QDs [43]. While these models are useful, they are well
beyond the scope of this work.

1.2.3

Microphotoluminescence Spectroscopy

The standard technique to measure the emission from single quantum dots
is called microphotoluminescence (-PL) spectroscopy. The way this is implemented experimentally for the measurements in this dissertation is diagrammed
in Fig. 1.10. Non-resonant excitation is done using a pulsed Titanium Sapphire
Laser (Spectra-Physics Tsunami). This laser has a 150 fs pulse width, 80 MHz
repetition rate, and average power of 1.5 W. The wavelength is tunable from
near 700 nm to 940 nm, but it typically used at 855 nm for wetting layer excitation
or 780 nm for above band gap excitation. Pump powers onto the sample surface
are less than 100 W and usually near 100 nW. The excitation light is passed
through a set of variable attenuators and then focused onto the sample surface at
normal incidence. An inexpensive charge-coupled device (CCD) array is used to
image the sample surface and enables the location of specific microcavities. Because the sample is located inside the cryostat and is a few mm from the cryostat
window, an extra long working distance objective is used to achieve a high numer-

40

1 Introduction
ical aperture. The objective used in this setup is a Mitutuyo M Plan Apo NIR
series 50X objective with N.A. = 0.42, 4.0 mm focal length, and 17.0 mm working
distance. The objective is infinity-corrected so that the collected luminescence is
collimated. The position of the beam can be navigated across the sample surface
by using the xyz manual micrometer stage that the objective is mounted to. Realignment of the beam using steering mirrors (not shown) is necessary to achieve
a circular spot. The objective is also mounted to a high precision xy piezo stage
(Physik Instrumente P-733.2CD) with 0.3 nm resolution and 100 m by 100 m
travel range. This can be used to spatially address single quantum dots because
the excitation diameter is on the order of the diffraction limit and the density of
QDs is usually around 10 m2 .
The emission from the QDs is spectrally separated from the reflected laser spot
by a dichroic beamsplitter (BS). This BS is designed to reflect (R > 0.99) light
with > 900 nm and transmit light with < 900 nm. From there, the emission
is directed into the monochromator or to a single mode (SM) fiber connected to
an avalanche photodiode (APD). The APD detects single photons near 940 nm
with 30% detection efficiency. The monochromator (Horiba Jobin-Yvon model
1250M) has a 1.25 m focal length with a f/9.0 F-number. The grating is a blazed
holographic plane grating with a groove density of 1200 lines/mm and is blazed at
900 nm. The monochromator has two outputs. The primary output is to a liquid

41

1 Introduction

Monochromator
Pulsed
Ti:Sa
Tunable CW

APD
APDs
SM Fiber

MM Fiber
BS
Imaging
CCD

CCD
Cryostat
Dichroic BS
Obj

Var Attn.

Figure 1.10: A schematic of the experimental setup for PL showing all relevant
optics, devices, and lasers.

42

1 Introduction
nitrogen cooled CCD array (Princeton Instruments Model Spec-10 100BR/LN)
with a quantum efficiency of 60% at 940 nm. It is composed of an array of 1340
by 100 pixels. In all, the spectral resolution is 30 eV (0.015 nm) and the
total detection efficiency including optical losses is 15 %. The array has a total
bandwidth of 16.95 nm. The other monochromator output is passed through an
exit slit and focused into a multi-mode (MM) fiber. The output bandwidth can be
adjusted by changing the width of the exit slit and is usually set to 0.1 nm. This
MM fiber is coupled to a 50:50 fiber beamsplitter and each output is connected
to an APD.
A typical spectrum of the photoluminescence from a single QD is shown in
Fig. 1.11. Emission from the states discussed in Sec. 1.2.2 are labeled accordingly.
The spectral locations of the different states correlate well to the predictions of
the simple model in the last section where the X + peak was expected to be found
towards higher energies and the X peak towards lower energies.

1.2.4

Non-classical states of Light

An important aspect of quantum dots is that they emit single photons, which
are a very non-classical source of light.

This was first demonstrated in the

InAs/GaAs quantum dot system in 2000 in Ref. [18]. It had previously been
demonstrated using atoms [44], a stored ion [45], a molecule [46], a different kind

43

1 Introduction

Counts Hs-1L

2000

1500
1000

500
0

1.355

XX

2-

1.360
1.365
Energy HeVL

1.370

Figure 1.11: The optical spectrum of photoluminescence from a single QD using


the PL setup under 50 nW excitation at 780 nm. Emission from several
different states is visible and each known state has been labeled. The location of
the different peaks very closely matches the simple analysis of Sec. 1.2.2. XX n is
used to denote XX 0 , XX + , and XX .

44

1 Introduction
of semiconductor QD [47], and a nitrogen-vacancy center in diamond [48, 49].
Since then, single photon emission has been shown for a variety of defect centers,
dopants, and even carbon nanotubes [50]. The standard way to verify emission of
single photons is by doing a Hanbury-Brown and Twiss type experiment [51]. In
this experiment, the collected luminescence is separated at a 50:50 beamsplitter
and photons are detected at two APDs, one acting as the start signal and the
other acting as the stop signal. A time-to-amplitude converter (TAC) builds a
histogram of start-stop events, labeled by a delay , which when normalized is
equivalent to the second-order photon correlation function g (2) ( ) in the limit that
the reciprocal of the average counting rate is much longer than the measured time
separation [51]. This requirement is easily met for all g (2) ( ) measurements in
this dissertation. g (2) ( ) for a single mode field is defined by

a
(t)a
(t
+

)a(t
+

)a(t)
g (2) ( ) =
.
ha (t + )a(t + )i ha (t)a(t)i

(1.60)

It is easy to show that as a consequence of the Cauchy-Schwarz inequality, classical


light must obey g (2) ( ) g (2) (0) and g (2) (0) 1. A thermal distribution of light
will give rise to photon bunching with g (2) (0) = 2 while Poissonian distributed
light (like a laser) gives rise to completely uncorrelated photons with g (2) (0) =
g (2) ( ) = 1. The transition between these two states of light is discussed at length
in Chap. 3 and calculated in Appendices B and C.

45

1 Introduction
On the other hand, an n-photon Fock state of light has
g (2) (0) = 1

1
,
n

(1.61)

and the 1-photon state has g (2) (0) = 0, called anti-bunching. Intuitively, it is
easy to see why g (2) (0) = 0 for a single photon stream. Because there is only 1
photon, quantum mechanics states that it can only be measured at one detector
after the beamsplitter. Therefore, there cant be any coincidences at = 0. Thus,
g (2) ( ) is an important measurement that can distinguish between classical and
quantum states of light. Additionally, a measurement of g (2) (0) < 0.5 proves
the light is primarily composed of single photons and acts as the experimental
benchmark.
Using the experimental setup of Fig. 1.10, g (2) ( ) can be measured for a single
QD and the result is plotted in Fig. 1.12. The value at = 0 is measured to be
0.0870.017, much less than 0.5. This small residual value is due to emission that
is collected from other nearby QDs.

1.2.5

Quantum Dots in Diode Structures

Quantum dots are also interesting to study because they can be easily incorporated into an electrically-gated semiconductor heterostructure like a pin diode.
The quantum dots are affected in three significant ways when placed in a diode

46

1 Introduction

1.0

gH2LHL

0.8
0.6
0.4
0.2
0.0

-80

-60

-20
HnsL

-40

20

40

Figure 1.12: A measurement of g (2) ( ) for a single QD excited by the pulsed Ti:Sa
laser. The value of g (2) (0) is 0.0870.017, much less than the value of 0.5 required
to prove single photon character.
structure, and the relative importance of each depends on the exact structure of
the diode. All three physical effects can be understood by considering the band
structure of the diode system as shown in Fig. 1.13.
Firstly, forward biasing a pin diode (which lowers the energy of the p-type
region with respect to the n-type region) will cause electrons to move from the
n-type region to the p-type region and holes to move in the opposite direction, creating an electric current. They will meet in the intrinsic region containing the QDs
and can recombine optically. Thus, the QDs will emit light even though carriers
arent being excited optically but rather by the current passing through the in-

47

1 Introduction

Vapp = 0
Ef

Electric Field
p-type
n-type

QD
Ef

Vapp > 0

Figure 1.13: Band structure along the growth direction of a QD embedded in


a pin diode. The top band structure corresponds to no applied bias while the
bottom corresponds to a small forward bias. Also shown are one confined level for
electrons and one for holes along with the Fermi energy Ef of the n-type region.

48

1 Introduction
trinsic region. This is called electroluminescence and was first demonstrated using
single QDs in 2002 by Yuan and coworkers [52]. An experimental demonstration
of electroluminescence using the setup described earlier is shown in Fig. 1.14a.
In this experiment, the spectrum of a several QDs interacting with the modes of
a micropillar cavity is monitored as a function of applied bias under no optical
excitation. Once current begins to flow, a large onset of luminescence is detected
corresponding to the optical relaxation of electrons and holes flowing through
the device. This effect is explored in more detail in Chap. 8 where it is used to
demonstrate a bright single photon LED.
Secondly, the sharply slanted band structure indicates that a strong electric
field is dropped over the intrinsic region containing the QDs. As a function of
applied bias Vapp , the electric field has a magnitude Eapp = (Vbi Vapp )/d for
Vapp < Vbi where Vbi is the built-in potential difference and is roughly equal to
the band gap of GaAs (1.5 V) and d is the size of the intrinsic region [53](usually
around 500 nm). Using this type of structure, fields can easily reach values of 300
kV/cm. The transitions of the quantum dot respond to the applied electric field,
and this phenomenon is called the quantum-confined Stark Effect [54]. The
effect is quadratic with the applied field and the change in energy of the emission

49

1 Introduction

1.335
1.3325

1.32

1.318

Energy HeVL

Energy HeVL

1.33
1.3275
1.325

1.316

1.314

1.3225
1.312

1.32
1.3175

2
3
Applied Bias HVL

10

1.328
1.326

1.324
Energy HeVL

6
8
Applied Bias HVL

1.322
1.32

1.318
1.316

2.8

2 -

X
3 X

2.9
3
3.1
Applied Bias HVL

3.2

Figure 1.14: (a) Spectra as a function of applied bias under no optical excitation
for several QDs in a pin structure capable of passing current through the active
region. Once current turn-on is surpassed, electroluminescence from the QDs and
cavity modes is detected. (b) Spectra as a function of applied bias under weak
optical excitation for several QDs in a Stark-tuning device. For small biases, the
electric field is strong enough to break apart excitons which quenches the emission.
Each transition line moves nonlinearly with bias due to the Stark Effect. (c)
Spectra as a function of applied bias under weak optical excitation for one QD
in a charge-tuning structure. As the bias is increased, the QD becomes more
negatively charged and the dominant transition line spectrally shifts with each
additional electron. Each peak is labeled by the initial state of the exciton for
that transition.

50

1 Introduction
is usually expressed as [55]
2
E = pEapp + Eapp
.

(1.62)

In this expression, p represents the permanent dipole character of the electron-hole


pair and has a value near 1.31028 Cm which corresponds to an electron-hole
separation of 8 nm [55]. is a measure of the polarizability of the quantum dot
and has a value near -51033 Cm2 /kV. An experimental demonstration of the
Stark Effect with QDs using the setup described earlier is shown in Fig. 1.14b. The
spectrum of emission for several QDs is monitored as a function of the applied bias
and all of the lines spectrally shift as predicted. The Stark Effect provides a fast
and repeatable spectral tuning method for QDs and has important applications for
achieving spectral resonance in cavity QED (see Chap. 9) as well as for generating
photons of a specific polarization (see Chap. 10).
Lastly, if the QD is near one of the doped layers, resonant tunnelling of electrons or holes is possible. As shown in the top band diagram of Fig. 1.13 at
Vapp = 0, the Fermi energy of the n-type layer Ef is above the confined single
electron level in the QD, but in the lower band diagram under forward bias, Ef
has moved below the confined level. That means that at some bias in between the
two biases in the diagram, Ef was energetically resonant with the s-shell of the QD
and resonant electron tunnelling could take place. Experimentally, measuring the

51

1 Introduction
single QD spectrum as a function of Vapp shows the emission peak spectrally shift
at specific biases, corresponding to resonant tunnelling such that the QD becomes
permanently charged with extra electrons. This was first demonstrated in 2000 in
an experiment by Warburton and coworkers [56] and is shown for a single QD in
Fig. 1.14c. The spectrum from a single QD is monitored as a function of applied
bias, and the peak clearly spectrally shifts at specific biases. The charge state
of each transition is labeled and closely matches the analysis of Sec. 1.2.2. This
ability makes the QD a very special atom in that its charge state can be easily
tuned with an applied bias and enables optical measurements of the interactions
between a small number of carriers. QD charge-tuning is explored in more detail
in Chap. 9 where it is combined with micropillar cavities to enable efficient single
photon-single spin interaction.

52

Chapter 2
Self-Tuned Gain in Photonic
Crystal Nanolasers
This chapter is based on Physical Review Letters 96, 127404 (2006). Copyright
2006 by the American Physical Society.

2.1

Introduction

Optical microcavities [57], in particular photonic-crystal membranes, offer the


ability to create new, efficient optical sources of specified wavelength through the
control of the dielectric environment. In photonic-crystal membranes, this control is obtained by making a periodic lattice of the index of refraction which,

53

2 Self-Tuned Gain in Photonic Crystal Nanolasers


in analogy to the formation of band gaps for electrons in semiconductors, creates
band gaps for photons [58]. Cavities are formed like defects in a semiconductor by
breaking this periodicity, forming a localized state inside the band gap. Extremely
low-threshold lasers can result from the appropriate match of small mode-volume
photonic-crystal cavities to optically active material. One measure of the optical efficiency of a laser is the spontaneous emission (SE) coupling factor . The
theoretical limit is one, corresponding to the hypothetical case of thresholdless
lasing. Microdisk [59, 60] and photonic-crystal lasers (PCLs) [61] with quantum
wells (QWs) or a high density of quantum dots (QD) as gain medium show values
ranging from 0.1 to 0.2, but a pronounced lasing threshold behavior remains [59
64]. For non-lasing devices values above 0.9 have been estimated [65]. Here we
report on PCLs sustained by only 2-4 QDs as active gain material showing ultralow lasing thresholds and a of 0.85. This is not expected to be possible within
the simple atomic-like model of a quantum dot (see Sec. 1.2.2) and we propose
that because of the presence of the photonic crystal, carrier-carrier interactions
play an important role in tuning the gain structure of QDs.

54

2 Self-Tuned Gain in Photonic Crystal Nanolasers

2.2

Photonic Crystal Cavity Design

To achieve lasing operation, both the gain medium and the cavity must have
exceptional features. The cavity design shown in Fig. 2.1a is based on a line defect
of three missing air holes (L3-type) within a triangular lattice formed in a 126 nm
GaAs membrane, acting as a mono-mode waveguide in the QD emission range.
The center of the membrane contains a single layer of low-density (8 2 m2 )
InAs QDs grown by the partial covered island technique [66]. This leads to QDs
with shallow confinement energies and a pronounced density of extended wetting
layer (WL) states [67]. Neighboring holes in the line have been shifted outwards
(as indicated by white arrows in Fig. 2.1a) to produce better confinement of the
mode [69]. To achieve optimal coupling of the lasing mode with the active QD
gain material, two new design concepts have been employed. First, the two holes
nearest to the defect have been reduced by 20% in size to decrease their overlap
with high field regions [70]. Second, the width w of the channel region containing
the three missing holes has been increased by 30 nm, as illustrated by the blue
shading in Fig. 2.1a. In culmination, these design concepts achieve a mode volume
of V = 0.68 (/n)3 and an effective refractive index nef f of 2.9, a very high value
for this class of porous cavity membranes. In addition, this design ensures that
the field maximum is in excellent spatial overlap with the gain medium while the

55

2 Self-Tuned Gain in Photonic Crystal Nanolasers

Figure 2.1: (a) Scanning electron micrograph around the defect region of a PCL
laser with a lattice constant of 260 nm. (b) Plot of the electric field intensity of the
lasing mode as calculated from 3D finite-difference time-domain simulations [68].
(c) Normalized lasing mode spectra for 9 different PCL devices taken at a pump
power of 5 W, T = 4.5 K. (d) Measured cavity mode intensity as a function of
polarizer angle showing a polarization ratio of 25:1.

56

2 Self-Tuned Gain in Photonic Crystal Nanolasers


overlap with the air-semiconductor interface is drastically reduced. Such interfaces are detrimental since the optical properties of QDs near surfaces degrade
critically [71], and surface defect states can be optically absorbent.
The wavelength of the PCL mode can be tuned through its lattice parameters
to the SE wavelength range of the QD gain medium. We studied 22 PCL devices
ranging from 908 to 975 nm covering the ground state (s-shell, centered at 965
nm) and excited state (p-shell) of the embedded QDs. The QDs were excited via
the GaAs matrix by a continuous wave (cw) 780 nm laser diode using a microscope
objective. The effective spot size including carrier diffusion was estimated from
lateral scanning of the single QD emission to be 3.4 m. Despite the fact that
there are on average only 2 to 4 QDs located within the mode volume and that the
exciton transitions of those QDs are in general off resonance with the cavity mode,
we claim that most devices show pronounced single mode lasing (Fig. 2.1c-d).

2.3

Experimental Lasing Results

Before discussing the remarkable laser gain process, we present three-fold experimental evidence for lasing operation of these devices. First, we recorded the
output power of the cavity mode as a function of cw-excitation pump power.

57

15

1.2

start

Threshold
124 nW

10

g (=0)

stop

1.1

Intensity (arb. units)

2 Self-Tuned Gain in Photonic Crystal Nanolasers

0
0

100

200

0
0

1000

2000

3000

1.0
0

P/Pthr

Pump power (nW)

Figure 2.2: (a) Lasing mode intensity (solid dots) and SE background (open dots)
as a function of excitation pump power for a cavity emitting at 960 nm. Red line:
linear fit. (b) Magnification of threshold region. (c) Photon correlation function
g 2 ( = 0) of the cavity mode as a function of normalized pump power P/Pthr .
Absorbed pump power at threshold, Pthr , for this device emitting at 940 nm is 25
nW. T = 4.5 K.
Figure 2.2a shows the resulting L-L-curve for a device emitting at 960 nm (solid
dots). Above a certain threshold pump power, a linear increase in output intensity
is observed, as indicated by the red line. The conventional estimation of the cw
lasing threshold, obtained by extrapolation of the red line to zero output power
(Fig. 2.2b), yields a value of 124 nW (1.3 W/cm2 ) corresponding to an absorbed
pump power of only 4 nW (49 mW/cm2 ). This is an improvement of 2 to 3 orders of magnitude over prior reports [57, 5964]. We measured similar ultra-low
threshold values for temperatures between 4 and 80 K. For higher temperatures
the QD emission is quenched due to the thermal break up of excitons. Figure 2.2b

58

2 Self-Tuned Gain in Photonic Crystal Nanolasers


shows no sharp threshold but a soft turn-on of the cavity mode intensity, in accordance with theory for a large SE coupling efficiency [72, 73].
A second characteristic of a laser is the emission linewidth narrowing profile.
Below and above threshold the linewidth E varies inversely with the output
power [74]. At threshold, the phase transition into lasing leads to a pronounced
kink in the profile [75], as demonstrated in Fig. 2.3a. To compare the linewidth
narrowing with the intensity data (Fig. 2.3b) the linewidth is plotted as a function of pump power. In principal, a contribution to linewidth narrowing could
also arise from loss saturation effects of QDs. These effects, however, do not lead
to a kink in the L-L curve, as has been confirmed by additional measurements
performed on non-lasing devices. Those measurements follow the dotted blue lines
in Fig. 2.3. Such behavior corresponds to operation in the light-emitting-diode
(LED) regime and is similar to the standard Schawlow-Townes linewidth narrowing below threshold, where there is a constant cavity loss but an increase in gain
[74]. At pump powers of 3-5 W, an overall saturation of the intensity sets in due
to saturation of the QD gain medium. We found that maximum output powers
for devices operating in the LED regime are an order of magnitude lower compared to those reaching into the lasing regime. For the best lasing devices the
E/E ratios reach values up to 32000 and the corresponding Q factors measured
at pump powers below lasing threshold range between 7800 and 18000.

59

2 Self-Tuned Gain in Photonic Crystal Nanolasers

Linewidth (eV)

10

200

10

10

10

10

150

Intensity (arb. units)

100
6

10

10

b
= 0.85

10

1.0
3

0.9

10

0.8

0.5

10

10

10

0.1
10

Pump Power (nW)

10

10

Figure 2.3: Cavity mode linewidth (a) and intensity (b) as a function of excitation
pump power (solid dots). The x-axis displays optical power sent into the cavity.
The corresponding absorbed pump powers are 3.6 % of these values. Solid lines:
Rate equation model solutions for various values of . Red line: Best fit corresponding to = 0.85 0.03. Dotted blue lines indicate the measured behavior
for non-lasing devices.

60

2 Self-Tuned Gain in Photonic Crystal Nanolasers


To provide a third and direct proof that lasing action occurs, we measured
the second order photon correlation function g (2) ( ) = hI(t)I(t + )i/hI(t)ihI(t)i,
where hI(t)i is the expectation value of the intensity in the cavity mode at time t,
as a function of excitation pump power (Fig. 2.2c). The cavity mode emission was
spectrally filtered with a 0.5 nm bandpass and sent into a Hanbury-Brown and
Twiss (HBT) interferometer, consisting of a 50/50 beamsplitter and two single
photon counting detectors, as shown schematically in the inset of Fig. 2.2c. Photon coincidences were recorded electronically as a histogram of start-stop events.
A characteristic feature of a laser is the phase transition from a thermal into a
coherent light state as a function of pump rate [76]. For a detailed theoretical
discussion of the photon statistical transition of a laser, see App. B and C. The
measured photon bunching signature (g (2) ( = 0) > g (2) ( )) around the threshold of the PCL indicates the thermal character of the emitted light field. Below
threshold g (2) ( = 0) should theoretically approach a value of two corresponding
to an ideal thermal state. However, the measured values for g (2) ( = 0) drop down
at lowest pump powers due to a decrease of the coherence length of the PCL emission approaching the temporal resolution limit (450 ps) of the HBT setup [77].
At pump powers above threshold the cavity mode picks up phase coherence due
to stimulated emission, leading to a transition into a coherent light state where
g (2) ( = 0) approaches unity (Fig. 2.2c). In addition, we performed experiments

61

2 Self-Tuned Gain in Photonic Crystal Nanolasers


on a series of photonic crystal cavities designed to span the range from = 0.3
to = 0.9 as well as on a commercial semiconductor laser diode confirming that
Fig. 2.2c is indeed characteristic for lasing action [77].

2.4

Discussion

To extract the fraction of SE into the lasing mode with respect to SE into
all available modes, we use a coupled rate equation model for the carrier density
N and the photon density P of a semiconductor laser diode [74]:
dN
dt
dP
dt

N
N

vg gP
sp nr
P
N
= vg gP +

sp tc
= Rp

(2.1)
(2.2)

where Rp is the pump rate, vg the group velocity, the confinement factor, 1/tc
the cavity photon loss rate, 1/tsp the spontaneous, and 1/tnr the non-radiative recombination rate. Steady state solutions have been found assuming a linear gain
function g = a(N Ntr ), where a is the differential gain and Ntr the transparency
carrier density. The solid lines shown in Fig. 2.3b are solutions of the rate equations for various values of keeping all other parameters constant. Parameters
used are determined as follows: A tc of 4.1 ps corresponds to the measured Q =
9800. Lifetime measurements at the cavity mode frequency reveal tsp = 0.14 ns.
62

2 Self-Tuned Gain in Photonic Crystal Nanolasers


The tnr has been taken to be 10 ns and does not influence the fit considerably
[72]. A of 0.004 has been estimated from the intersection of the mode volume
with the single layer of QDs and vg = c/nef f = 1 1010 cm/s. Best fits have
been found for a = 4.5 1010 cm2 and Ntr = 5.5 1014 cm3 . The corresponding
material gain of 2 105 cm1 is in good agreement with values for other InAs QD
lasers [78]. The transparency (threshold) carrier density corresponds to about 8
(18) electron hole pairs within the mode volume. The red curve in Fig. 2.3b is the
best fit for = 0.85 0.03 which implies an ultra-high SE coupling efficiency of
85%. Note that the nonlinear kink in the intensity data occurs at the same pump
power values as the kink in the linewidth narrowing (Fig. 2.3a) and the soft turnon starting at about 30 nW incident (1 nW absorbed) pump power corresponds
to the definition of the lasing threshold for high- lasers [72].
In general, the factor depends on the number of optical modes and the width
of the SE spectrum. Besides the fundamental cavity mode, the density of radiation modes is strongly suppressed. This is supported by measured lifetimes up
to 10 ns for single QDs located outside of the cavity mode region but within the
photonic crystal membrane. These values are one order of magnitude larger than
SE lifetimes (1 ns) measured in unprocessed areas of the wafer. The observed
inhibition directly reflects the low density of radiation modes within the photonic
band gap. It has been predicted that non-degenerate fundamental modes in PCL

63

2 Self-Tuned Gain in Photonic Crystal Nanolasers


structures with similar mode profiles can achieve values up to 0.87 assuming
a SE bandwidth of 25 nm [79]. While such a bandwidth can be achieved with
QWs or larger ensembles of QDs [64], it is indeed very surprising that at our low
QD densities such an efficient QD-cavity coupling occurs. The sharp exciton QD
emission is statistically distributed over 50 nm making a simultaneous spectral
and spatial coupling very unlikely (< 1%). Control samples containing zero QDs
but a dominant WL emission at 860 nm do not show lasing or any decoration of
the cavity mode. If there is only one QD spatially overlapping with the cavity
mode, as has been confirmed by deterministic QD positioning [80], decoration
of the mode occurs even for up to 20 nm spectral detuning [68]. Therefore, the
quantum dot cavity interaction has to be of an indirect nature.

2.5

Self-Tuned Gain

To obtain insight into the underlying gain mechanism, we studied the spectral
properties of a single QD (Fig. 2.4a-b). The sharp transitions of the s-shell emission evolve with increasing pump power into spectrally broad bands (Fig. 2.4a
top). This broad background is significant at pump powers around the lasing
threshold (Fig. 2.4b) and demonstrates that SE from a single QD occurs in a large

64

b 150 nW0

s-shell
10

150 nW
X- XX
1.5 nW

+
1

1.33

1.34

Cavity
mode
s-shell

XX

200 nW

10
1.33

1.34

II

WL

QD

cavity
mode

WL

II

QD4

QD2

QD3

QD1

p-shell

II

III
1.30

10

XX
X-

10

Intensity (arb. units)

3 W

SE Lifetime (ns)

DOS (arb. units) Intensity (arb. units)

2 Self-Tuned Gain in Photonic Crystal Nanolasers

1.35

1.40

Energy (eV)

1.45

0.1

1.32

1.33

Energy (eV)

Figure 2.4: (a) Single QD spectra recorded at pump powers below (1.5nW), around
(150 nW) and above (1.5 W) the lasing threshold of the PCLs. (b) 150 nW spectrum on a log scale illustrating the broad background. (c) Schematic illustration
of the possible transitions of a single QD adjacent to a WL (inset) and the corresponding density of states (DOS). (d) PCL spectrum taken at 200 nW (right
y-axis) and measured SE lifetimes as a function of energy using a 0.1 meV broad
spectral window (left y-axis). Sharp lines right next to the cavity mode have been
tentatively labelled by QD1-QD4.

65

2 Self-Tuned Gain in Photonic Crystal Nanolasers


frequency range. Several effects are responsible for the rich spectral properties.
Multiple charge states of the exciton (X + , X 0 , X ) [56] and bi-excitons (XX) [81]
give rise to a few recombination lines. Each transition deviates from a Lorentzian
line shape in form of a broad background extending about 4 meV (not shown)
due to electron-acoustic phonon coupling [82]. It is furthermore known that in
these QDs with shallow confinement transitions between localized QD states and
extended WL states provide a quasi-continuum spectrally covering the energy region between the WL and the p-shell [67], as has been indicated by regions I
and II in Fig. 2.4c, respectively. At pump powers around the lasing threshold,
these states are already appreciably populated. In this case, when multi-exciton
states are involved, the recombination energy of the s-shell strongly depends on
the exact exciton number and spin configuration in the p-shell [83] and extended
states. Therefore, this intra-dot Coulomb interaction produces a configuration
mixing between s- and p-shells and couples the extended state continuum around
the p-shell into the energy region of the s-shell emission, thereby causing a broad
emission background as shown in Fig. 2.4a top and indicated by region III in
Fig. 2.4c. The many-body coupling between s- and p-shells requires the Coulomb
interaction to be included into the emission process, and in this sense it is different
from a two-step, Auger-like relaxation process.
The essential role of the nanocavity in contrast to a large mode volume cavity

66

2 Self-Tuned Gain in Photonic Crystal Nanolasers


is the strong inhibition of sharp exciton transitions up to 1 order of magnitude,
as can be seen exemplarily for QDs 2 and 4 in Fig. 2.4d. Simultaneously, the
combined QD-WL states become crucial since their energy resonance condition
is easily fulfilled, leading to an effective transfer channel into the lasing mode.
Measured SE lifetimes on top of the cavity mode are as fast as 145 ps, which is
close to the resolution limit of our setup (Fig. 2.4d.). Note that not all of the
QDs experience inhibition (see QDs 1 and 3 in Fig. 2.4d). This is expected since
about 20 % of QDs are close to surfaces that lead to additional non-radiative loss
channels [71]. In culmination, the QD confinement potential harvests the excitons
and their coupling to the surrounding matrix efficiently transfers the energy into
the lasing mode.

2.6

Summary and Conclusion

In summary, we observe lasing action as demonstrated by the characteristic


L-L curve, linewidth narrowing profile and photon statistics. We have shown that
lasing in single-mode photonic-crystal cavities does not require an exciton/mode
resonance. These results challenge the conventional QD-laser design based on
incorporating several dense layers of QD material. This has clear technological

67

2 Self-Tuned Gain in Photonic Crystal Nanolasers


implications for future design of ultra-efficient nanoscale lasers, which may now
be formed with only a few, or even single, QDs that, with cooperation from the
surrounding matrix, self-tune their emission into the lasing mode with nearly
perfect efficiency.

68

Chapter 3
Validation of Lasing in PC
Nanolasers
This chapter is based on Applied Physics Letters 91, 031108 (2007). Copyright
2007 by the American Institute of Physics.

3.1

Introduction

Recent advances in the ability to design and fabricate active nanocavities with
tailored modal characteristics [62, 84, 85] have enabled the nanolasers with very
high factors, meaning virtually all spontaneous emission is directed into the lasing mode. Subsequently, substantial reduction in lasing threshold [62, 8488] and

69

3 Validation of Lasing in PC Nanolasers


high-speed modulation [89] have been demonstrated, stimulating further research
towards the highest-performance nanolasers [72, 90]. However, the realization of
high- nanolasers is complicated by the inherent difficulty in characterizing their
behavior. The characteristic nonlinearities in the optical output power around
the lasing threshold become so subtle that conventional criteria used to determine
a lasing threshold are not well defined. Theoretical and experimental studies
on second-order intensity correlation [87, 88, 91] or photon number fluctuation
[76, 92] have demonstrated their usefulness to confirm the high- lasing actions
in the nanocavities [87] and the microcavities [76, 88, 91, 92].

3.2

Line Defect PC Cavity Modes

In this chapter, we have investigated the evolution of the second-order intensity correlation function near the lasing thresholds of PC nanolasers with tailored
modal characteristics. Our high- PC nanolasers are comprised of self-assembled
InAs quantum dots (QDs), with a density of 5 109 cm2 , grown by molecular
beam epitaxy on a (100) GaAs substrate by the partially covered island technique [66]. The QDs show sharp and distinct exciton and muli-exciton spectra,
with clearly delineated s-shell and p-shell regions at 4 K. The nanocavities were
designed as triangular-lattice PC structures with 3, 7, and 11 missing holes in

70

3 Validation of Lasing in PC Nanolasers


the -K direction, denoted as L3, L7, and L11 cavities, respectively. The cavity
resonances are determined by the PC waveguide dispersion and the Fabry-Perot
condition [93]. The finite-difference time-domain simulation shows that the cavity
resonances exist on the dispersion curves of a single-line-defect PC waveguide as
shown in Fig. 3.1(a). Thus, by changing the number of missing holes, the number of modes within the QD s-shell spectrum can be engineered. To match the
s-shell spectrum with the low energy, high-Q even (e#) modes, we specified a
lattice constant of 260 nm, a hole radius of 65 nm, and a thickness of 126
nm. Micro-photoluminescence (-PL) measurements were carried out at 4 K using a He-cryostat and a monochromator under optical pumping with a 780 nm
laser diode. The -PL spectra in Fig. 3.1(b) show that L3, L7, and L11 cavities support 1, 2, and 3 non-degenerate modes polarized along the y direction
in accordance with our cavity design. With a simple mode counting argument,
this can stepwise change from approximately 1 to 0.5 to 0.33, and demonstrate
systematic variation in the g (2) (0) transition near lasing thresholds. For the measurement of g (2) ( ), the -PL was sent to two avalanche photodiodes arranged in
the Hanbury-Brown Twiss configuration [94].

71

3 Validation of Lasing in PC Nanolasers

Figure 3.1: (a) The dispersion curves of a PC waveguide and the resonances of the
L7 and L11 cavities; the shaded area is the leaky region. An SEM inset shows the
L7 cavity. (b) The -PL spectra of the L3, L7, and L11 lasers far above threshold.
Cavity modes are identified with the calculated Ey profiles.

72

3 Validation of Lasing in PC Nanolasers

3.3

Lasing Transition of a Conventional Laser

We begin by describing measurements taken on a conventional laser diode


that calibrates our technique for a low- laser. This laser has a clearly defined
threshold current of 36.25 mA as observed in the intensity versus current curve
of Fig. 3.2(a). Below threshold, the multi-quantum-well (MQW) gain spectrum
overlaps with about 100 cavity modes, which are discriminated from the lasing
mode by a monochromator for tracking g (2) (0) around the lasing threshold. As
shown in Fig. 3.2(b), g (2) (0) increases as the current is elevated to 0.986 Ith and
rapidly converges to 1 as the current is further increased to Ith . Therefore, a good
correlation can be found between the threshold current identified from Fig. 2(a)
and the narrow current region that the mode approaches a coherent state with
g (2) ( ) = 1. Spectrally, the lasing mode becomes dominant over the nonlasing
modes as shown in the insets of Fig. 3.2(b).

3.4

Lasing Transition of an L3 Laser

In sharp contrast to the MQW laser diode, for a single mode PC nanolaser
based on a small number of QDs, the achievement of transparency and lasing
threshold are expected to occur at very low pump powers. A conventional rate
equation analysis for a high-(1) laser predicts linear intensity versus pump

73

3 Validation of Lasing in PC Nanolasers

Figure 3.2: The lasing characteristics of the laser diode. (a) The output power is
plotted as a function of the operating current. (b) The g (2) (0) values as a function
of threshold-normalized current for the shaded region of (a). The insets show the
emission spectra at different currents.

74

3 Validation of Lasing in PC Nanolasers


characteristics without the standard kink [72, 90]. The actual intensity versus
pump curve of the L3 structure shows a soft-turn-on behavior, distinctly different
from that demonstrated by the laser diode. The initial sublinear increase of the
output intensity is followed by an approximately linear region for incident powers
between 2 W and 20 W, followed by the saturation region for powers greater
than 20 W as shown in Fig. 3.3. In contrast to the ambiguous intensity versus
pump characteristics, the g (2) (0) data taken as a function of pump power clearly
shows the onset of lasing at this soft-turn-on region. The pronounced bunching
behavior at low pump powers suggests the existence of correlated stimulated emission events, i.e., amplified spontaneous emission. With further increase of pump
powers, the L3 structure attains a coherent state, as indicated by the convergence
to Poissonian statistics, i.e. g (2) ( ) = 1, as shown in Fig. 3.3. The measurement
of g (2) ( ) in the L7 and L11 structures also demonstrated the onset of lasing while
the output intensity showed similar soft-turn-on behaviors. Notably, the lasing
transition region in the g (2) (0) versus pump data is broadest for the L3 structure,
sharp for the L7 structure, and sharper for the L11 structure, as shown in Fig. 3.4.

75

3 Validation of Lasing in PC Nanolasers

Figure 3.3: The lasing characteristics of the L3 laser. The g (2) (0) data and the
output intensity are plotted as a function of incident power. The g (2) ( ) spectra
taken at 0.6 W and 6.2 W are shown in the insets, where parallel lines indicate
the standard deviation noise.

76

3 Validation of Lasing in PC Nanolasers

3.5

Comparison of different lasers

The observed evolution of the onset of coherence in the family of QD PC


nanolasers can be analyzed by calculating g (2) (0) directly from the standard photon number probability distribution1
pn = p0

k=n
Y
k=1

Na
Nb + 2Tp (R1 + k)

(3.1)

[76], where p0 is the zero photon probability determined by normalization, n is the


number of photons in the cavity, and Na and Nb are the number of carriers in the
upper and the lower levels, respectively. Tp , the effective pump rate, is defined as
Tp = (0 + 0 )1 where 0 is the external pump rate. R is defined by R = 42 Tp T2 ,
and Na and Nb can be given by Na = N 0 Tp and N b = N 0 Tp , respectively. is
the cavity decay rate given by = 2/Q, where Q is the cavity quality factor
and is the cavity frequency. In this calculation, the material-dependent laser
parameters are determined experimentally. The cavity mode linewidth below
threshold is used to find . The value of 0 , i.e. the sum of all recombination rates
that do not add a photon to the cavity, can be determined from the lifetime of a QD
uncoupled from the cavity mode, which is measured to be 5 ns ( 0 1.26 GHz).
Direct optical measurement of the characteristic dephasing time T2 depends on
the exact state of the QD and its interaction with the local environment, which
1

For two alternative derivations and a more complete theoretical discussion, see App. B and

C.

77

3 Validation of Lasing in PC Nanolasers


is altered by the presence of the cavity mode [95]. Indeed, four-wave-mixing
measurements of a QD ensemble with no surrounding cavity have shown T2 values
ranging from 0.33 ps to 1 ns depending on excitation [96]. Our estimate of T2 is
obtained by measuring the coherence time using a Michelson interferometer and
PL decay time, T1 , of the photons emitted at the cavity wavelength at pump
powers well below the lasing threshold. In the bad cavity limit, where the cavity
linewidth is much greater than the QD linewidth, T2 can be determined from the
coherence time and T1 using 1/c = 1/2T1 + 1/T2 . We measured a coherence time
of 4.3 ps and a T1 of 280 ps, resulting in a T2 of 4.3 ps. Since the coherence time
matches the cavity linewidth, the cavity is most likely acting as a spectral filter
and artificially increasing the coherence time, implying we are in the good cavity
limit. Therefore, we can take 4.3 ps only as an upper bound for T2 . As a lower
bound, we note the literature value of 0.33 ps observed at high excitation [96]. To
analyze our data taken at a relatively low excitation, we choose T2 = 1 ps and
denote proper uncertainties for all variable fit parameters depending on T2 . The
number of carriers N and the effective coupling parameter between the cavity and
carriers are used as dependent fitting parameters. For the MQW laser diode,
the g (2) (0) data was analyzed with the typical values found in Ref. [76].
The theoretical results are summarized in Fig. 3.4 along with the experimental

78

3 Validation of Lasing in PC Nanolasers

Figure 3.4: (a) The g (2) (0) data of QD PC lasers are compared with the theoretical
curves obtained with [, , , N ] = [0.69 0.09, 287, 40 10, (3 1) 102 ] for L3,
[, , , N ] = [0.44 0.10, 432, 27 7, (9 3) 102 ] for L7, and [, , , N ] =
[0.19 0.08, 396, 13 2, (3 0.7) 103 ] and L11, respectively. (b) The g (2) (0)
data of the MQW laser is compared with the theoretical curves obtained with
[, , , N ] = [0.0001, 200, 0.7, 5 107 ].

79

3 Validation of Lasing in PC Nanolasers


data in terms of the factor, whose phenomenological expression is given by

0
= 1+
22

+
T2 2

1
(3.2)

[76, 97]. Here, the fitting was made to best describe the observed coherent-state
transition near and above the lasing threshold. The data deviates from theory
below threshold because of the limited timing resolution of our setup (300 ps).
The timescale of each g (2) ( ) measurement is set by the amplitude fluctuations of
the emitted light (see App. B for a complete discussion). Below threshold, this
timescale is faster than our timing resolution and the expected bunching peak
at = 0 is washed out. The result is presented as a function of a normalized
power P/Pth 2 and normalized current I/Ith . In the L3, L7 and L11 lasers shown
in Fig. 3.4(a), the systematic g (2) (0) variation near threshold can be compared
with the theoretical curves corresponding to factors of 0.69, 0.44, and 0.19,
respectively. The dash-dot lines reflect the upper and lower bounds of T2 , which
then set the uncertainty at about 0.1. In the MQW laser diode, the observed
g (2) (0) data at the vicinity of Ith agrees with the theoretical curve corresponding to
= 0.0001 as shown in Fig. 3.4(b). The lasing transition of the laser diode occurs
in a normalized pump region which is orders of magnitude narrower than those
of the QD PC nanolasers. As the nominal number of modes in the laser decrease,
2

A pump power, at which the slope of g (2) (0) versus pump power is maximized, is taken as
Pth . In our calculation, the photon number fluctuation is also maximized at Pth , in agreement
with the criteria of Ref. [92].

80

3 Validation of Lasing in PC Nanolasers


and the factors concomitantly increase, the transition to the coherent or lasing
state is broader, occurring over a wider range of pump power [92]. Similarly,
the maximum value and the slopes of g (2) (0) data decrease with the increase
of the , suggesting reduced photon number fluctuations in high- microlasers
[88]. The general trend found in these controlled active nanocavities and the
laser diode confirms ultrahigh- lasing, and allows the analysis of effective QDcavity coupling parameters. As the cavity decay rates are larger than the effective
coupling parameters, the observed lasing actions are in the weak-coupling regime.

3.6

Conclusion and Outlook

In conclusion, we have demonstrated the systematic correlation between the


coherent-state transition and the factor in a family of QD PC nanolasers. The
onset of coherence is confirmed by the convergence to Poissonian statistics, i.e.
g (2) (0) = 1, with increasing pump powers. The quantitative analysis reveals the
high factors of 0.69, 0.44, and 0.19, generally in agreement with a simple mode
counting argument. This systematic approach of mapping the onset of coherence
to a series of controlled active nanophotonic devices will be a critical measure
of lasing performance of nanocavities in general, and in particular for devices
reaching into the strong-coupling regime, where an emitter is coherently bound to

81

3 Validation of Lasing in PC Nanolasers


a single mode [98].

82

Chapter 4
Frequency control of photonic
crystal membrane resonators
This chapter is based on Applied Physics Letters 88, 043116 (2006). Copyright
2006 by the American Institute of Physics.

4.1

Introduction

Photonic crystal membrane microcavities (PCMs) are promising candidates


for applications ranging from quantum and classical communication [99], to microlasers [62, 85, 87] and sensing devices [100, 101]. Due to their ultrasmall mode
volumes [57] and large surface to volume ratio, the PCMs resonant frequency is

83

4 Frequency control of photonic crystal membrane resonators


highly sensitive to its environment. While this sensitivity may be exploited for
novel sensing applications, it complicates solid-state cavity quantum electrodynamic (QED) experiments that depend on a precise resonance condition between
a cavity mode and an embedded single quantum dot (QD) [15, 16, 18, 102], single atom [103] or single impurity [104]. This chapter describes a slow redshift of
the PCM mode emission frequency that can occur at low operation temperatures.
We ascribe this shift to molecular condensation on the PCM surface. We further
describe methods used to fully curtail the drift, and in addition, we report on the
first demonstration of a controlled redshift of the PCM mode by the adsorption
of a self-assembled monolayer (SAM) of polypeptide molecules.

4.2

Drift Tuning of a Photonic Crystal Cavity


Mode

We are particularly interested in PCM devices operated at low temperatures


in such a way that embedded QDs display a discrete energy spectrum. As a model
system a square lattice PCM geometry with one missing air hole (S1) has been
chosen, which is known to confine the fundamental mode in the proximity of the
air-semiconductor interface [68]. A single layer of self-assembled InAs QDs was
embedded in the 180-nm-thick GaAs membrane and emits around 950 nm [66]

84

4 Frequency control of photonic crystal membrane resonators


under nonresonant laser excitation at 780 nm. Figure 4.1(a) shows a spectrum of
the fundamental cavity mode taken at pump powers of 15 W, which has been
recorded with a microphotoluminescence (micro-PL) setup [105]. At these excitation conditions the cavity mode is clearly visible at 1.3293 eV with a quality
factor (Q factor) of 1900. The 2-m-diam laser excitation spot has to be positioned
with an accuracy of 0.5 m with respect to the cavity defect region [Fig. 4.1(c)],
demonstrating the strongly localized character of the mode. Individual QD transitions are visible under low pump power excitation of 50 nW, which have been
identified by their pronounced antibunching signature (not shown). Another set
of two spectra was taken 200 min later as shown in Fig. 4.1(b). While the single
QD emission energy at 1.317802105 eV did not change in the entire observation period, the cavity mode has now redshifted by 5.8 meV (4 nm) and has a
slightly lower Q factor of 1700. The energetically stable QD emission indicates
that no temperature drift or strain-induced modification of the electronic states
[106] occurred over time.
The cavity mode energy shift as a function of time is shown in Fig. 4.2(a). The
observed redshift slows down and saturates after a few hours. Measurements on
different samples show that the drift is mostly independent of the actual r/a ratio
of the S1 cavities. It is known that chemical wet etching of the PCM structures in
HF and selective removal of a self-formed native oxide will result in a systematic

85

4 Frequency control of photonic crystal membrane resonators

Figure 4.1: Low temperature (4.2 K) micro-PL spectra of a PCM shortly after
cooldown (a) and 200 min later (b). High pump power spectra (15 W, blue) show
the cavity mode and low pump power spectra (0.05 W, black) show a single QD
emitting at 1.318 eV. (c) Scanning electron micrograph of a PCM with a lattice
constant a of 290 nm and hole radius r of 110 nm around the defect region.

86

4 Frequency control of photonic crystal membrane resonators

Figure 4.2: (a) Drift of the cavity mode energy with time. Cycling the temperature
from 4 to 410 K and back to 4 K resets the mode back to 1.3295 eV and the drift
starts again as shown in Fig. 4.2(b). Data are taken at 4.2 K and 15 W pump
power.
blueshift of the cavity mode [68]. Therefore, we believe that the measured redshift
can be ascribed to material being added (adsorbed) onto the same surface through
cryo-gettering 1 . In confirmation of this hypothesis, we have found that the redshift of the cavity mode can be fully recovered by cycling the sample temperature
from 4 to 410 K and back to 4 K [Fig. 4.2(b)], demonstrating that the thin film
can be fully removed.
A more detailed study reveals that the temperature dependence of the mode
1

A similar drift of the PCM mode energy has been observed in four different optical cryostats
(Oxford), although the vacuum was better than 1106 mTorr.

87

4 Frequency control of photonic crystal membrane resonators


shift is varying as shown in Fig. 4.3(a). All temperature data have been recorded
5 h after initial cooldown to 4 K in the regime where the mode drift over time
is saturated. With increasing temperature the mode energy is nearly constant
between 4 and 30 K, blueshifts by 1 nm between 30 and 50 K and is followed by
a redshift of 2 nm up to 100 K (solid dots). In contrast, the single QD emission
energy [Fig. 4.3(b), crosses] follows the expected temperature dependence of the
semiconductor band gap [Fig. 4.3(b), red line], according to the BoseEinstein
model [107] with parameters as given in the inset. This demonstrates that the
actual temperature readout of the Si thermometer inside the cryostat is close to
the real sample temperature within the PCM defect region, where the single QD
is located. For temperatures above 50 K the slope of the mode energy follows
the expected linear redshift due to the temperature dependence of the effective
refractive index n(T ), as has been calculated by finite-difference time-domain
simulations (dashed line) assuming that n(T ) of the GaAs membrane changes with
temperature according to Ref. [108]. Thus, the observed anomalous blueshift at
temperatures around 40 K must be caused by an additional effect and is attributed
to a partial desorption and/or reconfiguration of the deposited film.
By way of comparison, we studied the behavior of the whispering-gallery mode
(WGM) of microdisk (MD) structures with 5 m diameter, which have been defined by optical lithography and transferred into the GaAs by a two-step wet etch

88

4 Frequency control of photonic crystal membrane resonators

Figure 4.3: (a) Temperature dependence of the mode emission energy for a PCM
(solid dots) and a MD (open dots). Dashed line: finite-difference time-domain
calculation of the mode shift with temperature for a PCM. (b) Temperature dependence of the single QD transition energy (crosses). The red line is a fit according to the BoseEinstein model. Mode energies are normalized to the T=4 K
values.

89

4 Frequency control of photonic crystal membrane resonators

Figure 4.4: Comparison of cavity mode redshift over time for a reference PCM
(blue dots), a SAM-coated PCM (red triangles), a microdisk (open green dots) and
a PCM capped with a thin glass slide (black dots). Mode energies are normalized
to the t = 0 min values. Data are taken at 4.2 K and 15 W pump power.
process based on HBr [64]. These devices show a largely reduced blueshift with
increasing temperature [Fig. 4.3(a) open dots] as well as a largely reduced frequency shift over time (Fig. 4.4, open green dots) compared to PCM reference
devices (Fig. 4.4, blue dots). The WGMs of the MD structures have some evanescent coupling in both lateral and vertical direction, but the PCM mode penetrates
much further into the air-hole region [68] and is therefore more susceptible to the
environment.

90

4 Frequency control of photonic crystal membrane resonators

4.3

Mode Tuning by monolayer deposition

On the one hand the pronounced sensitivity of the PCM mode to the actual
environmental conditions is promising for chemical sensing applications. On the
other hand, this sensitivity might significantly complicate the analysis of cavityQED experiments utilizing temperature tuning in order to establish a resonance
condition between the cavity mode and the emission energy of an embedded single
quantum dot. In order to fully stop the mode energy drift over time the PCM
devices have been capped with a thin glass slide2 . As a result, the energy drift is
now completely absent (Fig. 4.4 black dots). While this approach is satisfactory
for cavity-QED experiments at low temperatures, it would be highly desirable to
directly manipulate the PCM surface, allowing for selective sensing of chemical
species and to interface with functional molecules and/or colloidal QDs. To this
end we linked polypeptide SAMs to the GaAs surface of the PCM. The polypeptide
molecule is composed of eight alternating (Ala- Aib) sequences with a Glutamic
acid (Glu) attached to the C terminal of the peptide and is stable in -helix form
[109]. The peptide chemically binds to the GaAs surface by a carboxyl group
located in the Glu amino acid. It was adsorbed to the GaAs surface by first
removing the surface oxide in dilute HF, and then immediately immersing it in 1
2

The barely visible film gettered on the glass slide dissolves completely in acetone but not in
de-ionized water, indicating that the material is hydrocarbon related.

91

4 Frequency control of photonic crystal membrane resonators


mM solution of the molecule in absolute ethanol.
The thickness of the monolayer was measured by ellipsometry to be 2.20.2
nm, in agreement with the calculated length of the polypeptide of 2.6 nm. This
indicates that the molecules have formed a monolayer with the long molecule
axis almost perpendicular to the surface. Samples with an attached SAM and
initially capped with a glass slide show a redshifted PCM mode by about 35 nm
compared to reference samples without SAMs. This indicates that the SAM is
indeed attached to the GaAs surface and highlights the pronounced sensitivity
of the S1 cavity modes ability to sense layers only 2 nm thick with a frequency
response about 6-10 times larger than the full width at half maximum of the cavity
mode (0.5 nm). This sensitivity can be further increased by use of S1 cavities with
demonstrated Q factors as high as 10,000 [68].
Finally, we removed the glass slide from the SAM covered sample and found
a largely reduced magnitude of the cavity mode redshift over time by up to a
factor of 3 (Fig. 4.4, red triangles) compared to untreated PCM reference devices
(Fig. 4.4, blue dots). This demonstrates a partial surface passivation once a SAM
is attached. It is furthermore expected that the use of molecules with longer chains
and thus larger average film thickness would give rise to a further reduction of the
mode shift over time.

92

4 Frequency control of photonic crystal membrane resonators

4.4

Summary and Conclusion

In summary, our experiments show how one may control the environment of
a PCM to obtain either stable, stepped or continuous tuning operation, each of
which will be of interest in a variety of nanophotonic applications. We demonstrated a new method to attach self-assembled monolayers to GaAs photonic crystal membrane cavities opening novel possibilities for biofunctionalized photonic
devices.

93

Chapter 5
Optical Positioning of a Single
Quantum Dot

5.1

Introduction

Photonic crystal defect cavities are very interesting for cavity QED applications because of their ultrasmall mode volume. This can be understood by considering the strength of the coupling between the emitter and the cavity mode as
discussed in Sec. 1.1.1 of Chap. 1. This coupling g can be expressed in terms of
the cavity mode volume as
s
g=

e2 f
.
4mo n2ef f o Vef f

94

(5.1)

5 Optical Positioning of a Single Quantum Dot


It is clear that confining the mode to smaller volumes increases the coupling g to
the optical mode. Unfortunately, this small mode volume also decreases the yield
of cavities with a quantum dot spectrally near the mode and spatially near a field
anti-node. This technical difficulty arises because of the non-identical nature of
QDs and their random growth nucleation (see Sec. 1.2.1). Simply growing a large
density of QDs is an option, but then it is difficult to extract the cavity QED effects
due to one QD and not many. Several novel techniques have been developed to deal
with these two problems for low QD densities. As we have seen in Chap. 4, most
techniques for achieving spectral resonance in photonic crystal cavities rely on
modifying the cavitys electromagnetic environment through deposition [110, 111],
or removal of material [68]. Later, in Chaps. 7, 9, and 10, different forms of
electrical tuning mechanisms in micropillars are discussed.
Achieving spatial control of QDs has also been seen a significant amount of
research. Current techniques include growing QDs in site-specific locations [112]
or finding a way to accurately measure the position of a QD to roughly 10 nm
[80, 98]. The first type of technique has shown limited results, mostly owing to the
fact that the QDs grown in such a way are of poorer optical quality than standard,
self-assembled QDs [113]. Thus, the focus of this chapter will be on the accurate
measurement of a QDs spatial position. One technique of this type that has been
developed uses a slight QD growth modification to propagate strain due to the

95

5 Optical Positioning of a Single Quantum Dot


QD to the surface where it can be detected with atomic-force microscopy (AFM).
While implementing this technique is non-trivial, researchers have demonstrated
both weak coupling [80] and strong coupling [98]. Alternatively, the method we
have developed is an all optical measurement which relies on the fact that while a
sub-diffraction limit-sized object (like a quantum dot) cannot be resolved optically,
its spatial position can be determined with an uncertainty much smaller than the
diffraction limit by obtaining sufficient measurement statistics. This concept has
been used very successfully in the field of biophysics to measure the movement of
motor proteins such as kinesin and myosin [114116].

5.2

Experimental Setup

The experimental apparatus we developed for this measurement is depicted in


Fig. 5.1. A 780 nm CW diode laser is coupled to a single-mode fiber. This fiber
is then coupled to a 50:50 fiber beamsplitter with four ports. One of the ports
(with half of the total 780 intensity) is attached to a fiber in/out-coupler which
collimates the free-space beam. This beam is aligned to pass through the center
of an extra-long working distance objective with N.A. = 0.42. This objective
has been precisely aligned so that its piezo-controlled movement is parallel to the
surface of the sample. Using the piezo stage, the objective is swept in x and

96

5 Optical Positioning of a Single Quantum Dot


y across the sample. Simultaneously, the photoluminescence from the sample is
collected through the same objective and directed back to the fiber in/out-coupler.
Half of the photoluminescence is lost through the single mode fiber beamsplitter
(BS) to the 780 input port, but the other half is then directed to another fiber
beamsplitter. This multi-mode beamsplitter splits the single mode into two modes
with an intensity ratio of about 80:20. The brighter of these two modes is then
directed to a rotatable 1 nm bandpass filter in combination with a 900 nm long
pass filter which passes the QD emission and blocks everything else. The weaker
mode is directed to a rotatable 20 nm bandpass filter at 820 nm in combination
with a 800 nm long pass filter. This passes emission from the GaAs band gap.
The long pass filters are used to remove 780 nm light that couples weakly back
into the measurement mode. Each mode is then detected at an APD. In this way,
the excitation and collection paths are automatically simultaneously aligned with
respect to one another.

5.3

Sample Design and Experimental Procedure

The samples are fabricated by Susanna Thon and are comprised of a single
layer of InAs QDs 70 nm below the surface with an Al0.3 Ga0.7 As sacrificial layer
for the creation of the photonic crystal membrane. On top of the sample, an

97

5 Optical Positioning of a Single Quantum Dot

Figure 5.1: The experimental setup for optical positioning. The magenta (black)
line denotes the path of the 780 nm laser and photoluminescence in free space
(fiber). The blue line denotes the path of the white light source for imaging.
Each spectral filter box contains a long-pass filter and a bandpass filter, which is
attached to a rotatable stage so that the center wavelength can be adjusted.

98

5 Optical Positioning of a Single Quantum Dot

Figure 5.2: An SEM image of one box on the optical positioning sample. The
quantum dots position is determined relative to the center of the horizontal and
vertical gold alignment marks. Image courtesy of Susanna Thon.
array of gold boxes are formed, as shown in Fig. 5.2, each with an alpha-numeric
label. Each box is 2222 m large, with an alignment mark width of 2 m.
The positions of the QDs are determined as follows. First, before the deposition
of the alignment boxes, micro-photoluminescence spectroscopy is performed on
the sample to determine a suitable QD density region. The density of QDs in
an optimum region is roughly 1 to 2 QDs per box in the wavelength range of
interest. This ensures that the photonic crystal will interact with only one QD.
After this density region is found, alignment marks are deposited, and the sample
is cooled once again to 4 K. Now, the boxes are searched for good candidates

99

5 Optical Positioning of a Single Quantum Dot


for positioning. A good candidate box is one that contains a bright QD near
the center, so that a photonic crystal with dimensions roughly 88 m can be
easily fabricated without interference from the walls of the box. Once enough
candidates are found, the positions of each QD are determined by scanning the
objective over the size of the box1 using a computer-controlled piezo stage while
photons are collected simultaneously from the GaAs band gap at 820 nm and from
the candidate QD. Scans are performed alternately in the y direction (keeping the
x coordinate constant) and then the x direction, with the x(y) position of a y(x)
scan determined by the x(y) coordinate that corresponded with the maximum QD
intensity from the previous scan. In this way, even though the cryostat position
drifts with time, the scanning procedure follows the position of the QD and ensures
that an appreciable intensity is measured from the QD over the entire length of
the procedure. While the location of the QD is determined by a maximum of
intensity, the location of the walls are determined by a decrease in the intensity of
the 820 nm emission. This is because the walls do no transmit any emission from
the sample. The scanning procedure continues until sufficient statistical errors are
obtained. In practice, this is achieved after 30 scans in each direction which takes
roughly 30 minutes.
1

In practice the size of each scan is 40 m.

100

5 Optical Positioning of a Single Quantum Dot

5.4

Data Analysis

Each scan outputs a data file containing the x and y positions of the piezo stage
as well as the measured counts from QD and from the band gap as a function of
time. Then, each scan is fit to determine the absolute positions of the QD and
walls in that scan. The peak in the QD data is fit to a Gaussian and the dips in
the band gap data are fit to a Gaussian convolved with a step function as shown
in Fig. 5.3(a)-(b). Both of these functions fit the data very well and the results are
manipulated to determine the relative distance between the QD and the nearest
wall in x and y. This list of x and y relative distances are then averaged to give
the final coordinates of the QD relative to the walls, and the standard deviation
is taken to determine the error. This procedure yields a Gaussian distribution of
x and y coordinates as shown in Fig. 5.3(c), so the error in the mean of these

distributions (which is the precision of our measurement) is / N , where is


the standard deviation and N is the number of scans. In practice, the statistical
error in the mean of 30 scans is less than 5 nm.

5.5

Photonic Crystal Cavity Fabrication

Once the positions of several QDs are determined, photonic crystals are fabricated in the clean room. Susanna Thon fabricates the photonic crystals and the

101

5 Optical Positioning of a Single Quantum Dot

HcL

Counts

HbL

Counts

HaL
1200
1000
800
600
400
200
0

38 39 40 41 42 43
x HmL

15 000
10 000
5000
0

26

28 30
y HmL

32

Number of Scans

20
15
10
5
0
5.10

5.15
5.20
Relative x HmL

5.25

5.30

Figure 5.3: (a) Quantum dot APD counts versus position with Gaussian fit for
an x scan. (b) GaAs band gap APD counts versus position with fit for a y scan.
The fit function is a Gaussian convolved with a step function. (c) A histogram of
the extracted relative x position of a quantum dot for about 80 scans. The error
in the mean is about 4 nm.

102

5 Optical Positioning of a Single Quantum Dot


exact techniques used will be found in full detail in her thesis, so a short synopsis
is as follows. An e-beam writer is used to expose the resist and imprint the array
of circles which compose the photonic crystal cavity. To accurately position the
photonic crystal inside the gold box, the e-beam writer is used in the direct-write
mode to detect the gold markers. The e-beam writer then exposes a photonic
crystal cavity with lattice parameters such that the mode will be spectrally near
the positioned QD2 . Next, the resist is developed and the photonic crystal array
is transfered into GaAs by reactive ion etching. Finally, the sample is undercut by
using HF to selectively etch the sacrificial Al0.7 Ga0.3 As, forming a membrane. An
SEM image of a completed photonic crystal cavity membrane positioned about a
QD is shown in Fig. 5.4.

5.6

Experimental Results

To verify that the quantum dot is positioned near the cavity anti-node, the
coupling between the quantum dot and mode is investigated. Using the spectral
tuning mechanism described in Chap. 4, the resonance of the cavity mode drifts to
lower energy as a function of time. In this way, a cavity mode that is initially blueshifted with respect to the QD transition can reach spectral resonance. Because
2

This wavelength is usually slightly to the red of the QD so that the mode can be blue-shifted
using digital etching [68] and/or the QD can be red-shifted with temperature.

103

5 Optical Positioning of a Single Quantum Dot

Figure 5.4: An SEM image of a positioned photonic crystal cavity. The marks seen
perpendicular to the gold alignment marks are from the e-beam writers detection
of them for positioning. Image courtesy of Susanna Thon.

104

5 Optical Positioning of a Single Quantum Dot


this spectral drift is very slow (roughly 10 eVmin1 at the most), high quality
measurements of the coupling can be made by simply recording many subsequent
spectra. Experimentally, CCD integration times of 1 sec. are more than adequate.
Fig. 5.5 shows a set of these spectra. As the mode approaches the QD transition
near 1500 s, both emission lines experience repulsion which leads to an anticrossing. This is a definitive measurement of strong coupling between the QD and
the cavity mode. Thus, the quantum dot must be positioned within 50 nm the
anti-node of the cavity field.
The anti-crossing is more clearly seen in Fig. 5.6, which is a plot of the spectral
positions of both resonances in the spectra as a function of time. Also shown
are the linewidths of each resonance as the dotted lines surrounding the center
data points. The upper resonance is initially broad and is clearly identified as
cavity mode-like whereas the lower resonance is very narrow and is clearly QDlike. As the cavity mode approaches the QD, the upper branch narrows while
the lower branch broadens. This occurs because the resonances of the mode and
QD become indistinguishable in strong coupling. At much later times, far from
spectral resonance, the lower branch takes on the character of the mode and the
upper branch has QD-like character. Thus, the resonances passed through one
another without ever crossing, which is the definition of a spectral anti-crossing.

105

5 Optical Positioning of a Single Quantum Dot

Figure 5.5: A density plot of spectra as a function of time showing a PC cavity mode reaching spectral resonance with a QD transition. Near resonance, a
clear repulsion leads to an anti-crossing, a definitive signature of strong coupling
between the mode and the QD.

106

5 Optical Positioning of a Single Quantum Dot

Spectral Position HeVL

1.3580
1.3578
1.3576
1.3574
1.3572
1.3570
1.3568
0

2000

4000
6000
Time HsL

8000

Figure 5.6: A plot of the measured spectral positions of both resonances as a


function of time. Also, the linewidth of each resonance is shown by the dotted
lines near each resonance. The spectral position and linewidth are determined by
fitting each spectra with two Lorentzians and a constant background.

107

5 Optical Positioning of a Single Quantum Dot

5.7

Discussion and Conclusion

The strength of the coupling between the QD and the mode g can be determined from the vacuum Rabi splitting found in Fig. 5.6. At spectral resonance,
this splitting can be expressed as
r
=

g2

1
( QD )2 ,
16

(5.2)

where is the linewidth of the cavity mode and QD is the linewidth of the QD
transition and are determined from the fit of the first spectrum. Experimentally,
2 is taken to be the smallest detuning between the two resonances and g is
determined using Eq. 5.2. g is found to be 702 eV, about 70% of the maximum
value calculated from the oscillator strength and the mode volume. If the spatial
variation of the vacuum electric field is approximated as a gaussian with a width
of = 40 nm, the reduction in g can be attributed to a mis-positioning of 30
nm. This is more than the statistical error in the measured position of the QD,
but is reasonable considering systematic measurement errors of the QD position
and errors in the placement of the photonic crystal. Furthermore, this error is
similar to the error using AFM-type QD positioning found in Ref. [98].
We used this technique to position several other QDs and the yield of devices
that demonstrated strong coupling was 75%, demonstrating the remarkable repeatability and robustness of this positioning technique. The benefits of this
108

5 Optical Positioning of a Single Quantum Dot


simpler positioning technique (compared to AFM positioning) are enormous, and
the impact will be far-reaching in QD photonic crystal research. Future implementations could enable coupling different, pre-selected QDs in different cavities
via a waveguide for quantum information tasks as well as a host of other schemes.
Optical positioning of a single QD removes a significant obstacle in the use of
these ultrasmall cavities.

109

Chapter 6
Oxide-Apertured Micropillars
This chapter is based on Applied Physics Letters 87, 031105 (2005). Copyright
2005 by the American Institute of Physics.

6.1

Introduction

Optical microcavities combined with active emitters provide a great opportunity to study the light-matter interaction at a fundamental level. To produce
a high quality microcavity it is necessary to confine light to precise resonance
frequencies with little or no optical loss [117]. The measure of this optical confinement is referred to as the cavity quality factor (Q). In order to limit the
number of optical modes present in a cavity it is also important to reduce the

110

6 Oxide-Apertured Micropillars
effective optical mode volume (Vef f ). Figures of merit for optical microcavity applications are proportional to the ratio of these two values Q/Vef f [57]. Potential
applications include solid state cavity quantum electrodynamics (CQED) experiments, modification of single emitter lifetimes, and single photon emitters and
detectors for quantum cryptography [57, 117].
Several solid-state microcavity architectures including microdisks [118, 119],
photonic crystals [16, 21, 80], and micropillars [15, 120] have shown CQED effects
in III-V semiconductors using self-assembled quantum dots (QDs) as active emitters. Among these architectures, micropillars couple light normal to the semiconductor in a single lobed Gaussian pattern that is easily fiber coupled [121].
This high photon collection efficiency makes micropillars a better alternative for
device applications. However, micropillars exhibit higher Vef f [ 5(/n)3 ] and lower
Qs ( 2000-10000) when compared with photonic crystals [15, 117]. Vef f can be
reduced by decreasing pillar diameter, but scattering losses due to sidewall roughness have been shown to limit achievable Q values [117, 122].

6.2

Micropillar Cavity Design

Here we present an alternative approach, using oxide apertured micropillars in


order to reduce Vef f while maintaining high Q values. Oxidized micropillars have

111

6 Oxide-Apertured Micropillars
been used for vertical cavity laser (VCL) applications to produce low threshold
laser devices that are fabricated into inexpensive arrays for optical data networks
[123]. By confining the optical mode with a laterally oxidized aperture layer these
structures simultaneously provide optical mode and electrical current confinement
while eliminating the scattering loss due to sidewall roughness inherent to etched
pillar structures. This method has been applied to the field of QD-microcavity
coupling with limited success due to very low Qs (< 1000) as well as high Vef f
[ 35(/n)3 ] when compared with conventional micropillars [124, 125]. If these values are improved, the advantages in ease of fabrication will make oxide apertured
micropillars very attractive for coupled QD-microcavity device applications.
Micropillar samples investigated in this study were grown by molecular beam
epitaxy on a semi-insulating GaAs (100) substrate with a 0.1 m buffer layer.
There are four independent sections in the structure: the bottom mirror, the
active region, the aperture region, and the top mirror as shown schematically
in Fig. 6.1(a). Mirrors consist of alternating one quarter optical thickness distributed Bragg reflector (DBR) layers of GaAs and Al0.9 Ga0.1 As. 32 pairs of
Al0.9 Ga0.1 As/GaAs layers with thicknesses of 79.8/68.4 nm respectively form the
bottom DBR mirror, while the top DBR mirror is made of 23 pairs. The active
region is one optical wavelength in thickness, with two 135.4 nm layers of GaAs
embedding a centered InGaAs/GaAs QD layer. QDs self-assemble during epitaxy

112

6 Oxide-Apertured Micropillars
operating in the Stranski-Krastanow growth mode. InGaAs islands are partially
covered with GaAs and annealed before completely capped with GaAs. This procedure blue shifts the QDs emission wavelengths [26] toward the spectral region
where Si-based detectors are more efficient. The thickness of the aperture region is
three quarters optical wavelength and consists of a pure AlAs layer sandwiched by
Al0.89 Ga0.11 As and Al0.75 Ga0.25 As in order to produce the desired oxide aperture
qualities. It is designed to give a change in effective index, nef f = 0.08, between
the fully oxidized and unoxidized regions of the micropillar in addition to a linear
oxide taper with a length of 1.1 m after an approximate 10 m oxidation. A
scanning electron microscopy (SEM) image of a fabricated oxide apertured microcavity is shown in Fig. 6.1(b), while Fig. 6.1(c) shows a cross-sectional SEM
image of an oxidized mesa calibration sample with mirror, active, and aperture
regions corresponding to Fig. 6.1(a). Samples are fabricated by optical lithography and reactive ion etch (RIE) in Cl2 plasma which penetrates approximately 5
mirror periods into the bottom DBR. A hard SrF2 mask is used during the Cl2
plasma etch. Micropillars are fabricated in large arrays with diameters varying
from 21-25 m. The wet lateral oxidation is performed at 430C in order to oxidize
the aperture region by converting AlAs into Alx Oy .
Micro-photoluminescence (-PL) measurements are performed using a He-flow
cryostat (4-300K). QDs are excited non-resonantly by a continuous wave 780 nm

113

6 Oxide-Apertured Micropillars

Figure 6.1: (a) Schematic and layer structure for oxide apertured micropillars.
(b) SEM image of a fully processed 21 m micropillar. (c) SEM cross section
image of an oxidized mesa showing actual layer structure depicted in (a). (d) The
theoretical fundamental mode profile for the 21 m micropillar modelled in this
study with a Gaussian mode radius of 0.78 m.

114

6 Oxide-Apertured Micropillars
laser diode focused to a spot size of 2.5 m using a microscope objective with
numerical aperture of 0.55. -PL emission is collected through the same objective and recorded with a 1.25 m spectrometer equipped with a charge-coupled
device with 30 eV spectral resolution at 900 nm. A scanning Fabry-Perot cavity
along with a single photon counting avalanche photodiode detector and an integrated counting unit are used to experimentally determine linewidths beyond the
resolution of the spectrometer.

6.3

Model of Micropillar Cavity Modes

A two-dimensional model of the cavity has been developed based on the experimentally determined values for the oxide aperture taper length and core width
from SEM images (Fig. 6.1) along with the one-dimensional reflectivity spectrum
of the unprocessed sample. This produces a two-dimensional index profile determined by the effective index (nef f ) in the growth direction for the unoxidized
and oxidized layer stack. The nef f between the unoxidized and oxidized regions is evaluated by replacing the AlAs and Al0.89 Ga0.11 As with Alx Oy (n = 1.5).
It has been demonstrated [126] that the linear oxide taper shown in Fig. 6.1(c)
corresponds to a parabolic index grade over the length of the taper.
We used this model to solve for the eigenmodes of the two-dimensional scalar

115

6 Oxide-Apertured Micropillars
wave equation using a finite difference technique with a non-uniform mesh [74].
The solution for the fundamental mode of a 21 m pillar is shown in Fig. 6.1(d) and
has a Gaussian mode radius of approximately 0.78 m. Scattering and radiation
losses are determined by propagating a scalar field around the unfolded cavity
until the field no longer changes shape. This procedure is analogous to the classic
work of Fox and Li [127].
Mirror, scattering, and radiation losses determine the empty or cold cavity
linewidth of the apertured micropillar. Assuming that undoped AlGaAs regions
have no internal optical loss at 4K, the only optical loss mechanisms in the cavity
are due to mirror loss (m ), radiation loss (rad ), and aperture scattering losses
(scat ) [74]. Furthermore, rad and scat are very small for the fundamental mode,
1.7 103 cm1 and 1.7 cm1 . This leaves photon escape through the top DBR
mirror as the dominating loss mechanism in the cavity, calculated as 13.9 cm1 .
Cold cavity (Qcold ) values are determined by these cavity losses according to [74]

1
= g gth = g (scat + rad + m ).
=
Qcold
p

(6.1)

Here p is cavity lifetime, g is group velocity, is the frequency, is the confinement factor, and gth is the threshold material gain. The estimated Qcold for a 21
m pillar is 14500.

116

6 Oxide-Apertured Micropillars

Figure 6.2: (a) Normalized optical mode spectra for 2125 m micropillars measured by -PL at 4 K are shown. Fundamental, first-, and second-order modes
are labelled 0, 1, and 2, respectively. (b) Mode position is shown as a function of
pillar diameter for fundamental, first-, and second-order modes.

6.4

Experimental Results

Experimental mode spectra for 21-25 m diameter pillars are shown in Figure
6.2(a) with mode orders 0, 1, and 2 labelled. The lifting of higher order mode
degeneracy is due to asymmetry in the fabrication process. -PL data shows
cavity modes with lower fundamental energies as well as increasing mode spacing
as pillar diameter decreases. This effect is shown in Figure 6.2(b) for varying
pillar diameters. In addition, intensity decreases are observed for higher order
modes due to increased scattering losses, an intentional effect produced by the
oxide aperture and the larger effective radii of multi-lobed higher order modes.
A Fabry-Perot scanning cavity is used to determine experimental Q values.
A reference laser is used to calibrate the maximum measurable Q for the Fabry-

117

6 Oxide-Apertured Micropillars
Perot cavity as 600000, shown in Fig. 6.3(a). The fundamental cavity mode is then
directed through a 1 nm bandpass filter into the detector for sample measurement.
Experimental quality factor (Qexp ) is determined to be approximately 48000 for a
21 m pillar as shown in Fig. 6.3(b). This value is larger than the theoretical cold
cavity value due to linewidth narrowing caused by modal gain from the QD active
region. The experimental gain (gexp ) is determined by the relationship between
Qcold and Qexp values according to [74]

=
g gexp .
Qexp
Qcold

(6.2)

Q values show little change with increasing pump power and no laser threshold
behavior is observed up to 50 W pump power. This indicates that material gain
in the QD layer saturates before the threshold condition is achieved. Estimated
values show that the threshold modal gain in the QD active region saturates at
approximately 10.44 cm1 , corresponding to approximately 2864 cm1 material
gain from the active region. This is a reasonable value for the single QD layer in
this cavity [128]. At low pump powers, Q values are expected to increase as the
optically pumped QD region provides gain and causes linewidth narrowing. For
the small lateral area and narrow spectral linewidths in these devices, the modal
gain provided by the QD single-exciton emission saturates at pump powers below

118

6 Oxide-Apertured Micropillars

Figure 6.3: (a) Transmission through a FabryPerot cavity as a function of cavity


length for a reference laser. Using the given free spectral range of the measurement
cavity, the reference laser has Q > 600000 giving a maximum measurable Q. (b)
The FabryPerot measurement of the fundamental mode linewidth of a 21 m
micropillar cavity. This measurement was done at 50 W pump power at 4 K.

119

6 Oxide-Apertured Micropillars
1 W [80]. In this regime, a slight linewidth narrowing from 0.06 to the observed
0.02 nm is expected to occur. However, it is difficult to resolve this linewidth
narrowing with a scanning Fabry-Perot cavity at these low pump powers.

6.5

Discussion and Outlook

The Purcell factor determines the figure of merit for a coupled QD microcavity
in the weak coupling regime. Using the theoretical Vef f (/nef f )3 of 51 and
experimental Qexp values, it is possible to estimate the maximum achievable Fp
according to [17]
Fp =

3
4 2

Vef f

3 .

(6.3)

nef f

Experimentally observed values are decreased due to spectral and spatial detuning.
Estimating the Purcell factor with experimental linewidths gives a maximum value
of 72.
Oxidized micropillars have a promising outlook for QD-microcavity coupling
due to high Q values. Further improvement for these devices could be achieved
by reducing Vef f . Although oxide aperture micropillars reduce two-dimensional
mode areas while maintaining high Q values, Vef f remains relatively high. This
is due to the large effective cavity length in the growth direction of this structure,
approximately 1.39 m. Improvement lies in reducing this value, potentially by

120

6 Oxide-Apertured Micropillars
replacing AlGaAs/GaAs DBR mirrors with Alx Oy /GaAs DBR mirrors having
reduced mirror penetration depths.
Experimental data show very high Q (48000) optical microcavities using an
oxide apertured micropillar architecture. Devices exhibit controlled mode positions and sizes down to core widths of approximately 0.5 m. Unlike etched air
interface micropillars, which can be difficult to fabricate with acceptable loss values at smaller diameters, apertured micropillars accomplish this in a controllable
and repeatable fashion. Cavities exhibit low loss values resulting from gain related
linewidth narrowing effects due to stimulated emission from the QD active region.
Apertured micropillars show promise for QD-microcavity coupling applications
due to these high experimental Q values producing a maximum Purcell factor of
72.

121

Chapter 7
High-Frequency
Single-Photon-Source with
Polarization Control
This chapter is based on Nature Photonics 1, 704 (2007). Copyright 2007 by
the Nature Publishing Group.

7.1

Introduction

Optoelectronic devices that provide non-classical light states on demand have


a broad range of applications in quantum information science [99], including

122

7 High-Frequency Single-Photon-Source with Polarization Control


quantum-key-distribution systems [129], quantum lithography [130] and quantum
computing [10]. Single-photon sources [131, 132] in particular have been demonstrated to outperform key distribution based on attenuated classical laser pulses
[133]. Implementations based on individual molecules [134], nitrogen vacancy centres [48] or dopant atoms [104] are rather inefficient owing to low emission rates,
rapid saturation and the lack of mature cavity technology. Promising singlephoton-source designs combine high-quality microcavities [57] with quantum dots
as active emitters [18]. So far, the highest measured single-photon rates are 200
kHz using etched micropillars [135, 136]. Here, we demonstrate a quantum-dotbased single-photon source with a measured single-photon emission rate of 4.0
MHz (31 MHz into the first lens, with an extraction efficiency of 38%) due to the
suppression of exciton dark states. Furthermore, our microcavity design provides
mechanical stability, and voltage-controlled tuning of the emitter/mode resonance
and of the polarization state.
Excitons inside semiconductor quantum dots (QDs) interact efficiently with
light, making them attractive as an optically active material for single-photon
sources (SPS) [18, 52, 131, 132, 135139] LEDs (ref. [140]), low-threshold lasers
[87] and solar cells [141]. Under optical excitation or electrical-carrier injection,
electrons and holes are created in higher energy states and subsequently relax
into localized QD states. Random carrier capture leads to both geminate and

123

7 High-Frequency Single-Photon-Source with Polarization Control


nongeminate loading of the QD, causing limitations in device efficiency [142].
When a QD is loaded with a neutral exciton (X0 ), it can be either in a brightstate configuration with angular momentum Lz = 1 or in a dark state with
Lz = 2. Although the bright-state configuration is dipole allowed with a typical
spontaneous emission (SE) lifetime of 1 ns, the dark-state configuration is dipole
forbidden and thus a long-lived state. If a QD is loaded into a dark state, a
spin-flip must occur, for example involving the Fermi sea of an adjacent back
gate [143], or it can be pumped through the biexciton (XX) to the bright state.
Thus if a QD is triggered to emit single photons (SPs) at a high repetition rate,
it occupies on average more time in dark-state configurations, which limits the
maximal achievable emission rate (see Sec. 7.7). Additionally, the capture of
single carriers leads to submicrosecond correlation effects [136], which act in effect
as dark states.
To eliminate dark-state configurations we aimed to load a QD embedded in a
microcavity with a single electron. Subsequent electronhole pair capturing owing
to optical (or electrical) excitation would create predominantly charged excitons
(X ) in a singlet configuration. In such a case, recombination is always expected
to be bright and an SPS can continuously fire SPs with rates approaching the SE
lifetime limit. Additional electrons can either be injected through electrical gates
or provided by doping QDs or their surroundings or both.

124

7 High-Frequency Single-Photon-Source with Polarization Control

7.2

Electrical Contacts

Although controlled loading of single QDs has been demonstrated [56, 143],
implementation of electrical gates into high-Q cavities has not yet been achieved.
To this end we have developed oxide-tapered high-Q microcavities based on a
rugged trench design. Figure 7.1a shows a device schematic. A layer of QDs
inside a wavelength-long cavity (-cavity) is sandwiched between two distributed
Bragg reflectors (DBR). A top electrical gate and a tunnel back gate with 25nm barrier are formed adjacent to the QDs (see Sec. 7.6 for details). These
n-doped layers provide an embedded electrical gating structure to control the QD
loading process and local current heating within the cavity. Air-etched GaAs
micropillars are typically fabricated with submicrometre diameters to confine the
mode volume [57, 135]. To avoid difficulties associated with micrometre-scale
electrical contacts, we use trenches with various geometries etched down into the
bottom DBR (Fig. 7.1b). These trenches are used to define an AlOx oxidation
front, providing high-quality factors up to 50, 000 and simultaneously optical-mode
confinement [137]. This design provides good mechanical stability, the possibility
of controlling the mode degeneracy, and it allows global contacting by countersink
etching two contacts to an entire SPS array.
Figure 7.2a shows a schematic of the band structure including top and back

125

7 High-Frequency Single-Photon-Source with Polarization Control

Figure 7.1: (a) Schematic three-dimensional view of the SPS containing a cavity region embedded between a 32-period bottom AlGaAs/GaAs DBR and
a 23-period top DBR (white/blue layers). The cavity contains a single layer of
InAs QDs (grey) and a mode-confining tapered AlOx region (dark blue). Selective countersink etching allows separate global contacting (yellow) of the n-doped
top and bottom gate layers (red). (b) Scanning electron micrograph of various
geometries of the trench design (top view). The inner lateral cavity area has a
diameter of 20 m. The remaining small bars provide lateral electrical contacts.

126

7 High-Frequency Single-Photon-Source with Polarization Control


gates (blue areas). The bottom gate is grounded and voltage is applied to the
top gate. Optical emission is probed under non-resonant laser excitation into
the GaAs. At negative bias voltages below 22 V the QD emission completely
quenches, in analogy to gated structures without a cavity [56]. For voltages above
about 5.5 V, carriers overspill and create a current breakthrough, as shown in
the current-voltage (I V ) curve in Fig. 7.2b. This creates two useful regimes
for SPS operation. For voltages below breakthrough, one can manipulate the QD
loading process as shown in the following. Above breakthrough, the electrical
current causes localized heating (on-chip heater), which can be used to fine-tune
the QD/mode resonance (Fig. 7.2c,d).
Interestingly, a strong fivefold increase in QD emission is observed for voltages
of 25 V (Fig. 7.2c). In this regime, no spectral shift of the QD exciton occurs.
This excludes both QD charging (X formation) and spectral QD/mode detuning
(Purcell effect [132]) as an explanation. Controlled QD/mode tuning can either
be achieved at zero gate voltage using the cryostat heater (method A) or at a fixed
base temperature (4 K) and increasing the gate voltage above 5 V (method B), as
shown in Fig. 7.2d. For method A, the QD transition has a 40-fold intensity enhancement if tuned into resonance (blue circles). Method B creates an additional
eightfold enhancement for all detuning conditions (red triangles). Simultaneously,
the X0 transition displays a Purcell factor of two (see Fig. 7.3b), independent of

127

7 High-Frequency Single-Photon-Source with Polarization Control

Figure 7.2: (a) Schematic energy band diagram. The tunnel barrier width between
the back gate and the QD is 25 nm. CB, conduction band; VB, valence band.
(b) Currentvoltage characteristic of the entire SPS array recorded at 4 K. (c)
Single QD intensity as a function of gate voltage for a type B (from Fig. 7.1b)
device under 780 nm excitation. (d) Comparison of SPS intensity as a function
of QD/mode detuning energy using the cryostat heater at zero gate voltage (T
tuning, blue circles) and for fixed cryostat temperature (4 K) and gate voltages
between 5 and 9 V (V tuning, red triangles). Detuning energies are taken from
Lorentzian line-shape analysis.

128

7 High-Frequency Single-Photon-Source with Polarization Control


the tuning method. Thus the intensity enhancement is independent of the actual
SE rate, suggesting a relation to the QD loading process [144]. These findings
suggest that the voltage-induced enhancement (Fig. 7.2c,d) is related to fieldenhanced carrier capture processes. Indeed, if single QDs are optically pumped
into SE saturation, the voltage-induced intensity enhancement also saturates (not
shown). However, practically, SPSs are not operated at saturation because of
degraded SP purity. Therefore, the achieved voltage-controlled enhancement is of
high practical importance as it allows higher SP rates at a comparable injection
level while minimizing background contributions that otherwise compromise the
SP purity.

7.3

Single Photon Results: X0 and X

Single-photon sources are characterized by strong photon antibunching signatures [131, 132]. Figure 7.3a shows a spectrum of a single QD loaded with X
and tuned in resonance with a degenerate cavity mode. The inset shows the corresponding second-order correlation function, g (2) ( ), displaying a zero delay time
peak area of 4%. This corresponds to a 25-fold decrease in unwanted multiphoton
emission events compared with attenuated coherent light sources. Devices typically display a 40-fold enhanced SP emission rate if sharp exciton lines are tuned

129

7 High-Frequency Single-Photon-Source with Polarization Control


into spectral resonance (Fig. 7.2d). These results were achieved for QDs with an
SE lifetime of about 1 ns, similar to QDs in bulk GaAs. They are thus spectrally,
but not spatially, on resonance. In such a case intensity enhancement is entirely
caused by enhanced geometrical collection efficiency and not related to the Purcell effect. In about a third of devices, the QDs under consideration are spatially
on resonance and display a Purcell factor of two to four if tuned spectrally into
resonance. To separate effects related to random positioning of QDs from effects
caused by charging and applied voltages, we concentrate on devices with a Purcell
factor of approximately two (Fig. 7.3b,c).
To show that X is better suited for an SPS than X0 , we first have to identify
the X0 and X transitions of individual QDs. In high-Q cavities, clear identification based on spectral signatures, that is, the typical 5-6 meV Coulomb
charging separation [56], is rather difficult because spectra are largely altered by
the underlying mode distribution and stop band. However, as has been demonstrated [143], time-resolved measurements of QDs in gated structures display a
mono-exponential decay for the X transition and a pronounced bi-exponential
decay for X0 . Figure 7.3b,c shows lifetime measurements for single QDs located
in two different devices, clearly distinguishing between biexponential (Fig. 7.3b)
and mono-exponential (Fig. 7.3c) decay for both off-resonance (red circles) and
on-resonance (blue circles) conditions with a degenerate mode, supporting our

130

7 High-Frequency Single-Photon-Source with Polarization Control

Figure 7.3: (a) Cavity mode spectrum with a neutral QD coupled to a secondorder mode at 7.3 V bias. Inset: Second-order correlation function g (2) ( ) recorded
with 10 nW pump power. Note that submicrosecond correlation effects [136] are
absent under biased conditions for both QD charge states even at lowest pump
powers. (b) Lifetime measurements for QD loaded with X0 under off-resonance
(red circles) and on-resonance (blue circles) conditions. Black data points indicate
scattered laser light corresponding to 160-ps system resolution. (c) Comparable
lifetime measurements for a QD loaded with X . Note the mono-exponential
decay in contrast to the bi-exponential decay of X0 .
131

7 High-Frequency Single-Photon-Source with Polarization Control


assignment of X0 and X .
We investigated the SPS device performance by recording both intensity and
antibunching signatures as a function of pump power (Fig. 7.4ad).With increasing
pump power, antibunching signatures bleach out owing to QD saturation and
SE coupling from other QDs at higher pump powers [87], limiting the maximal
achievable SP rates. To compare devices pumped under pulsed and continuouswave (c.w.) excitation conditions we chose a working point close to QD saturation,
where g (2) ( = 0) = 0.4, that is, when the SPS still performs 2.5 times better than
an attenuated laser source. Before comparison, the measured intensity data had
been corrected for the multiphoton background to extract the bare SP rate (see
Sec. 7.6). Two devices with degenerate modes were chosen for comparison, one
with X0 on resonance (Fig. 7.4a,b) and one with X on resonance (Fig. 7.4c,d).
Under pulsed excitation, measured SP rates are 2.7 MHz for the X0 and 4.0 MHz
for the X transition, respectively. If corrected for the 13% detection efficiency,
these devices emit into the first lens with a record high SP rate of 21 MHz for X0
and 31 MHz for X . From these data and the given repetition rate (82 MHz) one
can estimate an SP extraction efficiency of 26% for X0 and 38% for X . If a
higher purity of SP is required [145], one can, for example, operate the X devices
at 1 W pump power (Fig. 7.4c) leading to g (2) (0) = 0.05, an SP rate of 21 MHz
and = 25%.

132

7 High-Frequency Single-Photon-Source with Polarization Control

Figure 7.4: (a)-(d) QD intensity (blue squares, left y axis) and corresponding
g (2) ( = 0) values from coincidence measurements (open red triangles, right y
axis). Intensity data are corrected for background emission (open black dots; see
Sec. 7.6). Two devices are compared: one with X0 (a, b) and one with X on
resonance (c, d). For each charging state we compare pulsed laser excitation at
780 nm (a, c) and c.w. laser excitation at 630 nm (b, d). Arrows indicate SP
intensities at g (2) ( = 0) = 0.4, close to QD saturation. Note the noticeably
improved count rates for charged QDs under c.w. excitation.

133

7 High-Frequency Single-Photon-Source with Polarization Control


If one drives the two-level system even faster than with an 82 MHz repetition
rate, one would expect even higher SP rates of about 870 MHz up to the SE lifetime limit (This limit is estimated using the measured (Purcell enhanced) exciton
lifetime of 440 ps (2.3 GHz) modified by the cavity-mode collection efficiency of
38%.) Pumping the single QD under c.w. excitation thus probes for the ultimate
limit. We performed the same SPS characterization under c.w. excitation at 632
nm (Fig. 7.4b,d). Surprisingly, for QDs loaded with X0 , maximal achievable SP
rates saturate at about 2.2 MHz, which is comparable to the pulsed excitation
experiment (Fig. 7.4b). This demonstrates that the emission rate of X0 is severely
limited by dark states. Most strikingly, c.w. excitation of an SPS with X on
resonance yields an SP rate of 10.1 MHz (Fig. 7.4c), corresponding to 78 MHz if
corrected for detection efficiency, which is a threefold improvement over the X0
case (see Sec. 7.6). In contrast to the case of X0 , the SP rate for X saturates
slowly and one can estimate 116 MHz at the highest pump powers near 150 W.
At lower pump powers the X /X0 ratio is close to unity. These findings are in
quantitative agreement with a rate-equation model (see Sec. 7.7). At low pump
rates there is enough time for a spin-flip to occur between excitation events, and
X0 dark states can become bright again, leading to a ratio of unity. At higher
pump rates the XX creates a path back into the bright state for X0 . In a cavity the XX is inhibited if X 0 is on resonance, which limits the X0 recovery time

134

7 High-Frequency Single-Photon-Source with Polarization Control


out of dark states compared with the Purcell enhanced recombination rate of the
X transition, which is always bright. Although these c.w. experiments clearly
demonstrate that a threefold improvement can be achieved by using X , the nature of c.w. excitation does not in principle allow for on-demand operation. Thus
the highest achieved count rate for on-demand operation of our SPS devices is 31
MHz ( = 38%), constituting a fivefold improvement compared with other SPSs
(refs [18, 48, 52, 57, 104, 131139]; see Sec. 7.6), which makes them very attractive
for quantum light applications.

7.4

Polarization Control

In addition to high SP rates, a practical device for quantum communication


can provide on-chip selection of the emitted polarization state. Non-degenerate
cavity modes allow the generation of an SPS emitting either horizontal (H) or
vertical (V) polarized SPs, as shown earlier [19, 146]. To actively switch between
polarization states on the chip, one must tune the QD emission either to the energy
of the H- or V-polarized mode. To this end, we explored several types of trenches
(Fig. 7.1b). Mode separation increases with decreasing aperture opening and/or
longer oxidation time [137]. Depending on trench geometry and oxidization times,
the cavity modes are either degenerate (type B, Fig. 7.5a) or non-degenerate (type

135

7 High-Frequency Single-Photon-Source with Polarization Control

Figure 7.5: (a) Cavity mode spectrum for a type B device (depicted in the inset)
showing polarization-degenerate modes (Q factors up to 50,000). (b) Polarizationresolved spectra of a type D device (depicted in the inset) showing linear polarization splitting of 270 eV for the first-order cavity mode (1st). H indicates
horizontal polarization and V vertical polarization. (c) Temperature tuning of a
single QD exciton transition through the non-degenerate cavity modes. Coupling
to either of the H- or V-modes is manifested by an enhanced lifetime from about
800 ps off resonance to 300 ps on resonance (y axis). Insets show spectra taken
for different QD/mode detuning conditions.

136

7 High-Frequency Single-Photon-Source with Polarization Control


D, Fig. 7.5b), with splitting energies of about 300 meV and a contrast ratio of 57:1
between H- and V-polarized modes. For comparison to the results of Figs 7.2-7.4,
we chose a type D device containing X0 . Method A (cryostat heater) was used to
achieve QD/mode resonance (Fig. 7.5c). Coupling of a single QD to either the H
or V mode is manifested by a Purcell factor of 2.7, which is comparable to the type
B devices. The measured intensity enhancement is about twofold lower compared
with the type B devices, as expected if only one QD polarization state is coupled
out using a nondegenerate mode. The ability to control the polarization state on
the chip will thus reduce the overall SPS device efficiencies by a factor of two.
The switching speed using temperature is nevertheless very limited. To overcome
these limitations Stark-shift tuning [147] in an applied electrical field can be used.
Our rugged cavity trench design offers possibilities to integrate additional lateral
gates; this will be the subject of another study.

7.5

Summary and Conclusion

Ultrahigh SP rates are a combination of several effects. First, large geometrical


extraction efficiency is provided by the cavity design. Second, the observed single
QD Purcell effect contributes a factor of from two to four. Third, the positive bias
voltage creates a fivefold intensity enhancement at comparable optical injection

137

7 High-Frequency Single-Photon-Source with Polarization Control


levels, reducing background contributions that otherwise compromise SP purity.
Finally, QDs loaded with X effectively reduce dark states, which further increase
the SPS efficiency by a factor of three. Together, these effects contribute to the
observed record high rates of SPs on demand up to 31 MHz, and promise rates
above 116 MHz as investigated by the c.w. experiments. Nevertheless, devices
still do not approach the SE lifetime limit of about 870 MHz (ref. [19]). One
reason for this is that non-geminate-carrier capture allows for a single-hole capture
directly after the X state has decayed into the single electron state. This leads to
formation of X0 rather than X , and dark-state configurations are again possible.
In conclusion, we fabricated an SPS by integrating high-Q (50,000) GaAsAlGaAs microcavities with embedded oxide-tapered apertures, a rugged trench
design, as well as buried electrical gates allowing controlled loading and Purcell
tuning of individual InAs QDs. We observed single photons at record high rates
up to 31 MHz, and SP extraction efficiencies up to 38% if negatively charged
excitons were used. Additionally, we demonstrated tunability of the spectral mode
degeneracy and also controlled coupling of single QDs to either H- or V-polarized
cavity modes. This type of SPS is of direct interest for applications in quantum
information science.

138

7 High-Frequency Single-Photon-Source with Polarization Control

7.6
7.6.1

Methods
Sample Preparation

Devices are grown by molecular beam epitaxy on a semi-insulating GaAs (100)


substrate with a 0.1-mm buffer layer. There are four sections in the structure, the
bottom mirror, the active region, the oxide-aperture region and the top mirror as
described in more detail in our previous study [137]. The optical GaAs cavity is
Si n-doped up to 25 nm below the single layer of InAs QDs, providing a tunnel
back gate. The fifth GaAs DBR period above the aperture region is Si n-doped,
providing a top gate. Two etch windows with selective countersink depth are
created using reactive-ion etching in Cl2 plasma away from the SPS devices and
contacted using a Ti/Pt/Au metallization. Cavity trenches were formed by optical lithography and reactive-ion etching to penetrate approximately five mirror
periods into the bottom DBR. This etches small trenches in the surface that are
used to define the AlOx oxidation front. Tapered AlOx apertures are formed laterally during steam oxidation within the cavity area, creating inner openings of
0.52 m, effectively confining the mode volume [137].

139

7 High-Frequency Single-Photon-Source with Polarization Control

7.6.2

Optical Characterization

Micro-photoluminescence (micro-PL) measurements were performed using a


microscope objective with numerical aperture N A = 0.4. Photon coincidences
were recorded as a histogram of startstop events and provide a measure of the
second-order correlation function g (2) ( ) = hI(t)I(t + )i/hI(t)i2 , where hI(t)i is
the expectation value of the photon intensity at time t. Additional details can be
found in refs [137] and [87]. Time-resolved single QD measurements have been
performed under pulsed TiSa-laser excitation (150 fs pulsewidth) at a repetition
rate of 82 MHz. The single QD PL signal was spectrally filtered with a 1-nm
bandpass and coupled into a single-photon detector with about 130 ps timing
jitter.

7.6.3

Data Analysis

Measured SP count rates were corrected for multiphoton emission events by


multiplying by (1g (2) ( = 0))0.5 as introduced in ref. [135]. A detection efficiency
of 13% was estimated by measuring losses of all optical components at 920 nm
and by using the detector efficiency. The total amount of single photons emitted
into the first lens is 7.7 times the value shown on the y axis in Fig. 7.4. Aperture
losses at the first lens can be neglected because our high-Q cavity emits into a

140

7 High-Frequency Single-Photon-Source with Polarization Control


far-field angle of only 10 , in contrast with earlier work [135, 136] with low-Q
cavities where 78% of the light was lost at the first lens. With these corrections,
our highest SP rate with on-demand character is 31 MHz. Highest measured
SP rates of 200 kHz were reported in ref. [136] under resonant excitation. For
comparison with our experiments the off-resonance value of 50 kHz should be
used, but no detection efficiency was given in ref. [136]. We have thus used the
stated SPS collection efficiency from ref. [136] of 8% multiplied by the repetition
rate, yielding 6 MHz. From this number we estimate a fivefold improvement
compared with former works.

7.6.4

X /X0 Enhancement Factor

The bare X /X0 ratio from the c.w. experiments is 4.6 (10.1 MHz/2.2 MHz).
Because it is difficult to control the charging state and simultaneously also the
QD/mode resonance within the same device, we compared two very similar devices, one loaded with X0 and tuned into resonance and one with X . Thus, it
is necessary to correct for the difference between the devices, which can be taken
from the pulsed experiments close to saturation (4.0 MHz/2.7 MHz), leading to a
corrected improvement factor of 3.1. (See Sec. 7.7 for a theoretical foundation in
a rate-equation analysis.)

141

7 High-Frequency Single-Photon-Source with Polarization Control

7.7

Rate Equation Model

To quantitatively describe the enhancement of single photon emission from


a quantum dot in the X state compared to an X 0 state, we developed a rate
equation model for the excitonic populations. The simpler case to consider is the
population of the X state and include the presence of the biexciton state, XX ,
when driven at a rate, Rp . We assume that the relaxation rate from higher level
states is much faster than all other rates and can be neglected, thus, the level
scheme for this system is shown in Fig. 7.6a.
The rate equations for the level populations are
d
X = Rp (1 X ) X X + XX XX Rp (1 XX )X
dt
d
XX = Rp (1 XX )X XX XX ,
dt

(7.1)
(7.2)

where X is the spontaneous emission rate from the X and XX is the spontaneous emission rate from the XX . Note that the term (1 X ) represents
the saturation of the X and the term (1 XX )X represents the saturation of
the XX and enables pumping of that state only if the X state is populated. In
contrast, the rate equations for the X 0 state are slightly more complicated by the
presence of the dark state. The level scheme is shown in Fig. 7.6b. If we include
the spin flip rate, f , which couples the bright state to the dark and only allow

142

7 High-Frequency Single-Photon-Source with Polarization Control

b
XX

XX

Xb

Xd

vac]

vac]

Figure 7.6: The level scheme for both the X and the X 0 .

143

7 High-Frequency Single-Photon-Source with Polarization Control


the XX to decay to Xb0 , the rate equations become
d 0 1
X = Rp (1 Xb0 Xd0 ) X 0 Xb0 f (Xb0 Xd0 )
dt b
2
+ XX XX Rp (1 XX)Xb0
d 0 1
X = Rp (1 Xd0 Xb0 ) + f (Xb0 Xd0 ) Rp (1 XX)Xd0
dt d
2
d
XX = Rp (1 XX)(Xd0 + Xb0 ) XX XX.
dt
Note that there is a factor of

1
2

(7.3)
(7.4)
(7.5)

in front of the pump rate for both the bright and

dark state due to the equal pumping assumption, which is to say that all carriers
have become uncorrelated during the above band gap pump and subsequent relaxation process. Also, the term (1 Xb0 Xd0 ) prevents simultaneous occupation
of Xb0 and Xd0 . Finally, note that the XX can be populated if either Xb0 or Xd0
is populated and that the XX solely decays to Xb0 . We solve these equations in
the steady state to obtain the level populations as a function of Rp . The resulting
single photon count rate from the Xb0 or X is determined by multiplying the
population by the appropriate spontaneous decay rate, X 0 or X .
The results are shown in Fig. 7.7a for parameters obtained in our experiment
with the exception of XX which was taken from literature and modified to account
for the presence of the cavity. Clearly, at sufficiently high pump rates there are
many more photons from the X than from the X 0 while at low pump rates
they are the same. Fig. 7.7b shows the ratio of the single photon rates for the
144

7 High-Frequency Single-Photon-Source with Polarization Control

b
1.00
0.50

2.5

Xo

2.0

0.10
0.05
0.01

3.0

XXo

Count Rate HGHzL

1.5

0.1
1
10
Pump Rate HGHzL

100

1.0
0.01

0.1

1
10
Pump Rate HGHzL

100

Figure 7.7: The rates and ratio as a function of Rp . Parameters measured in


the experiment: X = 2.27 GHz, X 0 = 2.56 GHz, f = 0.33 GHz. XX =
XX = 0.5 GHz was chosen based on literature values modified to account for the
presence of the cavity.
same parameter set, which changes from 1 at low pump rates to around 3 at
high pump rates. At low pump rates, the Xb0 and X have the same population
filling because there is enough time between excitation events such that the dark
state can undergo a spin flip, become a bright exciton and recombine optically.
Once the pump rate becomes fast compared to the spin flip time, the biexciton
can become populated and creates a path for the dark state to be pumped into a
bright state. One might think then that at high pump rates, the X only provides
an enhancement factor of 2 because of the biexciton pumping, but this is only true
if the biexciton lifetime is fast compared to the pump rate. In the case where a
cavity is involved, emission which is not in resonance with the cavity mode is
inhibited. This inhibition of the biexciton lifetime leads to a reduction in how fast
the dark state can be pumped into the bright state, and causes the X to have an

145

7 High-Frequency Single-Photon-Source with Polarization Control


emission enhancement which is greater than 2. Thus, the X 0 has a reduced single
photon rate due to the presence of a dark state, which even when pumped into the
bright state through the biexciton, becomes saturated at pump rates which are
comparable to the biexciton decay rate. Therefore, at pump rates greater than
the biexciton decay rate, the X can produce more photons than the X 0 . This
difference is exacerbated in a cavity because bright transitions which are resonant
with the cavity mode become enhanced and transitions which are off-resonant are
inhibited.

146

Chapter 8
Electrically-Injected
Single-Photon-Source

8.1

Introduction

A high flux of triggered single photons is a necessary component to many quantum computation and quantum communication schemes [10, 148, 149]. Many
different experimental avenues to achieve a bright source of on-demand single
photons are currently being pursued including heralded single photons from parametric downconversion [150], single molecules [134], single atoms in cavities [151],
and defect centers in diamond [48]. In order to produce a scalable, all-on-chip
device, electrical generation of single photons is desirable. Until recently, experi-

147

8 Electrically-Injected Single-Photon-Source
mental demonstrations of a single-photon diode structure have focused on using
single, self-assembled quantum dots (QDs) to create single photons, but the ultimate photon flux has been limited by poor capture efficiencies due to the high
index contrast between the growth semiconductor and vacuum [52]. Further improvements have been made by placing the QD in a low quality factor, Q = 47,
1-dimensional cavity structure, but is limited to a factor of 5 increase because of
the low Q [152]. The natural extension is to embed the QD in a full, 3-dimensional
cavity with a high quality factor, but contacting a single cavity without destroying
it is a considerable technological challenge. Two different groups have reported
success creating such a device but one was in a lossy, Q=100 cavity [153] and the
other could not verify single photon operation with electrical injection [154]. Very
recently, the first group reported electrical injection into a cavity with Q 3000
[155].

8.2

Micropillar Design for Electrical Injection

In this chapter, we demonstrate a bright, electrically-injected source of single


photons from a QD coupled to a high quality, oxide-apertured micropillar cavity.
The main advantage of our cavity design when compared to others such as etched
micropillars, microdisks, or photonic crystals is that the lateral confinement of

148

8 Electrically-Injected Single-Photon-Source
the cavity mode is due to a thin aperture of Alx Oy [123], as shown in Fig. 8.1(a).
Because the index of refraction of Alx Oy is lower than the unoxidized AlAs, the
cavity mode is gently confined to the middle of the cavity and is prevented from
scattering at the rough sidewalls, enabling a Q greater than 45,000 with a mode
volume, Vef f , of 35( n )3 [137]. To overcome the contacting obstacle, we use a robust
cavity design which enables global contacting of several micropillar cavities while
maintaining a very large quality factor [156]. The main innovation of this design
is to define the oxidation front by etched trenches, as shown in Fig. 8.1(b), so
the cavity is electrically connected to the rest of the sample by thin material
bridges. To achieve efficient current injection into the device, this work employs
an insulating mask to contact a single microcavity.
The sample structure is grown by molecular beam epitaxy and is comprised
of an active region and an aperture region in between two distributed Bragg reflector (DBR) mirrors. The mirrors are made of alternating periods of GaAs and
Al0.9 Ga0.1 As, each

4n

thick. The

-thick
n

active region contains a single layer of

self-assembled InGaAs/GaAs QDs. The QDs are partially covered with GaAs
before annealing, and then are capped with GaAs, resulting in a blue-shift of the
QD emission to 940 nm [26]. The aperture region is composed a pure AlAs layer
between a Al0.89 Ga0.11 As and a Al0.75 Ga0.25 As layer. This is designed to provide
a linear taper to the aperture. The bottom mirror is Si n-doped up to 500
A

149

8 Electrically-Injected Single-Photon-Source

(a)

Ohmic p-contact
p-DBR
Oxide Aperture
QD layer

Ohmic n-contact
n-DBR

(b)

Figure 8.1: (a) A layer schematic of the device for efficient electrical injection of
single photons. (b) A CCD image of electroluminescence from the device operating
at 4 K with 10 mA DC excitation.

150

8 Electrically-Injected Single-Photon-Source
below the QD layer. The top mirror is carbon p-doped from the AlAs in the
aperture region up through the top DBR period. All DBR mirrors have a graded
Al content region between the GaAs and the Al0.9 Ga0.1 As DBR periods in order
to minimize the total electrical resistance of the device. N-contacts are fabricated
using a AuGe/Ni/Au recipe while p-contacts are fabricated using a Ti/Pt/Au
recipe. Electrically injected devices are isolated and individually contacted in order to minimize device resistances to acceptable values. A schematic of the device
structure is shown in Fig. 8.1(a). The resulting cavity modes can be imaged under
electrical injection onto a low-noise CCD camera and such an image is shown in
Fig. 8.1(b), demonstrating the Gaussian-like character of the fundamental mode.

8.3

Experimental Demonstration and Results

To study and verify the single photon stream emitted by the device under
electrical injection, the sample is placed in a Helium flow cryostat at 4 K. Carriers
are injected by a high frequency voltage source and subsequently relax into the
quantum dots where they optically recombine. Optical spectra are taken with a
1.25 m spectrometer and imaged onto a liquid nitrogen cooled CCD array, yielding
a combined spectral resolution of 30 eV at 900 nm with a total detection efficiency
of 15%. The electroluminescence spectra for a micropillar as a function of DC bias

151

8 Electrically-Injected Single-Photon-Source
is shown in Fig. 8.2(a) with the corresponding I-V curve, showing a clear turnon in current and electroluminescence near 3 V. The darkest regions correspond
to the cavity modes of the micropillar supported by QD emission. In between
modes, emission from non-resonant QDs is also visible. The turn-on voltage is
greater than what is expected in such a p-i-n device, and this can be attributed
to contact resistance. The spectrum of the micropillar cavity modes under optical
excitation is shown in Fig. 8.2(b). The slight ellipticity of the aperture gives rise
to a splitting of the fundamental mode into two linearly polarized modes [156]. A
two-Lorentzian fit is used to determine that Q > 11, 000 for both of these modes,
a record value for an electrically-driven device.
For some devices, emission from a spectrally nearby QD can be temperature
tuned into resonance with the fundamental cavity mode. In this way, single photons can be harvested into the mode via the Purcell enhancement, which can then
be mode-matched efficiently into our collection optics [156]. To verify single photon operation under pulsed electrical injection, we measure electroluminescence
using a small square wave at 50 MHz in addition to a DC bias just below turn-on
at 3 V. The electroluminescence spectrum is shown in Fig. 8.3(a) for the device
in Fig. 8.2 under an additional 1.1 V AC peak bias at 17 K. At this temperature, a QD transition is resonant with the higher energy polarization mode of the
split fundamental mode. The single photon quality is measured by selecting a

152

8 Electrically-Injected Single-Photon-Source

(a)

(b)

Figure 8.2: (a) A density plot of electroluminescence spectra as a function of DC


bias. Inset: The I-V curve taken simultaneously with the spectra. (b) Micropillar
cavity modes under optical excitation. Inset: A two-Lorentzian fit to the nonpolarization degenerate fundamental modes. The fit yields Q > 11, 000 for both
modes.
153

8 Electrically-Injected Single-Photon-Source
small bandwidth (100 eV) at the exit of the spectrometer and sending the signal
into a Hanbury-Brown & Twiss setup composed of two avalanche photodiodes
(APDs) and a time-to-amplitude converter card. This setup measures the second
order photon correlation, g (2) ( ), of the emitted light. The g (2) ( ) measurement
corresponding to the spectrum in Fig. 8.3(a) is shown in Fig. 8.3(b) under the
same excitation conditions. Because of the large count rates, sufficient statistics
could be reached with measurement times less than 5 minutes. The reduced size
of the 0 time delay peak in Fig. 8.3(b) is a measure of the single photon quality
of the emission and the value of g (2) (0) < 0.5 proves the single photon character.
Figure 8.3(c) plots the loss-corrected count rate (red squares) and g (2) (0) (blue circles) as a function of peak drive bias. The single photon character of the emitted
light saturates near 1.14 V peak bias, which corresponds to a single photon count
rate of 500 kHz. If single photons are generated at 50 MHz, this corresponds to a
capture efficiency of 1%, a comparable efficiency to the highest value reported in
literature [155].

8.4

Comparison to Optical Excitation

While the capture efficiency is large compared to other types of electrical


devices, it is small compared to optically generated single photons [156]. This

154

8 Electrically-Injected Single-Photon-Source

(a)

(b)

(c)

Figure 8.3: (a) AC electroluminescence spectra of a QD on resonance with the


H polarized fundamental mode from Fig. 8.2(b) at 17 K excited by a peak AC
bias of 1.1 V. (b) g (2) ( ) measurement of the QD emission from (a) through a 100
eV spectral filter, clearly showing single photon character. (c) Summary of the
loss-corrected count rate and g (2) (0) values as a function of peak AC bias. The
dashed blue line represents the maximum g (2) (0) for a single photon source. In this
device, the single photon character lasts until 1.14 V peak bias which corresponds
to a count rate of 500 kHz.

155

8 Electrically-Injected Single-Photon-Source
is due to the fact that the electrical generation of carriers creates uncorrelated,
free electrons and holes, many of which recombine before being captured by the
confining potential of the QD. Additionally, free carriers create a multitude of
recombination possibilities for QDs such as charged excitons and other charged,
multi-excitonic complexes. All of these transitions become more readily populated
under high excitation and thus light from other QDs can couple to the mode as
well, degrading the single photon character. In order to increase the single photon
count rate and capture efficiency, care must be taken avoid populating these other
states. Additionally, at low peak bias we measure long-lived correlations extending
as long as 200 ns, similar to what has been measured under optical excitation [136].
These correlations are no longer present at peak biases greater than 1.3 V. In order
to reduce the turn-on bias and to restrict current to pass solely through the cavity,
a ring contact, single VCSEL type processing scheme will need to be developed,
but the present devices act as a proof of concept.

8.5

Conclusion and Outlook

In conclusion, we have demonstrated a bright source of electrically-injected


single photons using quantum dots embedded in a high quality, electrically-gated,
oxide-apertured micropillar cavity. We verify the single photon character of the

156

8 Electrically-Injected Single-Photon-Source
emitted light stream by measuring the second-order coherence, g (2) ( ), under 50
MHz electrical excitation. From a measured single photon count rate of 75 kHz
we extract a count rate of 500 kHz into the first lens of our setup, corresponding
to a capture efficiency of 1%. These results were enabled by coupling the quantum
dot to a cavity with Q = 11, 500, a record high for an electrically driven device.
This demonstration verifies that this type of device has great promise for future
applications in quantum communication and computation.

157

Chapter 9
Cavity QED with Charged
Quantum Dots

9.1

Introduction

Hybrid quantum information schemes combine the coherence properties and


ease of manipulation of photons with the scalability and robustness of local quantum systems. Examples of proposed local quantum systems include electron
spins in quantum dots, defect centers in diamond, and trapped atoms or ions
[8, 157, 158]. Hybrid schemes such as quantum repeaters and distributed quantum networks use the coupling between the local quantum system (qubit) and the
optical field to reversibly map the quantum state of an injected photon onto the

158

9 Cavity QED with Charged Quantum Dots


state of the local system [159162]. Other hybrid schemes use a joint measurement of emitted photons, which are entangled with their respective local qubit,
to perform remote gate operations on the two spatially separated local systems
[11, 148, 163, 164]. This latter scheme can be used to create entanglement between
many local qubits as needed for cluster state quantum computation.
Within this architecture, the coupling between the photon and the local quantum system should be efficient, which means the quantum system must be placed
in a high-quality microcavity where it can interact with photons in the cavity
mode, as described by cavity quantum electrodynamics (QED). The cavity mode
should also be matched to an external optical mode. In general, such schemes
usually operate deep in the weak coupling (Purcell) regime, just below the onset
of strong coupling, and employ optical pulse shaping to maximize the information transfer from the photon to the local qubit [159, 160, 162]. Furthermore,
for cluster-state and distributed quantum computation, the hybrid system must
be scalable. Here we report on the first experimental demonstration of a system that satisfies these requirements in the solid-state. This system is composed
of self-assembled quantum dots (QDs) (density 10m2 ) at the center of an
oxidation-apertured micropillar cavity with integrated doped layers that enable
an external bias to apply an electric field across the QD. This field causes carriers
to tunnel in and out [56] of the QD, changing the QD charge state, and induces

159

9 Cavity QED with Charged Quantum Dots


the quantum confined Stark effect [53, 165], shifting the emitted photons energy.
While cavity QED has been studied using quantum dots for several years, this has
been done using a neutral exciton [22, 23]. Neutral excitons, bound electron-hole
pairs, have been proposed as qubits [166] but seem problematic due to their quick
spontaneous decay, 1ns in GaAs, and fast dephasing [167]. The local qubit
in our system is the spin of a trapped electron [168171], which interacts with
the cavity mode through the addition of an exciton, forming a short-lived trion
state. Since the polarization of the emitted photon is correlated with the spin
state of the remaining electron, the trion acts as a readout channel of the spin
[161]. Additionally, the micropillar cavity geometry is such that the fundamental
mode is a doubly degenerate HE1,1 mode, which mode-matches well to external
modes due to its Gaussian-like shape [121]. Here, we report on two variations of
the solid-state cavity QED system, one optimized to operate in the charge-tuning
regime, and the other in the Stark-tuning regime.

9.2

Sample Design

The demonstration of an electrically-gated QD embedded at the anti-node of


a high Q cavity mode has become feasible through a series of scientific advances.
Firstly, the development of vertical-cavity surface-emitting lasers (VCSELs) with

160

9 Cavity QED with Charged Quantum Dots

b 60
y HmL

50
40
30
20
20

DBR

Dot Layer
Quantum
BR
Graded D

re

u
Oxide Apert
n doping
p doping

40
x HmL

50

60

c 44
42

y HmL

21 period top

30

40
38

m DBR
riod botto

32 pe

36
34

36

38 40
x HmL

42

Figure 9.1: (a) Schematic of the scalable solid-state cavity QED system based
on electrically gated self-assembled QDs embedded in oxide-apertured micropillars.(b) Two-dimensional reflectivity scan of a micropillar cavity taken with a laser
resonant with the cavity mode. The mode in the center can be seen clearly as a
dip in the reflected signal. (c) Higher resolution reflectivity scan taken in a 10 m
by 10 m area containing the mode, as depicted in (b).

161

9 Cavity QED with Charged Quantum Dots


oxide apertures in the GaAs/AlGaAs material system enabled the creation of cavities with small mode volumes, Vef f = 35(/n)3 , while maintaining a very high
Q [123]. Secondly, the addition of single, self-assembled InAs/GaAs QDs embedded at the axial anti-node of the cavity mode provided an atomic-like emitter to
couple to the optical mode [137]. Thirdly, the use of etched trenches to define the
oxidation front, as shown in Fig. 9.1, enabled both control over the polarization
degeneracy of the cavity modes as well as global electrical connection to an array
of solid-state cavity QED systems [156]. In the experiments presented here, the
oxidation time of the aperture is such to maximize the quality-factor, Q, while
minimizing the mode volume. As shown in Fig. 9.1b,c, the cavity mode is to
a good approximation Gaussian in lateral profile and fits to a waist of 2.2 m,
in agreement with measurements of the spacing between different lateral modes
[137].
Using a voltage source to create an electric field which properly drops over
the QD active region is complicated by the presence of nearby material interfaces
at each distributed Bragg reflector (DBR) period and at the oxide aperture region. These interfaces trap charges and result in the formation of charge domains,
which reduce the field dropped across the QD region and hence obstruct controlled
charging and Stark tuning [156]. To overcome these problems, a novel P-I-N device
structure was developed in which the intrinsic region does not include the oxide

162

9 Cavity QED with Charged Quantum Dots


aperture and the nearby p-doped Al0.9 Ga0.1 As DBR period is Al-content graded
to and from the adjacent GaAs layers as shown in Fig. 9.1a. The Al-content
grading prevents the formation of triangular potential wells that arise at abrupt
Al0.9 Ga0.1 As/GaAs interfaces. Furthermore, all doping concentrations are graded
such that the doped regions are easily contacted by countersink etching without
introducing unnecessary dopants near the QD region. The two variations of the
solid-state cavity QED system presented here have nominally the same growth
structure, but the average doping levels for the charge tuning system are 3.5 1018
cm3 (2.5 1018 cm3 ) for the n-doped (p-doped) layer whereas for the Stark tuning
system they are 7.0 1017 cm3 (7.5 1017 cm3 ).

9.3

Intra-cavity Charging

To investigate the intra-cavity charging, we first characterized QDs outside


of the etched cavities in the surrounding DBR mirror region, where the Purcell
effect is negligible. We monitored the photoluminescence spectrum (using a 1.25
m monochromator coupled with a CCD array) under 150 fs, 860 nm Ti:Sa laser
excitation with 50 nW average power as the applied bias is varied. A typical
trace for a single QD is shown in Fig. 9.2a. At approximately 18 V applied bias,
there is a transition for the QD emission line to a line that is 6 meV to lower

163

9 Cavity QED with Charged Quantum Dots


energy. This is the characteristic energy separation for the transition between
the neutral exciton, X 0 , and the singly electron-charged exciton, X [39, 56].
To verify the charge designations as X 0 and X , we also measured the timeresolved decay of the photoluminescence using an avalanche photodiode with a
time-to-amplitude converter. The result is shown in Fig. 9.2b, with curves taken
at biases below (above) 18 V labelled as dashed (straight). Because of the presence
of optically dark states, the X 0 decay traces have a distinctive bi-exponential
behavior, whereas the X decay are single exponentials [143, 156].
The same lifetime measurement is performed for QDs in the cavity region which
are on resonance with a polarization-degenerate fundamental mode as shown in
Fig. 9.2c (this was possible for approximately 10% of cavities), and the results
qualitatively replicate that of the bulk QDs. However, the effect of the high Q
cavity strongly reduces the emission lifetime by the Purcell effect. For some X
cases, this lifetime approaches the timing resolution of our experiment, which is
150 ps and is due to the APD timing jitter. Nonetheless, a deconvolved lifetime
of 13721 ps was obtained for the fastest X transition and 32115 ps for the
X 0 . This yields a Purcell enhancement, Fp = o /cav , of approximately 7 for
the X . We measure that on average (Fp = 2.8 0.22 for 4 X 0 transitions and
5.9 0.96 for 6 X ) the Purcell enhancement is stronger for X than for X 0 .
Because both transitions have similar lifetimes in the DBR region, resulting from

164

9 Cavity QED with Charged Quantum Dots

c 1.00

1.324

1.322

Counts Harb. unitsL

Energy HeVL

1.32

1.318

X-

1.316

0.50
0.20
0.10
0.05
0.02
0

1.314

17

4
6
time HnsL

10

d 1.0

19

0.8

1.00
0.50

gH2LHL

Counts Harb. unitsL

18
18.5
Applied Bias HVL

17.5

0.20

0.6
0.4

0.10
0.05

0.2

0.02
0

4
6
time HnsL

0.0
8

10

-40

-20

0
time HnsL

20

40

Figure 9.2: (a) Photoluminescence spectra as a function of applied bias for a QD


in the mirror region at 4 K. Near 18 V a clear transition from X 0 to X takes
place. (b) Emission decay traces for 10 QDs in the mirror region. Straight traces
are taken with 18.5 V applied bias (X ) and dashed traces are taken with 17.5
V applied bias (X 0 ). (c) Emission decay traces for QDs on resonance with cavity
modes for several different cavities. The straight (dashed) traces correspond to an
X (X 0 ) decay. (d) g (2) ( ) measurements for an X (straight) and X 0 (dashed)
transition.

165

9 Cavity QED with Charged Quantum Dots


similar oscillator strengths, one would expect both to have similar Purcell effects.
However, this is not found experimentally and may be due to a better matching
of the transition dipole moment to the cavity mode polarization for the X . This
could be explained by theoretical calculations beyond the standard techniques for
calculating the optical transitions of a QD [172]. In addition, the second-order
photon correlation function, g (2) ( ), was measured as shown in Fig. 9.2d for an X 0
and an X transition. While both clearly demonstrate single photon behavior,
the X is much cleaner due to its fast, single exponential decay. The measured
single photon (g (2) (0) < 0.25) count rate was typically 3 106 s1 for an X with
an 80 MHz pump rate, yielding a 25% extraction efficiency for the QD when
corrected for the optical and detection losses of the setup, which correspond to a
net efficiency of 15%.

9.4

Coherent Cavity Reflection Spectroscopy

While the lifetime measurements indicate that the emission is coupled to the
cavity mode, it does not yield a quantitative measure of the coupling strength,
g, directly. To do this, one must probe the coupled system coherently and we
accomplish this by measuring the reflectivity of the cavity-QD system [22, 23].
The reflection spectrum of the QD-cavity system can be derived from the Jaynes-

166

9 Cavity QED with Charged Quantum Dots


Cummings Hamiltonian using the input-output formalism and under sufficiently
weak probing of a symmetric cavity [173176], can be expressed as

R() = 1

[ i( QD )]
,
[ i( QD )][ i( c )] + g 2

(9.1)

where g is the emitter-cavity coupling, QD (c ) is the emitter (cavity) resonance,


is the dipole decay rate, and is cavity field decay rate. If there are no QD
transitions coupled to the cavity mode, the spectrum shows a single dip at the
cavity resonance with a width equal to the cavity field decay rate, , as shown in
Fig. 9.3a. For this micropillar, we can fit the data to obtain = 24.1 eV, which
corresponds to a Q factor of 27,000. The depth of this dip is a measure of how
well the probe beam is mode-matched to the cavity, and in this case the coupling
efficiency is greater than 96%. This remarkably high efficiency implies that the
mode has a Gaussian-like profile and that reliable information transfer at the
single photon level is feasible. If a QD transition is coupled to the microcavity,
the reflection spectrum is drastically altered. Figure 9.3b shows the reflection
spectrum of the cavity mode interacting with a single QD transition. By fitting
the reflection spectrum to Eq. 9.1, we obtain an emitter-cavity coupling of g
= 9.7 eV and an emitter decay rate of = 1.9 eV. Since g/ = 0.402, the
emitter-cavity system is deep in the Purcell (weak-coupling) regime and at the
precipice of the strong-coupling regime, g/ > 0.5, exactly in the region ideally

167

9 Cavity QED with Charged Quantum Dots

1.0

0.8

0.8

0.6

0.6

a 1.0

0.4

0.4

0.2

0.2

0.0

1.32408

1.32412 1.32416
Energy HeVL

1.3242

0.0

1.32408

1.32412 1.32416
Energy HeVL

1.3242

Figure 9.3: (a) Cavity reflection spectrum of an unloaded micropillar cavity measured by recording the reflected signal of a tunable-wavelength laser. Eq. 9.1 with
g = 0 plus a linear background is used to fit the data, yielding = 24.1 eV, which
corresponds to Q = 27,000. The depth of the dip in reflection is a measure of
the mode-matching efficiency, , to the cavity mode, and in this case we achieve
> 0.96. (b) Cavity reflection spectrum of a QD coupled to the micropillar cavity.
Eq. 9.1 plus a linear background is used to fit the data and we obtain g = 9.7 eV
and = 1.9 eV.
suited for hybrid quantum information schemes [11, 161163]. The spectrum,
with resolution limited by the probe laser linewidth, completely characterizes the
system. Additionally, it reveals the natural linewidth of the QD transition with
a signal much greater than achieved in transmission or differential transmission.
In conclusion, the combination of these results for the cavity QED system in the
charge-tuning regime demonstrates that it is ideal for hybrid quantum information
processing.

168

9 Cavity QED with Charged Quantum Dots

9.5

Intra-cavity Stark Tuning

We now turn to the Stark-tuning cavity QED system. Since the coupling
between the QD transition and the cavity mode depends on the relative spectral
detuning, an external control is necessary to reach spectral resonance. In QD
systems without electrical gating, this control is commonly achieved by adjusting
the sample temperature [23]. However, this control typically decreases coherence
within the system through higher phonon occupations. An applied electric field
also tunes the QD transition energy via the Stark Effect without the negative
effect of temperature. In order to illustrate this effect and potential applications,
we utilized a polarization non-degenerate cavity mode. As mentioned in Ref. [156],
an engineered ellipticity of the aperture lifts the polarization degeneracy, creating
two orthogonal linear polarization modes (denoted as H and V) as illustrated
in Figure 9.4a. Because the Q factor is very high (40,000), the modes can be
spectrally separated by as little as 50 eV and still be resolved. This enables the
quantum dot transition to be Stark-shift tuned into resonance with two modes as
shown in Fig. 9.4b. Note that the dependence is nonlinear with bias as expected
for the Quantum Confined Stark Effect [53, 165].
By Stark-shift tuning the QD emission, a polarization-dependant Purcell effect
is observed on resonance with each mode. Stark shift tuning as opposed to current

169

9 Cavity QED with Charged Quantum Dots

Counts Harb. unitsL

Counts H105 s-1L

4
3
2
1
1.314

1.316

1.318 1.320
Energy HeVL

1.322

1.324

0.02
0.01

H mode reference QD

1.316
1.3155

0.7
Lifetime HnsL

Energy HeVL

0.20
0.10
0.05

d 0.8

0.5

6
7
8
Applied Bias HVL

2
3
time HnsL

9.5 V
8.33 V
9.2 V

5.0 V

0.4

7.62 V

7.0 V

8.86 V
0.2
1.3154 1.3155 1.3156 1.3157 1.3158 1.3159 1.3160
Energy HeVL

1.3145
4

0.6

0.3

V mode QD of interest

1.315

7.62 V
8.2 V
8.86 V
9.5 V

-1

1.317
1.3165

1.00
0.50

10

Figure 9.4: (a) Micro-photoluminescence spectra of non-degenerate optical modes


in an etched-trench oxidized microcavity. H (V) polarized modes are black (grey).
Inset: SEM image of an etched trench microcavity. (b) Photoluminescence spectra
as a function of applied bias for two QDs (labeled QD of interest and Reference
QD) and two nondegenerate fundamental cavity modes (labeled H mode and V
mode). (c) Lifetime traces for a few bias settings; 7.62 V, 8.2 V, 8.86 V, 9.5
V. (d) Deconvolved lifetimes as a function of emission energy with fit clearly
demonstrating a Purcell effect at both mode resonances.

170

9 Cavity QED with Charged Quantum Dots


induced heating was confirmed by observing a constant QD linewidth over the
tuning range. For several applied biases, the QD emission decay curve is measured,
see Fig. 9.4c, and the extracted lifetime is plotted as a function of spectral position
as shown in Fig. 9.4d. The dips in the transition lifetimes measured on resonance
are a clear consequence of the Purcell effect. The lifetimes at the resonance of
each mode is measured to be around 220 ps as shown in Fig. 9.4d, approximately
5 times shorter than the bulk lifetime. The appearance of a bi-exponential, most
prevalent for the on resonance biases, in Fig. 9.4c is attributed to a small fraction
(4%) of photons collected from QDs outside the mode volume. Stark-shift tuning
when used in addition to charge tuning constitutes a completely bias-controlled,
solid-state cavity QED system.

9.6

Summary and Conclusion

In conclusion, we presented a solid-state cavity QED system which has near


ideal properties for photon electron-spin coupling as needed for hybrid quantum
information processing. The unique features of our system are intra-cavity electron
charging, near perfect mode-matching, polarization control of the cavity modes,
and operation deep in the Purcell regime. In addition, the cavity-QD coupling
can be controlled via the Stark Effect, which has applications for quantum and

171

9 Cavity QED with Charged Quantum Dots


classical communication. The combination of this work with spin initialization,
manipulation, and readout [3942, 177] will bring the implementation of solidstate hybrid quantum information protocols within reach.

172

Chapter 10
Polarization-switchable Single
Photon Source using the Stark
effect
This chapter is based on Applied Physics Letters 93, 091118 (2008). Copyright
2008 by the American Institute of Physics.

10.1

Introduction

Self-assembled quantum dots (QDs) embedded in microcavities have been used


to demonstrate a range of novel optical devices with potential for future tech-

173

10 Polarization-switchable Single Photon Source using the Stark effect


nologies. To date, this has included bright sources of single photons [135, 156],
ultra-low threshold lasers [87], and cavity quantum electrodynamic (QED) devices
[15, 16, 98]. One of the advantages of the QD-microcavity system is that the QD
behaves like an atom which is permanently trapped in its semiconductor matrix,
negating the use of trapping lasers as in atomic cavity QED systems [178, 179].
The disadvantage of using QDs is that they are not identical like atoms. Because
of this fact, the QD microcavity system must take advantage of an external control to achieve spectral resonance. To date, this has been done by shifting the
QD with temperature [15, 16] or by shifting the microcavity mode by depositing
a small amount of material onto the structure [110, 111]. In principle, the QD
energy could be tuned with an electric field by the quantum confined Stark effect
[53, 180]. In this chapter, the Stark effect is used to demonstrate a novel single
photon device where a QD embedded in an oxide-apertured micropillar cavity is
electrically tuned into resonance with two orthogonally-polarized cavity modes,
creating a polarization-switchable single photon source.

10.2

Sample Design and Fabrication

The types of cavities used in the experiments are oxide-apertured micopillars


[123, 137]. In these micropillars, optical confinement in the growth direction is

174

10 Polarization-switchable Single Photon Source using the Stark effect


provided by two distributed Bragg reflector (DBR) stacks made of alternating
layers of GaAs and Al0.9 Ga0.1 As. Lateral confinement is provided by the index
contrast between the unoxidized AlAs and the AlOx in the aperture region. This
type of gentle confinement has been shown to achieve a quality factor Q of up
to 48,000 [137]. Additionally, the aperture can be engineered to be non-circular,
for example by exploiting the fact that the oxidation proceeds at different rates
along different crystalline axes [181]. The ellipticity of the aperture breaks the
polarization degeneracy of the fundamental HE1,1 mode [146, 156], leaving two
linearly polarized modes with a small energy splitting, as shown in Fig. 10.1. In
the cavity considered here, the Qs were 18, 920 and 19, 320 for the horizontally
and vertically polarized modes respectively; the polarization visibility was 0.95
for each mode and the mode splitting was 194 eV.
Samples are grown by molecular beam epitaxy on a semi-insulating GaAs
[100] substrate. There are four sections in the structure, the bottom mirror,
the active region, the oxide-aperture region, and the top mirror as described in
more detail in a previous study [137]. The active region contains a centered
InAs QD layer grown in the Stranski-Krastanow growth mode. The optical GaAs
cavity is Si n-doped up to 20 nm above the single layer of QDs providing a back
gate. The first GaAs DBR period in the bottom mirror is C p-doped providing
a top gate. Two etch windows with selective countersink depth are created using

175

10 Polarization-switchable Single Photon Source using the Stark effect

Counts Ha.u.L

(a)
1.317

1.3148

Energy HeVL

1.3151

1.3154

1.3157

1.315
Counts Ha.u.L

Energy HeVL

1.316

H mode

(b)

1.314

1.3148

1.313
-50

V mode

(c)

Energy HeVL

1.3151

1.3154

0
50
100
Polarizer Angle HdegL

1.3157

150

Figure 10.1: (a) The spectra of the non-degenerate fundamental modes as a function of polarizer angle. (b),(c) The spectra of the H and V modes respectively
with Lorentzian fits yielding Q = 18, 920 and Q = 19, 320.

176

10 Polarization-switchable Single Photon Source using the Stark effect


reactive ion etch (RIE) in Cl2 plasma away from the cavities and contacted using
a Ti/Pt/Au metallization. Cavity trenches were formed by optical lithography
and RIE penetrating approximately 5 mirror periods into the bottom DBR. This
etches small trenches in the surface that are used to define the AlOx oxidation
front. Tapered AlOx apertures are formed laterally during steam oxidation within
the cavity area creating inner openings of 0.5-2 m. The doped layers enable
electrical gating of the sample that in turn creates an adjustable electric field over
the QD region. This external field causes the QD emission to spectrally shift, a
phenomena known as the quantum confined Stark effect [53].

10.3

Polarization Switching

All measurements are performed in a He-flow cryostat at 4.2 K. Microphotoluminescence spectroscopy is implemented by focusing 500 nW of 860 nm laser
through a high-power (NA = 0.42) objective onto the sample surface. The emission is then directed to a spectrometer equipped with a charge-coupled device
(CCD) array yielding a spectral resolution of 30 eV or to two avalanche photodiodes (APDs) for single photon correlation measurements. The sample is electrically connected to a function generator capable of operating up to 50 MHz. As
shown for the device in Fig. 10.2(a)-(c), emission from a QD can be electrically

177

10 Polarization-switchable Single Photon Source using the Stark effect


tuned into resonance with the modes from Fig. 10.1. Near 7.62 V, the QD emission is in resonance with the H-mode. Because of the Purcell effect, QD emission
into this mode is enhanced by the Purcell factor, Fp , while V-polarized emission
is inhibited by a factor denoted as in . As a result, the emission is dominantly Hpolarized with ratio Fp /in . Similarly, at 8.86 V the emission is in resonance with
the V-mode, and is dominantly V-polarized. In this way, the polarization of the
single photons emitted by the QD can be electrically switched between dominantly
H and V. Fig. 10.2(d) shows the spectra as a function of time with a 100 mHz
sinusoidal drive bias demonstrating the principle of polarization switching. This
frequency was chosen in order to provide a sufficient drive frequency to sampling
frequency ratio for the CCD array. The QD emission clearly oscillates at the drive
frequency between the two modes, emitting single photons whose polarization are
switched at the same drive frequency. In addition, another off-resonant QD can
be seen slightly higher in energy and spectrally oscillates at the drive frequency.
For pump powers used here, the second order correlation function g (2) ( ) always
has a value g (2) (0) < 0.15 as shown in Fig. 10.2(e), well below the minimum of
0.5 required to prove single photon operation. Note that in comparison to QD
Stark effect literature [53], the applied biases used here are much larger. This is
due to large contact resistances, but based on the amount of spectral tuning, the
electric field at the QDs is approximately 50 kV/cm and varies by approximately

178

10 Polarization-switchable Single Photon Source using the Stark effect

1.3165

(a)

(d)

1.3155
(e)

1.315

1.
0.75
0.5
0.25

gH2LHL

(b)

Energy HeVL

1.316

(c)

1.3145

Energy (eV)

-80 -40 0
t HnsL

10

15
t HsL

20

25

40

30

Figure 10.2: (a)-(c) The spectra for applied biases of 7.62 V, 8.33 V, and 8.86 V,
showing the QD in resonance with the H mode, in between both modes, and in
resonance with the V mode. (d) The spectra as a function of time with a 100
mHz sinusoidal driving bias. (e) The second order correlation function g (2) ( ) for
the QD in resonance with the H mode.
40 kV/cm.
To probe the quality of the polarization switching and to find its temporal
limit, we measured switching as a function of drive frequency. This was accomplished by using a dc bias to place the QD emission spectrally equidistant from
each mode, and then applying an additional ac square wave to oscillate the QD.
The amplitude of the square wave was set so that the QD could reach both modes
at low frequency. H and V polarized photons were separated at a polarizing
beamsplitter and each detected at an APD. The signal from the APDs was sent

179

10 Polarization-switchable Single Photon Source using the Stark effect

(b) 1.6
1.4
1.4
1.2
1.2
1.0
1.0
0.8
0.8
0.6
0.6
0.4
0.4
0 20 40 60 80 100 120
0
t HsL
(c) 1.6
(d) 1.6
1.4
1.4
1.2
1.2
1.0
1.0
0.8
0.8
0.6
0.6
0.4
0.4
0 2 4 6 8 10 12
0
t HsL
Counts

10

20
30
t HsL

40

2
3
t HsL

Counts

Counts

Counts

(a)1.6

Figure 10.3: (a)-(d) Normalized and time-integrated APD counts for polarization
switching at drive frequencies of 10, 30, 100, and 300 kHz. H(V)-polarization is
denoted in dashed magenta (straight blue).
to a multi-channel analyzer triggered by the function generator. Normalized integrated photon counts as a function of time for each polarization are shown in
Fig. 10.3 for drive frequencies of 10, 30, 100, and 300 kHz at constant drive amplitude with H (V) detection events in dashed magenta (straight blue). As can be
seen clearly in the 10 kHz measurements, the signals have both a slow and fast
temporal component, with the slow component vanishing for the higher frequency
measurements. This will be explored in detail in the next section.

180

10 Polarization-switchable Single Photon Source using the Stark effect

10.4

Theoretical Model and Discussion

For a given drive frequency, the quality of the polarization switching can be
quantitatively described by the average of the visibilities of the count rate oscillations as shown in Fig. 10.3, where the visibility is defined as Vp (Imax
Imin )/(Imax + Imin ) for each polarization measurement. Figure 10.4 summarizes
the average visibility as a function of drive frequency with constant drive amplitude. Because the system has a limited frequency response, the visibility is
expected to drop off once the frequency cutoff is surpassed and the amplitude
of the spectral oscillation of the QD isnt large enough to reach the modes. As
Fig. 10.4 clearly shows, there are two cutoff frequencies, one around 10 KHz and
another near 1 MHz. This was expected as two temporal components were visible
in Fig. 10.3(a). We hypothesize and later verify that the lower frequency cutoff
corresponds to the current-induced heating of the sample and the higher frequency
corresponds to the RC electrical cutoff. To model the dependence of the visibility
on drive frequency, the emission intensity of each polarization as a function of QD
spectral position must be known. This dependence can be well approximated by
a Lorentzian with width (height) determined by the Q-factor (Purcell factor, Fp ),
and a constant background due to the inhibited decay into modes other than the

181

10 Polarization-switchable Single Photon Source using the Stark effect


cavity, in . The average visibility can then be expressed as
Vp =

42 Fp
,
in ( 2 42 )2 + (Fp + 2in ) ( 2 + 42 ) 2 + (in + Fp ) 4

(10.1)

where is the QD spectral tuning amplitude, is the mode splitting, and is


the mode linewidth. and are known from the optical spectrum (Fig. 10.1) and
Fp was measured by time-resolved spectroscopy to be approximately 5 for both
modes. in can be determined as follows. At sufficiently low drive frequency the
QD spectrally reaches both modes so = /2. This value of gives the highest
possible visibility for this device and is measured to be 0.71 0.007, determining
in to be (0.85 .03)% of the bulk decay rate. The maximum switching visibility
is less than the pure mode polarization visibility of 0.95 because of Fp and in .
The frequency dependence of the QD spectral tuning amplitude = () can
be described as the sum of two incoherent processes, each with a specific cutoff
frequency. This function is expressed as
#
"
/2
T AT
RC
p
,
() =
+p
2
AT + 1
4 2 2 + RC
4 2 2 + T2

(10.2)

where T (RC ) is the thermal (RC) cutoff frequency, and AT is the relative
strength of the thermal process. By substituting Eqn. 10.2 into Eqn. 10.1, the visibility data can be fit and the curve with best fit parameters is shown in Fig. 10.4.
From this fit, we determine the thermal cutoff frequency, T , to be 14 4.1 kHz
and the RC cutoff frequency, RC , to be 1.1 0.10 MHz. A simple estimate of the
182

10 Polarization-switchable Single Photon Source using the Stark effect

1.00

Visibility

0.70
0.50
0.30
0.20
0.15
0.10
10

100

1000
104
Frequency HHzL

105

106

Figure 10.4: The average polarization visibility as a function of drive frequency


with constant amplitude. A fit to the data using Eqs. 10.1 and 10.2 yields a
thermal cutoff frequency of 14 4.1 kHz and an RC cutoff frequency of 1.090.10
MHz.

183

10 Polarization-switchable Single Photon Source using the Stark effect


thermal cutoff using the low-temperature specific heat and thermal conductivity
of GaAs as well as the device size, yields a value of approximately 10 kHz, in agreement with the fit. Furthermore, additional low-temperature capacitance-voltage
measurements were performed to verify the origin of the high frequency cutoff.
The resistance and capacitance near the working voltages of the experiment were
found to be 1.7 k and 0.49 nF respectively, yielding an RC frequency cutoff of
1.2 MHz, in agreement with the fit.

10.5

Conclusion and Outlook

In conclusion, we have demonstrated a method for on-chip, single photon polarization switching by Stark tuning emission from a single QD into resonance with
two, high-quality, orthogonally-polarized cavity modes. While our demonstration
was temporally limited by the thermal conductivity of the device and ultimately
by the RC cutoff, future implementations can be made faster by reducing the
current and by decreasing the contact resistance and parasitic capacitance. In
principle, switching rates should be able to approach gigahertz frequencies, similar to recent results on fast QD charging [182, 183]. Additionally, the polarization
visibility was limited in this case to 0.71 by Fp and in . The visibility could be
increased to 0.96 if in and Fp had values like found in photonic crystal defect cav-

184

10 Polarization-switchable Single Photon Source using the Stark effect


ities [87] and the mode splitting was increased by a factor of 2. The combination of
these advances will enable high visibility switching at gigahertz frequencies. This
kind of switching could enable high frequency lock-in measurements of single QDs
(Ref. [182]) or fast loading of a cavity for bias-controlled optical nonlinearities
in cavity QED [184186]. Furthermore, this type of single photon polarization
switching could have applications as an excitation mechanism for other quantum
systems or for classical optical communication.

185

Chapter 11
Conclusion
In summary, this dissertation detailed several novel quantum optics experiments that used self-assembled quantum dots embedded in semiconductor microcavities. The first set of these experiments revealed that a remarkably small
number of QDs are necessary to act as a sufficient gain medium for lasing in
photonic crystal cavities. This surprisingly result requires that the QDs act far
outside of a simple, atomic-like model. Due to the weak turn-on behavior of
these nanolasers, the second experiment used a family of these cavities to study
the onset of lasing for different system sizes by measuring the photon statistical
transition of the cavity mode. These results are compared to measurements of
a conventional laser diode. The next set of experiments finds solutions to the
difficulties inherent in achieving strong coupling in QD-photonic crystal cavity

186

11 Conclusion
QED. The first uses the cold adsorption of material from inside the cryostat onto
the sample surface to spectrally shift the photonic crystal resonance. The second
experiment illustrates a technique to optically locate a single QD to a precision
of 30 nm. Then, a photonic crystal cavity is fabricated about this position and
strong coupling is unequivocally identified.
The experiments in the second half of the dissertation use quantum dots embedded in electrically-gated micropillar cavities. The first experiment describes
the optical microcavity design and relevant cavity QED parameters. Next, these
cavities are shown to be an ultra-bright source of single photons by using the
charged exciton state of the QD rather than the neutral exciton. Then, using the
gates to drive a current through the cavity enabled a demonstration of a bright
source of triggered, electrically-generated single photons. The next experiment
used the intra-cavity electrical gates to controllably charge the QD. In addition,
coupling deep in the Purcell regime as well as a very high mode-matching efficiency are precisely measured by coherent reflection spectroscopy, demonstrating
the potential for efficient implementation of quantum information schemes using
single electron spins. Finally, these gates are used to generate an intra-cavity electric field which spectrally tunes the QD emission by the Stark Effect. This tuning
is used to create a polarization switchable single photon source when combined
with an elliptical micropillar cavity.

187

11 Conclusion
Taken as a whole, these experiments explore the coupling between a quantum
dot and a microcavity mode from the single photon regime all the way to the
lasing regime. Concurrently, the optical properties of quantum dots are explored
from the simple atomic-like model to a more broadband emitter model. Finally,
the capabilities inherent in semiconductor device technology are used to manipulate the QD in novel ways, which may lead to important contributions to the
burgeoning field of solid-state quantum information.

188

Appendix A
Quantum Dot Fine Structure
The fine structure of the quantum states of self-assembled QDs like used in this
dissertation is largely determined by the short-range exchange interaction which
arises due to electron-hole spin coupling. The Hamiltonian for this interaction in
GaAs can be expressed as [187]
Hex =

ai Ji Si + bi Ji3 Si ,

(A.1)

i=x,y,z

where Ji and Si are the hole and electron spin operators in the ith direction, the
ai s and bi s are the spin-coupling coefficients, and the z direction is taken to be
the growth direction. Because the light-hole band is split off from the heavy-hole
band, holes in the valence band have Jz = 3/2 and electrons in the conduction
band have Sz = 1/2. Furthermore, neglecting the light hole band leaves reduces

189

A Quantum Dot Fine Structure


Eq. A.1 to
Hex = [az Jz Sz +bz Jz3 Sz +

bx 3
by
(J+ +J3 )(S+ +S )+ (J+3 J3 )(S+ S )], (A.2)
16
16

where J = Jx Jy and S = Sx Sy . Note that the action of these operators


on the state | j, jz i is [188]
J+ | j, jz i =

p
(j jz )(j + jz + 1) | j, jz + 1i ,

(A.3)

J | j, jz i =

p
(j + jz )(j jz + 1) | j, jz 1i .

(A.4)

The first case to consider is the effect of Hex on the neutral exciton X 0 . As
stated in Sec. 1.2.2, there are four degenerate X 0 states, two of them are opticallybright states with Mz = Sz + Jz = 1 and two are dark states with Mz = 2
as shown in the leftmost diagram of Fig. A.1. In the basis of these four states
| +1i , | 1i , | +2i , | 2i, Hex can be expressed in matrix form as

Hex

b
=

0
0

0 d

0 d

190

(A.5)

A Quantum Dot Fine Structure


where
9
3
= (az + bz ),
4
4
3
b = (by bx ),
8
3
d = (by + bx ).
8

(A.6)
(A.7)
(A.8)

Thus, Hex energetically separates the bright and dark subspaces by 2 and mixes
the states within each subset. There is no mixing between the bright and dark
states because Eq. A.1 doesnt contain any terms like Jx Sz or Jz Sx . The bright
exciton subspace of Hex can be diagonalized to obtain energies
Eb, = b

(A.9)

1
| ib = (| +1i | 1i).
2

(A.10)

and eigenstates

In the same way, diagonalization of the dark subspace results in new dark exciton
states
1
| id = (| +2i | 2i)
2

(A.11)

with energies Ed, = d .


The exchange interaction has several important consequences for neutral excitons. Firstly, the new eigenstates | ib no longer have transitions which emit
circularly polarized photons, but rather linearly polarized photons x and y .
191

A Quantum Dot Fine Structure

No Exchange

Cylindrical
-1]

-1]

2]

+1]

-]b

2db
+]d

-]d

-]d

2dd

0]

+]b

+1]
2D

+]d
-

Elliptical

0]

0]

Figure A.1: X 0 level structure for a QD without exchange interaction, a cylindrically symmetric QD, and a slightly elliptical QD. The optical transitions are
denoted with arrows as well as polarization labels. In addition, the level splittings
and eigenstates are labelled.
Secondly, the non-degeneracy of the bright excitons destroys the polarization entanglement that could be generated by the biexciton-exciton cascade as discussed
in Chapter 1. However, it is important to note that the degeneracy (and the photon polarization) is restored if the QD has cylindrical symmetry (bx = by ). Figure
A.1 shows the X 0 quantum levels for a cylindrically symmetric QD (middle diagram) and an elliptical QD (rightmost diagram). Thus, the splitting of the bright
states is the result of the combination of an ellipticity in the QD shape and the
exchange interaction.
The next case to consider is that of the trion states X + and X . In the case

192

A Quantum Dot Fine Structure


of X , the two electrons in the s-shell must be in a spin-singlet state with Sz = 0.
This means there is no exchange interaction with the hole and the degeneracy
of the optical transitions remain. Clearly, the same argument must hold true for
X + except that the holes are in the spin-singlet with Jz = 0. This argument is
not valid if the carriers are weakly confined in the QD. In that case, the Coulomb
interaction is comparable to the confinement energy and the s-shell and p-shell
will strongly mix, leading to spatial configurations where the two electrons (two
holes) of the X (X + ) are separated. This will cause the local spin density to be
non-zero and the exchange interaction will again play a role [187].

193

Appendix B
g (2)( ): A Semiclassical-Langevin
Approach
To calculate g (2) ( ) for a cavity mode during the onset of lasing, one must
go beyond the rate equation analysis of Chap. 2, but not much further. This
type of calculation requires the use of stochastic methods, which in practice is
incorporated by adding a Langevin noise source. The calculation presented here
proceeds similar to the work in Ref. [189191]. With the noise source due to
spontaneous emission, Fsp , added, the rate equations for the intra-cavity photon

194

B g (2) ( ): A Semiclassical-Langevin Approach


number, P , and the number of excited carriers, N ,become
N
+ Fsp (t)
P = (G )P +
sp

(B.1)

N
GP,
N = Rp
sp

(B.2)

where is the spontaneous emission coupling factor, sp is the spontaneous emission rate, G is the gain, is the cavity field decay rate, and Rp is the pump rate.
These equations can be solved under steady-state to obtain P as a function of
Rp and the resulting curves with different values are shown in Fig. B.1. The
increase of leads to the soft turn-on behavior characteristic of microlasers. To
go beyond steady-state and into the stochastic dynamics of the system, a Markovian approximation (the system has no memory) is made such that Fsp (t) has the
following properties:

where Dpp =

NP,
sp o o

hFsp (t)i = 0

(B.3)

hFsp (t + )Fsp (t)i = 2Dsp ( ), ,

(B.4)

a function of the steady-state values of N and P . Since the

region of interest is below the onset of lasing, a linear gain model is sufficient and
can be expressed as
G=

a vg
(N Ntr ) ,
Vef f

195

(B.5)

B g (2) ( ): A Semiclassical-Langevin Approach

0.1

0.001
10-5
10-7
10-9
0.01

0.1

1
R pRth

=0.85
=0.085
=0.0085
=0.00085
=0.000085

10

100

Figure B.1: Plots of the steady state solution of Eq. B.1 and B.2 for P as a function
of normalized pump rate r = Rp /Rth for = 0.85, 8.5 102 , 8.5 103 , 8.5
104 , 8.5 105 . All other parameters are set to values found in Chap. 2, namely:
a = 4.5 1010 cm2 , vg = 101 0 cm/s, Vef f = 2.17 1014 cm3 , sp = 140 ps,
= 244 GHz, and Ntr = .048. The normalized pump rate is Rp divided by the
standard threshold pump rate which occurs when stimulated emission surpasses
cavity loss.

196

B g (2) ( ): A Semiclassical-Langevin Approach


where a is the differential gain, vg is the group velocity, Ntr is the carrier transparency number, and Vef f is the effective mode volume. Since the noise characteristics are of interest, Eq. B.1 and B.2 can be linearized by defining
P (t) = Po + p(t)

(B.6)

N (t) = No + n(t),

(B.7)

G = Go + G0 n(t)

(B.8)

G
.
N

The linearized rate equations

where No and Po are stable values and G0 =

for p and n are obtained by substituting Eq. B.6-B.8 into Eq. B.1 and B.2 and
retaining terms up to linear order. This results in the following:
p = (Go )p + G0 Po n +

n =

n
+ Fsp (t)
sp

1
0
+ G Po n Go p.
sp

(B.9)
(B.10)

Before proceeding with solving these equations, it is important to consider exactly


what is needed to calculate g (2) ( ). By definition,
hE () (t)E () (t + )E (+) (t + )E (+) (t)i
.
g ( ) =
hE () (t)E (+) (t)ihE () (t + )E (+) (t + )i
(2)

(B.11)

The intra-cavity electric field is proportional to the photon number by


E (+) (t) =

p
P (t)e(t)+c t ,

197

(B.12)

B g (2) ( ): A Semiclassical-Langevin Approach


and since the field is stationary (hP (t)i = hP (t + )i = Po ), g (2) ( ) can be written
as
g (2) ( ) = 1 +

hp(t + )p(t)i
,
Po2

(B.13)

where the terms linear in hp(t)i have been averaged out. Because Eq. B.9 and
B.10 are linear, it is auspicious to work in frequency space. Thus, the quantity of
interest, hp(t + )p(t)i, can be expressed more usefully as
1
hp(t + )p(t)i =
2

Z
dd 0 e h
p()
p ( 0 )i.

(B.14)

Eq. B.9 and B.10 can be solved by Fourier transform to obtain an expression for
p(),
p() =

1
sp

+ G0 Po

( )( 1sp + G0 Po ) + Go ( sp

1
)
sp

Fsp ().

(B.15)

For simplicity, this can be written as p() = H()Fsp () and using the Fourier
transform of Eq. B.3, Eq. B.14 can be expressed as
Dsp
hp(t + )p(t)i =

Z
d e |H()|2 .

(B.16)

The integral in Eq. B.16 can be done most easily using contour integration in the
complex plane. Since g (2) ( ) must be symmetric in time, the integral only has
to be solved for > 0, which corresponds to a contour in the lower half plane.
|H()|2 has resonances at
= r r
198

(B.17)

B g (2) ( ): A Semiclassical-Langevin Approach


and their complex conjugates, where
1
r = (n Go + )
2
1
n = + G0 Po

s
Go
( 1) + n .
r = r2 +
sp

(B.18)
(B.19)
(B.20)

Using the laser parameters from Chap. 2, the poles within a contour in the lower
half plane are = r r for all possible values of the pump rate. For these
types of lasers, r r as shown in Fig. B.2. Physically, this means that the
relaxation oscillations created by spontaneous emission events are very strongly
damped over the entire working range of the laser. It is these oscillations which
give rise to the bunching peak observed in the experiments, and the decay of
these oscillations sets the timescale of the correlation. Note, this is in direct
contradiction to the work of Ref. [88], which states that this timescale is set by
the coherence length of the light. This must not be the case because the coherence
length is mostly determined by phase noise, whereas g (2) ( ) is a phase-insensitive
measure (it does not depend on (t)).
By using the residue theorem, hp(t + )p(t)i is readily calculated to give the

199

B g (2) ( ): A Semiclassical-Langevin Approach

1.5

rr

1.4
1.3
1.2
1.1
1.0
0.01

0.1
1
10
Normalized Pump Rate

100

Figure B.2: A plot of r /r as a function of normalized pump rate from below to


above threshold. All laser parameters are the same as Fig. B.1 and = 0.85.

200

gH2LHL

B g (2) ( ): A Semiclassical-Langevin Approach

2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.0

r=0.1
r=0.18
r=0.32
r=0.56
r=1.
r=1.78
r=3.16
r=5.62
r=10.

0.5

1.0

1.5

2.0

2.5

3.0

Figure B.3: A plot of g (2) ( ) as a function of for several values of the normalized
pump rate r = Rp /Rth . All other parameters are set to that of Fig. B.1 with
= 0.85.
desired expression for g (2) ( )

er t
Dsp
r (n2 + r2 + r2 ) cos r t
g ( ) = 1 +
2P02 r r (r2 + r2 )

2
2
2
+ r (n r r ) sin r t) .
(2)

(B.21)

This result is plotted in Fig. B.3 for several pump rates. As one can see, for low
pump rates the relaxation oscillations decay quickly, but achieve g (2) (0) = 2, the
value of a thermal state. As the pump rate increases, the oscillations decay more
slowly and then gzero begins to decrease as towards 1, the value of a coherent
state. Using Eq. B.21, g (2) (0) can also be calculated as a function of pump rate, as
shown in Fig. B.4 for several values of . As expected from Fig. B.1, the transition
201

B g (2) ( ): A Semiclassical-Langevin Approach

2.0

=0.85
=0.085
=0.0085
=0.00085

gH2LH0L

1.8
1.6
1.4
1.2
1.0
0.01

0.1

1
R pRth

10

100

Figure B.4: A plot of g (2) (0) as a function of r = Rp /Rth for several values of .
All other parameters are set to that of Fig. B.1. As increases, the transition to
a coherent takes place over a much large range of pump rates.
to a coherent state above threshold takes place much slower for high lasers. The
results shown in Fig. B.4 differ from the experiments of Chaps. 2 and 3 because
of the fact that these measurements are limited by the timing resolution of the
detectors. This has several effects. Firstly, the oscillations found in Fig. B.3 are
averaged out, leaving only the decay. Secondly, because the timescale of the decay
r changes with the pump rate, values of g (2) (0) far below threshold are completely
averaged over. As r decreases near threshold, g (2) (0) becomes measurably larger
than 1, but the value is reduced due to averaging. While this is true and explains
some of the measurements, it is important to note that a full quantum calculation
of g (2) (0) shows that for high lasers, g (2) (0) < 2 for pump rates far below

202

B g (2) ( ): A Semiclassical-Langevin Approach


threshold. This will be calculated in App. C.

203

Appendix C
g (2)(0): A Quantum Density
Matrix Approach
In the last chapter, a semiclassical laser model was developed using Langevin
forces to determine g (2) ( ), more specifically it was used to determine the time dependence of g (2) ( ). Of course, g (2) (0) could be determined from this calculation,
but the low pump rate limit of g (2) (0) did not match the experimental data found
in Chap. 3. Presumably, this occurs because the quantum mechanical nature of
the electric field must become apparent at very low photon numbers. High
lasers are an interesting class of lasers because spontaneous emission events play
a dramatic part and become important at very low pump rates. Hence, it seems
reasonable that a quantum description is necessary when trying to understand the

204

C g (2) (0): A Quantum Density Matrix Approach


photon statistics of the lasing transition for these lasers. While a full quantum description including the time dependence is outside of the scope of this dissertation,
a quantum calculation of g (2) (0) is presented in this chapter.
The approach used is a density matrix formulation and closely follows Chapter
11 of Ref. [12]. The main differences are the level scheme and the added calculation
of hn2 i. The approximate level scheme for this lasing model is shown in Fig. C.1
and is comprised of four levels; the conduction and valence bands denoted as
| CBi and | V Bi and two lasing levels, | ai and | bi. Also included in the figure
are the pump rate from | V Bi to | CBi, the decay rate from | ai into modes other
than the lasing mode and the relaxation rate from (into) the bands a (b ). The
starting point is the interaction term of the Hamiltonian for several atomic systems
interacting with a single cavity field
H=

i
i
~g(+
a + a
)=

Hi ,

(C.1)

where a (a ) are annihilation (creation) operators for photons in the cavity field
i
i
and +
= | ai hb| (
= | bi ha|) are the raising (lowering) operators for the ith

atom. The reduced density matrix for the cavity field can be expressed as

n,n0 = Tratoms [H, ]n,n0 + (L)n,n0


~

(C.2)

where Tratoms denotes a partial trace over the atomic subspace and (L)n,n0 is the
damping of the cavity field due to its finite Q and is characterized by a loss rate
205

C g (2) (0): A Quantum Density Matrix Approach

CB]
Ga
a]
Rp

H
g
Gb

b]

VB]

Figure C.1: The level scheme used in the quantum density matrix approach. The
scheme includes the conduction and valence bands as well as two lasing levels. Also
included are other phenomenological rates such as the pump rate and various level
decays.

206

C g (2) (0): A Quantum Density Matrix Approach


. The gain part of this expression can be simplified using Eq. C.1 to obtain
( n,n0 )gain

Xn
Han,bn+1 ibn+1,an0 ian,bn0 +1 Hbn0 +1,an0
=
~ i
o
Hbn,an1 ian1,bn0 ibn,an0 1 Han0 1,bn0 ,

(C.3)

where

Han,bn+1 = ~g n + 1

(C.4)

Hbn0 +1,an0 = ~g n0 + 1.

(C.5)

Since the density matrix for the cavity field is of interest, each individual
atomic contribution is collected into a generalized population matrix such that
an,bn+1 =

ian,bn+1 .

(C.6)

Physically, this does not correspond to letting g N g because not all of the
atoms are excited simultaneously. This procedure instead results in an effective
coupling for the collective action of the atoms (or quantum dots). Thus, the sum
can be removed from Eq. C.3, and the effective coupling g will be understood.
This leaves,
n
Han,bn+1 bn+1,an0 an,bn0 +1 Hbn0 +1,an0
~
o
Hbn,an1 an1,bn0 bn,an0 1 Han0 1,bn0 ,

( n,n0 )gain =

as the expression for the gain part of n,n0 .


207

(C.7)

C g (2) (0): A Quantum Density Matrix Approach


The terms i,j in Eq. C.7 each have their own equation of motion governed by
the Schrodinger equation, phenomenological loss mechanisms and the pump rate.
In this work, the important loss mechanisms included are spontaneous emission
from |ai into modes other than the cavity mode and atomic dephasing . For
the sake of notational simplicity, the decay rate from |bi into |V Bi will be taken
as b = . The equations of motion can be written as
an,an0 = a CBn,CBn0 an,an0

(Han,bn+1 bn+1,an0 an,bn0 +1 Hbn0 +1,an0 ) ,


~

(C.8)

an,bn0 +1 = an,bn0 +1

(Han,bn+1 bn+1,bn0 +1 an,an0 Han0 ,bn0 +1 ) ,


~

(C.9)

bn+1,an0 = bn+1,an0

(Hbn+1,an an,an0 bn+1,bn0 +1 Hbn0 +1,an0 ) ,


~

(C.10)

bn+1,bn0 +1 = bn+1,bn0 +1

(Hbn+1,an an,bn0 +1 bn+1,an0 Han0 ,bn0 +1 ) ,


~

(C.11)

CBn,CBn0 = Rp V Bn,V Bn0 a CBn,CBn0 ,

(C.12)

V Bn,V Bn0 = Rp V Bn,V Bn0 + (an,an0 + bn,bn0 ) .

(C.13)

The goal is to solve Eqns. C.8-C.13 for an,bn0 +1 and bn+1,an0 in terms of n,n0 . First,
in the steady state Eq. C.12 yields Rp V Bn,V Bn0 = a CBn,CBn0 , so CBn,CBn0 can

208

C g (2) (0): A Quantum Density Matrix Approach


be removed from Eq. C.8. Then, V Bn,V Bn0 can be written in terms of n,n0 by
noting that by definition,
n,n0 = CBn,CBn0 + an,an0 + bn,bn0 + V Bn,V Bn0 .
This can be simplified by noting that CBn,CBn0 =

Rp
0

a V Bn,V Bn

(C.14)

and assuming that

a is the fastest rate in the problem. This means CBn,CBn0 0 because once an
electron is excited into | CBi, it immediately decays into | ai. Thus,
n,n0 V Bn,V Bn0 = an,an0 + bn,bn0 .

(C.15)

Eq. C.13 can be solved in the steady-state, to obtain


Rp
V Bn,V Bn0 = an,an0 + bn,bn0 .

(C.16)

If these equations are used in combination, V Bn,V Bn0 can be written as


V Bn,V Bn0 =

n,n0 .
+ Rp

(C.17)

Once Eq. C.17 is substituted in Eq. C.8, Eq. C.8-C.11 form a closed set with
external drive term rp n,n0 where the effective pump rate rp = Rp /(Rp + ). This
closed set can be solved using linear algebra if written as
R = M R + A,

209

(C.18)

C g (2) (0): A Quantum Density Matrix Approach


where

an,an0

an,bn0 +1

,
R=

0
bn+1,an

bn+1,bn0 +1
and

g n0 + 1

M =

g n + 1

g n0 + 1

g n + 1

1


0

A = rp n,n0
,
0



0

g n + 1

(C.19)

0
g n + 1

g n0 + 1

0
g n + 1

(C.20)

The steady-state solution for R is given by


R = M 1 A.

(C.21)

The resulting expressions for an,bn0 +1 and bn+1,an0 are


an,bn0 +1
bn+1,an0

rp g n0 + 1 2 0
g (n n) + 2
=
|M |

rp g n + 1 2
g (n n0 ) + 2 ,
=
|M |

(C.22)
(C.23)

where
|M | = 4 + 2g 2 2 (2 + n + n0 ) + g 4 (n n0 )2 .

(C.24)

The expressions for an1,bn0 and bn,an0 1 can be obtained from Eq. C.22 and C.23
by substituting n n 1 and n0 n0 1.
210

C g (2) (0): A Quantum Density Matrix Approach


It is worth noting that the calculation of g (2) (0) requires only knowledge of
the diagonal elements (n) = n,n of n,n0 . This fact simplifies the expressions
considerably and henceforth n0 = n. Now, using Eq. C.22 and C.23, Eq. C.7 can
be written as,

(n)
gain

(n + 1)A
nA
=
(n) +
(n 1),
1 + (n + 1)B/A
1 + nB/A

(C.25)

where
A=

2g 2
rp

(C.26)

4g 2
A

(C.27)

is Scullys linear gain coefficient and


B=

is Scullys self-saturation coefficient. If the loss term (L)n,n is added, the full
expression for equation of motion is

(n + 1)A
nA
(n)

=
(n) +
(n 1)
1 + (n + 1)B/A
1 + nB/A
n (n) + (n + 1)(n + 1).

(C.28)

In the steady-state, Eq. C.28 can be reduced by requiring detailed balance to


two expressions

(n + 1)A
0=
(n) + (n + 1)(n + 1)
1 + (n + 1)B/A

nA
0=
(n 1) n (n)
1 + nB/A

211

(C.29)
(C.30)

C g (2) (0): A Quantum Density Matrix Approach


that are clearly equivalent. Each sets the gain from a lower photon number state
equal to the loss from the higher number state. Solving Eq. C.30 yields a recursive
equation for (n),

A/
(n) =
(n 1).
1 + nB/A

(C.31)

The solution to this recursion is clearly


(n) = (0)

n
Y
j=1

and (0) can be determined from


usefully as

P
n=0

A/
,
1 + jB/A

(C.32)

(n) = 1. Eq. C.32 can be expressed more

A A2 n
! B
B
.
(n) = (0)
!
n+ A
B

(C.33)

This enables (0) to be expressed in closed form as


2 AB
A2
e B A
B
A A2 ,
(0) = A A

,
B
B
B B

(C.34)


R
is the Euler Gamma function defined as (x) = 0 dt et tx1 and
where A
B

R
A2
,
is the plica function defined as (x, y) = y dt et tx1 .
A
B B
The photon number distribution can then be plotted using Eq. C.33 and C.34
as shown in Fig. C.2 for a .005 laser for five values of the pump rate. For
pump rates much less than threshold the distribution is thermal, which is a characterized by an exponentially decaying dependence on n. Above threshold, the
distribution approaches Poissonian, which is the distribution for a coherent state.
212

C g (2) (0): A Quantum Density Matrix Approach

r p= 0.4rth
r p= 0.6rth
r p= 1.rth
r p= 1.8rth
r p= 3.2rth

HnL

0.1
0.01
0.001
10-4
10-5
0

50

100

150
n

200

250

300

Figure C.2: A plot of (n) for five values of the pump rate: rp =
{10.5 , 10.25 , 1, 10.25 , 10.5 } rth where rth is the threshold pump rate. These
curves are for a .005 laser.
Thus, Eq. C.33 reproduces what is expected; the photon statistics of the cavity
field change from thermal to Poissonian as the threshold is reached.
To obtain the more traditional laser quantity of output power vs input power,
hni must be calculated. This can be accomplished from Eq. C.33 and the definition

213

C g (2) (0): A Quantum Density Matrix Approach


of hni in the following way,
hni =

n(n)

n=0

= (0)

A A2 n
! B
B

)!
(n + A
B

A2 n
A
A
X

n
+

B
B
B
A
!
= (0)
A
B n=1
(n + B )!
2 n1
2 n
A
A
X
X

2
B
B
A
A
A
A
(0)
(0)
!
!
=
A
B
B n=1 (n 1 + B )! B
B n=1 (n + A
)!
B
n=1

A2 A
[1 (0)] ,
B B

(C.35)

leaving a nice, closed expression. Eq. C.35 is plotted in Fig. C.3 for seven different
values of as a function of threshold-normalized pump rate rp /rth . The onset
of lasing is clearly shown to broaden as increases, just like for the simple rate
equation model used in Chap. 2 and App. B. Within this density-matrix approach,
can be defined by comparing the Purcell-enhanced spontaneous rate into the
cavity mode 4g 2 / to the total emission rate from | ai,
=

1
4g 2 /
=
.
2
4g / +
1 + 4g
2

(C.36)

Using Eq. C.33, g (2) (0) can also be calculated. From the definition,
(2)

g (0) =


a a aa

214

ha ai2

(C.37)

C g (2) (0): A Quantum Density Matrix Approach

104

Xn\

100

0.01
0.01

1
r prth

0.1

10

100

Figure C.3: A plot of hni for seven different values of = {5104 , 5103.5 , 5
103 , 5 102.5 , 5 102 , 5 101.5 , 5 101 }. The curve with the largest (smallest) kink near threshold corresponds to = 5 104 (5 101 ). Each curve is
plotted as a function of threshold-normalized pump rate rp /rth .

215

C g (2) (0): A Quantum Density Matrix Approach


Using the commutation relation [a, a ] = 1, and making the identifications n = a a
n2 = a aa a, g (2) (0) can be expressed as
g (2) (0) =

hn2 i hni
.
hni2

(C.38)

Thus, an expression for hn2 i must also be obtained. Similar to the derivation of
hni, hn2 i can be calculated from the definition as follows,
2

hn i =

n2 (n)

n=0

A2
X

B
A A
A

!
n n+
= (0)
B n=1
B
B n+ A
!
B
2 n
2 n
A
A
X
X

B
B
A
A
A
(0)
!
n
!
n
= (0)
A
B n=1 (n 1 + B )! B
B n=1 (n + A
)!
B
2 n1
A
X

2
B
A
A
(0)
!
(n 1)
=
B
B n=1
)!
(n 1 + A
B
2 n1
A
X

2
B
A
A
A
hni
(0)
!
+
A
B
B n=1 (n 1 + B )! B
2

A
A
A2
+

hni,
(C.39)
=
B
B B
yielding again a nice, closed form expression. Now, Eq. C.38 is plotted in Fig. C.4
for the same seven values as Fig. C.3. At very low pump rates, the photon
statistics are bunched demonstrating a thermal-type distribution. As the pump
rate increases, the system undergoes the lasing transition and the photon statistics
show an uncorrelated Poissonian distribution with g (2) (0) = 1. Furthermore, as
216

C g (2) (0): A Quantum Density Matrix Approach

2.0
=5E-4
=5E-3.5
=5E-3.
=5E-2.5
=5E-2.
=5E-1.5
=5E-1.

gH2LH0L

1.8
1.6
1.4
1.2
1.0
0.01

0.1

1
r prth

10

100

Figure C.4: A plot of g (2) (0) for seven different values of = {5 104 , 5
103.5 , 5 103 , 5 102.5 , 5 102 , 5 101.5 , 5 101 }. Again, each curve is
plotted as a function of threshold-normalized pump rate rp /rth .
is increased, not only does the transition to a coherent state broaden, but the low
pump rate asymptotic value decreases, just like what was measured in Chap. 3.

217

Appendix D
Reconvolution
Due to the Purcell Effect as explained in Chap. 1, the spontaneous emission
rate of a quantum dot transition can be drastically modified by the presence of a
cavity. Experimentally, this effect is confirmed by measuring the decay rate of a
QD transition as a function of spectral detuning from the cavity mode. There are
several examples of this measurement in this dissertation and it is performed in
the Bouwmeester Lab using a technique called time-correlated photon-counting
(TCPC). Basically, a fast electronic device called a time-to-amplitude converter
(TAC) receives a trigger signal from the ultrafast laser (in our case a SpectraPhysics Tsunami with 100 fs pulse width and repetition rate of 80 MHz) used
the excite the QD. This pulse acts as the start signal for the TAC. It then waits
for an electronic pulse from the single photon detector (a Si Avalanche Photo-

218

D Reconvolution
diode (APD)) after it detects a photon which is the stop event. This detector
measures the photons emitted by the QD. The TAC then makes a histogram of
these start-stop events with a user selected bin size, which is usually 68.8 ps for
our measurements. The result of this measurement then shows the initial increase
to the excited state of the QD and the subsequent decay. Provided that the decay of the QD is much slower than the instrument response and bin size, the
measurement accurately measures the decay rate. In our setup, the instrument
response time is dominated by the timing jitter of the APD, which has a timescale
of approximately 200 ps. The width of the excitation pulse and other electronic
errors are much faster and can be neglected. Importantly, the bulk spontaneous
emission lifetime of a QD is typically near 1 ns [26] and a Purcell enhanced lifetime
of 200 ps or shorter is perfectly within reason. Thus, for decay times around 500
ps, the effect of the instrument response function is significant and must be taken
into account. The point of this appendix is to clearly describe the process used
to determine fast lifetimes.
Fig. D.1 is a plot of both the instrument response (in black) of a TCPC measurement and the data from a QD on resonance with a cavity mode (in red). The
QD emission is almost as fast as the instrument response, so simply fitting the
decay to an exponential (or in this case a bi-exponential) would underestimate the
decay rate. The instrument response data is taken by detecting the fast laser light

219

D Reconvolution

Counts

0.1
0.01
0.001
10-4
0

2
3
time HnsL

Figure D.1: TCPC raw data for the instrument response in black and emission
from a QD on resonance with a cavity mode in red.
that is reflected from the sample surface. Since the pulse width is much shorter
than the APD timing jitter, this provides an adequate approximation to a deltafunction like signal so that the instrument response data is a good measure of the
APD response. In the limit of no noise, the actual QD decay could be deconvolved
from the measured signal by dividing the Fourier transform of the measured signal
by the Fourier transform of the instrument response and then Fourier transforming back to the time domain. However, in practice the high-frequency noise of
the measurement will dominate the transform and the result will look nothing
like what is wanted. The solution to this problem involves spectral filtering and
many people have devoted significant portions of their scientific careers to opti-

220

D Reconvolution
mally solving it [192]. I have tried using this technique in the past with TCPC
data without much success, probably due to the fact that the measured data and
instrument response are virtually the same. Instead, I have found that the best
technique to solving this problem for TCPC data is by reconvolution. The basic idea of this technique to fit the measured data using a function which is the
convolution of the instrument response data with the form of the actual decay.
This procedure works as follows. First, the form of the actual decay is taken
to be a bi-exponential expressed as

Dactual (t0 , A, B, f , s ; t) = A(t to ) e(tto )/f + Be(tto )/s ,

(D.1)

where (tto ) is the Heaviside step function, f and s are the fast and slow decay
rates, B sets the relative weighting the slow decay, A sets the overall maximum
value, and to is the start of the decay. This function is convolved with the instrument response function, with the added difficulty that the instrument response is
not a function, but a set of N data points IR [t] = {(t1 , y1 ), (t2 , y2 ), ..., (tN , yN )}
with t = tn tn1 . Thus, the usual integral turns into a sum which can be
expressed as
Dexp (t0 , A, B, f , s ; t) = t

N
X

Dactual (t0 , A, B, f , s ; t tj )IR [tj ],

(D.2)

j=1

and so the function to be used to fit the experimental data is actually composed
of a sum of N bi-exponential functions each weighted by the instrument response.
221

D Reconvolution
The next step is to determine the initial values for the five fitting parameters
{to , A, B, f , s }. This can be done simply by plotting the function with some
parameters and seeing how well that fits. The quality of a fit for a given set of
fitting parameters to a set of N data points dat[t] = {(t1 , y1 ), (t2 , y2 ), ..., (tN , yN )}
is determined by 2 , where
2

(t0 , A, B, f , s ) =

N
X

{ln (dat[tj ]) ln [Dexp (to , A, B, f , s ; tj )]}2 ,

(D.3)

j=1

and ln is used because the fit should be optimized on a log scale. For all of
the fitting parameters except to , 2 is very nearly quadratic in the neighborhood
of the optimal value of the parameter as shown in Fig. D.2 for the parameter
f . Since these parameters behave nicely, they can be easily optimized. to has
to be handled individually by minimizing the 2 for a specific set of the other
parameters, although a reasonable value of to is easy to find.
Because there must exist some optimal fit which minimizes 2 , it seems reasonable to surmise that 2 (to , A, B, f , s ) must be able to be expressed as
4
4
4
X
1 XX
M x x +
V x + C,
(to , A, B, f , s ) =
2 =1 =1
=1
2

(D.4)

a general quadratic function of the fit parameters x = {A, B, f , s }, where M is


a 44 matrix and V is a vector. The minimum of this function will take place
where
4

2 X
=
M x + V = 0
x
=1
222

(D.5)

D Reconvolution

0.20

0.18
0.16
0.14
0.12
0.200

0.205

0.210
f

0.215

0.220

Figure D.2: 2 as a function of the parameter f for the QD data in Fig. D.1
keeping all other parameters constant with quadratic fit.
Thus, if M and V are known, then the vector of best fit parameters xo is
xo = M1 V.

(D.6)

M can be determined from a set of initial guess parameters xi by taking the


second derivative about those points because
2 2
= M .
x x

(D.7)

V can also be determined from xi by


4

2 X

M xi
V =
x =1

(D.8)

Because the 2 function is hard to work with directly, it is more convenient to

223

D Reconvolution
take numerical first and second derivatives defined by
o
f
1n
f (x + /2) f (x /2) ,

2 n

o
2f
1
f
(x
+

2f
(x
)

+
f
(x

)
,

x2

2f
1 n
f (x + /2, x + /2) f (x /2, x + /2)

x x

o
f (x + /2, x /2) + f (x /2, x /2) .

(D.9)
(D.10)

(D.11)

Then, in terms of xi , xo is
xo = M1

2
+ xi ,
x

(D.12)

and they will be equal once 2 /x vanishes. The only iterative part required
in setting up a fitting algorithm is to set the value of . For the numerical
derivatives to approximate the real derivatives well,
|xi xo |.

(D.13)

This can be written in terms of the desired precision of the fit P as P xo .


In practice, a factor of 100 is sufficient and so
= P xo /100.

(D.14)

In summary, the process for reconvolution is as follows:


1. Perform a discrete convolution of the instrument response data with Eq. D.1.
224

D Reconvolution
2. Find good initial guess fitting parameters by playing around with the function resulting from Step 1.
3. Find a good value for to by noting where the actual decay begins.
4. Take numerical first and second derivatives of the 2 function (Eq. D.3)about
the initial guess parameters. The initial size of the s should be determined
by the required precision using Eq. D.14. In practice an initial value of
P = 0.01 is sufficient.
5. From the derivatives, determine new fit parameters using Eq. D.12.
6. Check to see if 2 has been lowered using the new parameters. If so, go back
to Step 4. If not, reduce by a factor of 10 and go back to Step 4 unless
a sufficient final precision has been reached. This means the algorithm has
succeeded and the best parameters have been found.
Using the aforementioned technique on the data from Fig. D.1 results in the
fit shown in Fig. D.3. The fit is very good for t > 0, and the poor fit at earlier
times is due to the lack of a rise timescale in Eq. D.1. This can be easily fixed
in future implementations. This technique was used to extract the fast lifetimes
found in the measurements of Chap. 2, 7, 9, and 10.

225

D Reconvolution

Counts

1.000
0.500
0.100
0.050
0.010
0.005
0

2
time HnsL

Figure D.3: Lifetime plot of the QD data from Fig. D.1 with fit using the
reconvolution technique. The fit best fit parameters are {to , A, B, f , s } =
{0.10, 6.043, 0.02375, 0.2094, 0.9870} with the times in units of ns.

226

Appendix E
Using the Cryostat
This appendix should serve as a manual for the unloading/loading of samples
into the Oxford Instruments Microstat He cryostat as well as for the proper cool
down procedure to begin measurements.

E.1

Sample Unloading

This section of the appendix describes the procedure to properly remove a


sample from the cryostat.
1. Remove the cold finger of the He transfer line from the cryostat and store
on the hooks located on the wall next to the He dewar.
2. Remove all connections to the cryostat such as the temperature sensor and
227

E Using the Cryostat


BNC cables.
3. Close the valve that connects the cryostat to the hose from the turbo pump.
This is very important because venting the hose with the valve open may
lead to vacuum grease on the sample or cryostat window.
4. Vent the hose from the turbo pump and cap the end with the copper seal.
5. Vent the cryostat to atmosphere by slowly opening the vacuum valve.
6. Remove the connector that seals the cryostat to the outer cover and remove
the cryostat from the cover.
7. Take the cryostat over to the lab bench and unscrew the screw that attaches
the radiation shield to the cryostat.
8. Very carefully remove the radiation shield. This must be done very carefully
because the temperature sensor has two weak solder joints. It is not fun to
have to re-solder them, so keep that in mind when removing the radiation
shield.
9. Remove the sample using tweezers. If the sample is electrically connected
to the cryostat, desolder the wires either at the sample chip carrier or the
gold contact pins.

228

E Using the Cryostat


10. Remove the left-over vacuum grease that was used to mount the sample
to the cryostat. I have found that first wiping it the excess away with a
kimwipe and then using isopropanol to remove the rest works best. Be sure
the isopropanol is cleaned off to leave a nice surface for the next sample.

E.2

Sample Loading

This section describes the procedure to properly load a sample onto the cryostat for measurement. To mount an electrically connected sample, follow all steps.
For non-electrical samples, do steps 1, 5, 6, 10, 11, and 14.
1. Be sure the end of the cryostat is nice and clean. Clean with isopropanol if
necessary.
2. Cut three pieces of wire from the Cu magnet wire spool that are long enough
to go from the sample to the gold contact pins. These wires will be the leads
for V 1, V 2, and GN D.
3. Because the magnet wire has a coating of insulating varnish, the leads need
to have this varnish removed at the tips. This can be done by scraping
roughly 5 mm off at the end using an exacto-knife. Take care not to cut
the wire by using too much force. It should be more reflective at the places

229

E Using the Cryostat


where the varnish has been removed.
4. Solder the leads to the chip carrier containing the sample using the Kester
Gold solder and some flux. This is not that easy because the gold contacts
on the chip carrier are small, but it is certainly possible.
5. Place some vacuum grease on the sample area using the end of a Q-tip.
Then, using a straight edge, wipe across the sample area to leave a smooth,
thin layer of grease. The best sample temperatures have been achieved using
a very thin layer of grease.
6. Place the sample onto the grease. Using some tweezers, push the sample
down into the grease and slightly twist the sample back and forth. This
ensures that the back of the sample has an even coating of grease and that
the sample is thermally contacted to the sample mount.
7. Solder the leads on the chip carrier to the gold contact pins on the cryostat.
With the cryostat oriented such that the sample side is to the right and the
vacuum flange is to the left when viewing the sample, the gold contact pin
nearest you corresponds to V 1 as shown in Fig. E.1. This pin is connected
to the BNC connection that is on the front of the cryostat. That is, it
is the connection closest to you when viewing from the sample side. The
middle contact pin is GN D and the farthest pin is V 2, which connects to the
230

E Using the Cryostat

GND
V2
V1
Figure E.1: A diagram of the cryostat and the sample mount area. Leads are
labelled as V 1, V 2, and GN D. The sample is mounted in a chip carrier like used
in experiments.
BNC connection on the back side. These leads are labelled in the described
orientation in Fig. E.1.
8. Make sure that the electrical connections are good by checking the resistance
from the BNC connection to where the lead contacts the chip carrier. The
V 1 connection should have a resistance of 10 , V 2 should be 15 , and
GN D should be 5 .
9. Push the sample down into the grease again with tweezers to make sure the
sample hasnt moved while soldering the leads. The sample should look like
231

E Using the Cryostat

Figure E.2: A picture of a properly mounted micropillar sample with contacts.


The vacuum grease can be seen around the chip carrier.
the sample in Fig. E.2.
10. Carefully place the radiation shield back on the cryostat. The radiation
shield is not symmetric, so the side with front scratched into it should be
towards the sample side of the cryostat. The back side will be too close
to the sample and may disconnect the leads.
11. Put the cryostat back into the outer cover on the optical table.
12. Re-connect the temperature sensor and the vacuum hose from the turbo
pump.

232

E Using the Cryostat


13. Attach the BNC cable from the Source-meter to V 1 and V 2 and check that
each has the expected I V characteristic.
14. Start the turbo pump.

E.3

Cool-Down Procedure

This section describes the proper cool-down procedure. This procedure should
be done with 2 people for the first few attempts because working with pressurized
cryogenics can be dangerous.
1. Wait until the pressure reads less than 1.0105 mbar before beginning
cool-down.
2. Prepare the liquid He dewar for installation of the transfer line. At the top
exhaust port (where the transfer line goes), place the o-ring followed by the
gold spacer. This spacer compressed the o-ring and creates a tight seal with
the transfer line. Place the nut over both of these and screw down until it
comes into contact with the spacer. It should be more difficult to screw at
this point.
3. Open the valve for the top exhaust port and make sure the valves are closed
for both exhaust ports on the side. (One of these is the 3 psi safety port

233

E Using the Cryostat


and the other is a standard exhaust port.)
4. Make sure the black cap covering the entrance port for the cold finger on
the cryostat is removed.
5. Put the cryo-gloves on.
6. Pick the transfer line off of the hook using your dominant hand/arm to hold
the part that goes in the dewar and your weaker hand/arm to hold the coldfinger part. The transfer line is awkward to manipulate, but be careful not
to bend it too much.
7. Lift up the transfer line and start to push it down into the dewar. You may
need to use a stool to reach high enough. Do this slowly so as not to create
too much gaseous He and waste the liquid. As you do this, use the wrench
to further screw the nut down. Screw it down far enough that a lot of gas
doesnt escape but you can still push the transfer line in deeper. This is
quite a balancing act and someone should be there to help you until you are
comfortable doing it yourself.
8. Continue lowering the transfer line into the dewar while monitoring the
pressure. The pressure should reach 10 psi or so. If not, you may need to
start over. If the pressure reaches 15 psi, release some of it by opening the

234

E Using the Cryostat


valve on the side standard exhaust port (not the 3 psi safety port!). The 15
psi safety should go off if the pressure is too high, but some dewars go off for
smaller or higher pressures. In practice, try to keep the pressure below 15
psi. Make sure the nut is screwed down enough that pressure isnt escaping
through the top. Lower the transfer line until the distance between the top
of the transfer line and the nut on the dewar is about 10 in or so.
9. Remove the cover from the cold-finger and insert the cold-finger into the
cryostat. Make sure it goes in all of the way which may necessitate moving the dewar to access the correct angle. Tighten the cold-finger down
completely.
10. Push the transfer line about 5 more inches and completely tighten the nut
down.
11. Open the needle valve at the top of the transfer line by about an eighth of
a turn. This will begin the flow of liquid He through the transfer line. You
should be able to feel He gas exhaust from the transfer line exhaust hose.
12. Wait about 20 minutes and then check the sample temperature. You should
be able to reach a minimum temperature of 4.2 K, so the needle valve may
need to be opened a little more, although it shouldnt be opened by more
than a quarter to a third of a turn. Sometimes better temperatures can be
235

E Using the Cryostat


reached by slightly adjusting the position of the dewar, which strains the
transfer line. Best temperatures have been reached when the transfer line is
slightly compressed.
13. If the temperature starts to creep up during measurements, you may need
to lower the transfer line all of the way into the dewar.
14. Keep in mind that the last 15 L or so of liquid He is not accessible using
the transfer line because it is a bit too short to reach the bottom of a 100 L
dewar.
15. Though the temperature sensor may read 4.2 K, it takes about 20 mins to 1
hour for the sample to reach its base temperature depending on the quality
of the thermal contact. You may notice the sample spatially drifting with
time (a m/min or even faster) until it settles completely.

236

Appendix F
List of Samples
This is a list of all important samples on which most of the experiments in
this dissertation were performed. Of course, many more samples were measured
but these gave the best results. Each photonic crystal sample should be resting
in a container filled with isopropanol. Each container should be labelled (usually
with a date) on the removable lid. Each micropillar sample is labelled by its MBE
wafer piece. All samples should be found on the sample shelf in the lab with the
correct label. For the photonic crystal samples, each cavity has an alpha-numeric
label. For the micropillars, each block of 47 cavities is marked with a single
label as shown in Fig. F.1. The label for a specific cavity is given by the label for
the entire block followed by the row number and column number. For example,
cavity A4R2C3 is a cavity in block A4 at row 2, column 3. Rows and columns are

237

F List of Samples

R1

R2

R3

R4
C1
C2
C3
C4
C5
C6
C7

Figure F.1: An optical microscope image of a micropillar sample with Au contacts.


The 47 block of cavities is marked by an alpha-numeric label (A4 in this case)
which is circled in red. The rows (R) and columns (C) are also marked in red.
Also visible are n-type contact and the p-type contact.
counted up from the device nearest the block label as shown in Fig. F.1.

F.1

Photonic Crystal Samples

Sample Label
ueberlaser cav0406
opt.pos.
080608D
BB 8/3/08

Experiments
and
Notes
Lasing and g (2) ( )
Optical positioning

238

Good Cavities
Most of the cavities
lase
24, 23W, 23H, 23J

F List of Samples

F.2

Micropillar Samples

Sample Label
061206C #10

Notes
Bright SPS.

070213C #6, #10

Bright SPS.

070419C

PIN for EL but shows


nice charging X + to
X 2
Controlled charging
experiment X 0 to
X at V 18 V
HV Switching and
Stark Effect
Single Photon EL
Latest
Charging
sample has X 0 to
X at V 0.6 V

070516C #10

070315A
070124C #6
080608C

239

Good Cavities
See Stefans notebook
See Stefans notebook
NA

B4R2C3, B3R2C7

A2R2C1
B1R2C3
Many

Bibliography
[1] F. Schmidt-Kaler, H. Haffner, M. Riebe, S. Gulde, G. P. T. Lancaster,
T. Deuschle, C. Becher, C. F. Roos, J. Eschner, and R. Blatt, Nature (London) 422, 408 (2003).
[2] C. F. Roos, M. Riebe, H. Haffner, W. Hansel, J. Benhelm, G. P. T. Lancaster, C. Becher, F. Schmidt-Kaler, and R. Blatt, Science 304, 1478 (2004).
[3] M. D. Barrett, J. Chiaverini, T. Schaetz, J. Britton, W. M. Itano, J. D.
Jost, E. Knill, C. Langer, D. Leibfried, R. Ozeri, et al., Nature (London)
429, 737 (2004).
[4] J. Chiaverini, D. Leibfried, T. Schaetz, M. D. Barrett, R. B. Blakestad,
J. Britton, W. M. Itano, J. D. Jost, E. Knill, C. Langer, et al., Nature
(London) 432, 602 (2004).
[5] J. Chiaverini, J. Britton, D. Leibfried, E. Knill, M. D. Barrett, R. B.
Blakestad, W. M. Itano, J. D. Jost, C. Langer, R. Ozeri, et al., Science
308, 997 (2005).
[6] M. Brune, F. Schmidt-Kaler, A. Maali, J. Dreyer, E. Hagley, J. M. Raimond,
and S. Haroche, Physical Review Letters 76, 1800 (1996).
[7] P. R. Berman, Cavity Quantum Electrodynamics (Academic Press, San
Diego, 1994).
[8] H. Mabuchi and A. C. Doherty, Science 298, 1372 (2002).
[9] L. Childress, J. M. Taylor, A. S. S. rensen, and M. D. Lukin, Physical Review
Letters 96, 070504 (2006).
[10] M. Knill, R. Laflamme, and G. Milburn, Nature (London) 409, 46 (2001).
[11] S. D. Barrett and P. Kok, Physical Review A 71, 060310 (2005).

240

BIBLIOGRAPHY

BIBLIOGRAPHY

[12] M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University


Press, Cambridge, 1997).
[13] L. C. Andreani, G. Panzarini, and J.-M. Gerard, Physical Review B 60,
13276 (1999).
[14] C. W. Gardiner and M. J. Collett, Physical Review A 31, 3761 (1985).
[15] J. P. Reithmaier, G. Sek, A. Loffler, C. Hofmann, S. Kuhn, S. Reitzenstein,
L. V. Keldysh, V. D. Kulakovskii, T. L. Reinecke, and A. Forchel, Nature
(London) 432, 197 (2004).
[16] T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. M. Gibbs, G. Rupper,
C. Ell, O. B. Shchekin, and D. G. Deppe, Nature (London) 432, 200 (2004).
[17] E. M. Purcell, H. C. Torrey, and R. V. Pound, Physical Review 69, 37
(1946).
[18] P. Michler, A. Kiraz, C. Becher, W. V. Schoenfeld, P. M. Petroff, L. Zhang,
E. Hu, and A. Imamoglu, Science 290, 2282 (2000).
[19] E. Moreau, I. Robert, J. M. Gerard, I. Abram, L. Manin, and V. ThierryMieg, Applied Physics Letters 79, 2865 (2001).
[20] M. Bayer, T. L. Reinecke, F. Weidner, A. Larionov, A. McDonald, and
A. Forchel, Physical Review Letters 86, 3168 (2001).
[21] T. D. Happ, I. I. Tartakovskii, V. D. Kulakovskii, J.-P. Reithmaier,
M. Kamp, and A. Forchel, Physical Review B 66, 041303 (2002).
[22] K. Srinivasan and O. Painter, Nature (London) 450, 862 (2007).
[23] D. Englund, A. Faraon, I. Fushman, N. Stoltz, P. Petroff, and J. Vuckovic,
Nature (London) 450, 857 (2007).
[24] D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder,
U. Meirav, and M. A. Kastner, Nature (London) 391, 156 (1998).
[25] N. G. Stoltz, Ph.D. thesis, University of California Santa Barbara, Santa
Barbara, CA (2007).
[26] P. M. Petroff, A. Lorke, and A. Imamoglu, Physics Today 54, 050000 (2001).

241

BIBLIOGRAPHY

BIBLIOGRAPHY

[27] D. Leonard, M. Krishnamurthy, C. M. Reaves, S. P. Denbaars, and P. M.


Petroff, Applied Physics Letters 63 (1993).
[28] J. Y. Marzin, J. M. Gerard, A. Izrael, D. Barrier, and G. Bastard, Physical
Review Letters 73, 716 (1994).
[29] R. J. Warburton, B. T. Miller, C. S. D
urr, C. Bodefeld, K. Karrai, J. P. Kotthaus, G. Medeiros-Ribeiro, P. M. Petroff, and S. Huant, Physical Review
B 58, 16221 (1998).
[30] P. Hawrylak, Physical Review B 60, 5597 (1999).
[31] M. Korkusinski, Physical Review B 74, 075317 (2006).
[32] P. Hawrylak, Physical Review Letters 71, 3347 (1993).
[33] P. Y. Yu and M. Cardona, Fundamentals of Semiconductors: Physics and
Material Properties (Springer, Berlin, 1996).
[34] O. Benson, C. Santori, M. Pelton, and Y. Yamamoto, Physical Review Letters 84, 2513 (2000).
[35] T. M. Stace, G. J. Milburn, and C. H. W. Barnes, Physical Review B 67,
085317 (2003).
[36] D. Gammon, E. S. Snow, B. V. Shanabrook, D. S. Katzer, and D. Park,
Physical Review Letters 76, 3005 (1996).
[37] R. M. Stevenson, R. J. Young, P. Atkinson, K. Cooper, D. A. Ritchie, and
A. J. Shields, Nature (London) 439, 179 (2006).
[38] N. Akopian, N. H. Lindner, E. Poem, Y. Berlatzky, J. Avron, D. Gershoni,
B. D. Gerardot, and P. M. Petroff, Physical Review Letters 96, 130501
(2006).
[39] M. Atature, J. Dreiser, A. Badolato, A. Hogele, K. Karrai, and A. Imamoglu,
Science 312, 551 (2006).

[40] B. D. Gerardot, D. Brunner, P. A. Dalgarno, P. Ohberg,


S. Seidl, M. Kroner, K. Karrai, N. G. Stoltz, P. M. Petroff, and R. J. Warburton, Nature
(London) 451, 441 (2007).
[41] J. Berezovsky, M. H. Mikkelsen, O. Gywat, N. G. Stoltz, L. A. Coldren, and
D. D. Awschalom, Science 314, 1916 (2006).
242

BIBLIOGRAPHY

BIBLIOGRAPHY

[42] M. H. Mikkelsen, J. Berezovsky, N. G. Stoltz, L. A. Coldren, and D. D.


Awschalom, Nature Physics 3, 770 (2007).
[43] M. Ediger, G. Bester, A. Badolato, P. M. Petroff, K. Karrai, A. Zunger, and
R. J. Warburton, Nature Physics 3, 774 (2007).
[44] H. J. Kimble, M. Dagenais, and L. Mandel, Physical Review Letters 39, 691
(1977).
[45] F. Diedrich and H. Walther, Physical Review Letters 58, 203 (1987).
[46] T. Basche, W. E. Moerner, M. Orrit, and H. Talon, Physical Review Letters
69, 1516 (1992).
[47] P. Michler, A. Imamoglu, M. D. Mason, P. J. Carson, G. F. Strouse, and
S. K. Buratto, Nature (London) 406, 968 (2000).
[48] C. Kurtsiefer, S. Mayer, P. Zarda, and H. Weinfurter, Physical Review Letters 85, 290 (2000).
[49] R. Brouri, A. Beveratos, J.-P. Poizat, and P. Grangier, Optics Letters 25,
1294 (2000).
[50] A. Hogele, C. Galland, M. Winger, and A. Imamoglu, Physical Review
Letters 100, 217401 (2008).
[51] R. Hanbury Brown and R. Q. Twiss, Nature (London) 178, 1447 (1956).
[52] Z. Yuan, B. E. Kardynal, R. M. Stevenson, A. J. Shields, C. J. Lobo,
K. Cooper, N. S. Beattie, D. A. Ritchie, and M. Pepper, Science 295, 102
(2002).
[53] P. W. Fry, I. E. Itskevich, D. J. Mowbray, M. S. Skolnick, J. J. Finley, J. A.
Barker, E. P. OReilly, L. R. Wilson, I. A. Larkin, P. A. Maksym, et al.,
Physical Review Letters 84, 733 (2000).
[54] S. A. Empedocles and M. G. Bawendi, Science 278, 2114 (1997).
[55] J. J. Finley, M. Sabathil, P. Vogl, G. Abstreiter, R. Oulton, A. I. Tartakovskii, D. J. Mowbray, M. S. Skolnick, S. L. Liew, A. G. Cullis, et al.,
Physical Review B 70, 201308 (2004).
[56] R. J. Warburton, C. Schaflein, D. Haft, F. Bickel, A. Lorke, K. Karrai, J. M.
Garcia, W. Schoenfeld, and P. M. Petroff, Nature (London) 405, 926 (2000).
243

BIBLIOGRAPHY

BIBLIOGRAPHY

[57] K. J. Vahala, Nature (London) 424, 839 (2003).


[58] J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, Photonic
Crystals: Molding the Flow of Light (Princeton University Press, Princeton,
2008), 2nd ed.
[59] R. I. Slusher, A. F. J. Levi, U. Mohideen, S. L. McCall, S. J. Pearton, and
R. A. Logan, Applied Physics Letters 63, 1310 (1993).
[60] L. Zhang and E. Hu, Applied Physics Letters 82, 319 (2003).
[61] H. Y. Ryu, M. Notomi, E. Kuramoti, and T. Segawa, Applied Physics Letters 84, 1067 (2004).
[62] O. Painter, R. K. Lee, A. Scherer, A. Yariv, J. D. OBrien, P. D. Dapkus,
and I. Kim, Science 284, 1819 (1999).
[63] M. Rohner, J. P. Reithmaier, A. Forchel, F. Schafer, and H. Zull, Applied
Physics Letters 71, 488 (1997).
[64] P. Michler, A. Kiraz, L. Zhang, C. Becher, E. Hu, and A. Imamoglu, Applied
Physics Letters 77, 184 (2000).
[65] A. Kress, F. Hofbauer, N. Reinelt, M. Kaniber, H. J. Krenner, R. Meyer,
G. Bohm, and J. J. Finley, Physical Review B 71, 241304 (2005).
[66] J. M. Garca, T. Mankad, P. O. Holtz, P. J. Wellman, and P. M. Petroff,
Applied Physics Letters 72, 3172 (1998).
[67] A. Vasanelli, R. Ferreira, and G. Bastard, Physical Review Letters 89,
216804 (2002).
[68] K. Hennessy, A. Badolato, A. Tamboli, P. M. Petroff, E. Hu, M. Atat
ure,
J. Dreiser, and A. Imamoglu, Applied Physics Letters 87, 021108 (2005).
[69] Y. Akahane, T. Asano, B.-S. Song, and S. Noda, Nature (London) 425, 944
(2003).
[70] L. C. Andreani, D. Gerace, and M. Agio, Photonics and Nanostructures 2,
103 (2004).
[71] C. F. Wang, A. Badolato, I. Wilson-Rae, P. M. Petroff, E. Hu, J. Urayama,
and A. Imamoglu, Applied Physics Letters 85, 3423 (2004).

244

BIBLIOGRAPHY

BIBLIOGRAPHY

[72] G. Bjork, A. Karlsson, and Y. Yamamoto, Physical Review A 50, 1675


(1994).
[73] M. P. van Exter, G. Nienhuis, and J. P. Woerdman, Physical Review A 54,
3553 (1996).
[74] L. A. Coldren and S. W. Corzine, Diode Lasers and Photonic Integrated
Circuits (Wiley, New York, 1995).
[75] G. Bjork, A. Karlsson, and Y. Yamamoto, Applied Physics Letters 60, 304
(1992).
[76] R. Jin, D. Boggavarapu, M. Sargent, P. Meystre, H. M. Gibbs, and
G. Khitrova, Physical Review A 49, 4038 (1994).
[77] Y.-S. Choi, M. T. Rakher, K. Hennessy, S. Strauf, A. Badolato, P. M.
Petroff, D. Bouwmeester, and E. L. Hu, Applied Physics Letters 91, 031108
(2007).
[78] D. Bimberg, N. Kirstaedter, N. N. Ledentsov, Z. I. Alferov, P. S. Kopev,
and V. M. Ustinov, IEEE Journal of Selected Topics in Quantum Electronics
3, 196 (1997).
[79] J. Vuckovic, O. Painter, Y. Xu, A. Yariv, and A. Scherer, IEEE Journal of
Quantum Electronics 35, 1168 (1999).
[80] A. Badolato, K. Hennessy, M. Atat
ure, J. Dreiser, E. Hu, P. M. Petroff, and
A. Imamoglu, Science 308, 1158 (2005).
[81] E. Moreau, I. Robert, L. Manin, V. Thierry-Mieg, J. M. Gerard, and
I. Abram, Physical Review Letters 87, 183601 (2001).
[82] L. Besombes, K. Kheng, L. Marsal, and H. Mariette, Physical Review B 63,
155307 (2001).
[83] M. Bayer, O. Stern, P. Hawrylak, S. Fafard, and A. Forchel, Nature (London) 405, 923 (2000).
[84] M. Fujita and T. Baba, Applied Physics Letters 80, 2051 (2002).
[85] H.-G. Park, S.-H. Kim, S.-H. Kwon, Y.-G. Ju, J.-K. Yang, J.-H. Baek, S.-B.
Kim, and Y.-H. Lee, Science 305, 1444 (2004).

245

BIBLIOGRAPHY

BIBLIOGRAPHY

[86] J. Hendrickson, B. C. Richards, J. Sweet, S. Mosor, C. Christenson, D. Lam,


G. Khitrova, H. M. Gibbs, T. Yoshie, A. Scherer, et al., Physical Review B
72, 193303 (2005).
[87] S. Strauf, K. Hennessy, M. T. Rakher, Y.-S. Choi, A. Badolato, L. C. Andreani, E. L. Hu, P. M. Petroff, and D. Bouwmeester, Physical Review
Letters 96, 127404 (2006).
[88] S. M. Ulrich, C. Gies, S. Ates, J. Wiersig, S. Reitzenstein, C. Hofmann,
A. Loffler, A. Forchel, F. Jahnke, and P. Michler, Physical Review Letters
98, 043906 (2007).
[89] H. Altug, D. Englund, and J. Vuckovic, Nature Physics 2, 484 (2006).
[90] H. Yokoyama, Science 256, 66 (1992).
[91] C. Gies, J. Wiersig, M. Lorke, and F. Jahnke, Physical Review A 75, 013803
(2007).
[92] P. R. Rice and H. J. Carmichael, Physical Review A 50, 4318 (1994).
[93] S.-H. Kim, G.-H. Kim, S.-K. Kim, H.-G. Park, Y.-H. Lee, and S.-B. Kim,
Journal of Applied Physics 95, 411 (2004).
[94] R. Loudon, The quantum theory of light (Oxford University Press, Oxford,
2000), 3rd ed.
[95] S. Ghosh, W. H. Wang, F. M. Mendoza, R. C. Myers, X. Li, N. Samarth,
A. C. Gossard, and D. D. Awschalom, Nature Materials 5, 261 (2006).
[96] P. Borri, W. Langbein, S. Schneider, U. Woggon, R. L. Sellin, D. Ouyang,
and D. Bimberg, Physical Review Letters 89, 187401 (2002).
[97] I. Protsenko, P. Domokos, V. Lef`evre-Seguin, J. Hare, J. M. Raimond, and
L. Davidovich, Physical Review A 59, 1667 (1999).
[98] K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atat
ure, S. Gulde,
S. Falt, E. L. Hu, and A. Imamoglu, Nature (London) 445, 896 (2007).
[99] D. Bouwmeester, A. K. Ekert, and A. Zeilinger, The Physics of Quantum
Information (Springer, Berlin, 2000).
[100] M. Loncar, A. Scherer, and Y. Qiu, Applied Physics Letters 82, 4648 (2003).

246

BIBLIOGRAPHY

BIBLIOGRAPHY

[101] E. Chow, A. Grot, L. W. Mirkarimi, M. Sigalas, and G. Girolami, Optics


Letters 29, 1093 (2004).
[102] J. Vuckovic, D. Fattal, C. Santori, G. S. Solomon, and Y. Yamamoto, Applied Physics Letters 82, 3596 (2003).
[103] J. Vuckovic, M. Loncar, H. Mabuchi, and A. Scherer, Physical Review E
65, 016608 (2002).
[104] S. Strauf, P. Michler, M. Klude, D. Hommel, G. Bacher, and A. Forchel,
Physical Review Letters 89, 177403 (2002).
[105] B. D. Gerardot, S. Strauf, M. J. de Dood, A. M. Bychkov, A. Badolato,
K. Hennessy, E. L. Hu, D. Bouwmeester, and P. M. Petroff, Physical Review
Letters 95, 137403 (2005).
[106] T. Nakaoka, T. Kakitsuka, T. Saito, and Y. Arakawa, Applied Physics Letters 84, 1392 (2004).
[107] E. Grilli, M. Guzzi, R. Zamboni, and L. Pavesi, Physical Review B 45, 1638
(1992).
[108] D. C. Reynolds, K. K. Bajaj, C. W. Litton, G. Peters, and P. W. Yu, Journal
of Applied Physics 61, 342 (1987).
[109] Y. Miura, S. Kimura, Y. Imanishi, and J. Umemura, Langmuir 14, 6935
(1998).
[110] S. Mosor, J. Hendrickson, B. C. Richards, J. Sweet, G. Khitrova, H. M.
Gibbs, T. Yoshie, A. Scherer, O. B. Shchekin, and D. G. Deppe, Applied
Physics Letters 87, 141105 (2005).
[111] S. Strauf, M. T. Rakher, I. Carmeli, K. Hennessy, C. Meier, A. Badolato,
M. J. A. Dedood, P. M. Petroff, E. L. Hu, E. G. Gwinn, et al., Applied
Physics Letters 88, 043116 (2006).
[112] Y. Nakamura, O. Schmidt, N. Jin-Phillipp, S. Kiravittaya, C. Muller,
K. Eberl, H. Grabeldinger, and H. Schweizer, Journal of Crystal Growth
242, 339 (2002).
[113] T. Tran, A. Muller, C. K. Shih, P. S. Wong, G. Balakrishnan, N. Nuntawong,
J. Tatebayashi, and D. L. Huffaker, Applied Physics Letters 91, 133104
(2007).
247

BIBLIOGRAPHY

BIBLIOGRAPHY

[114] A. Yildiz, J. N. Forkey, S. A. McKinney, T. Ha, Y. E. Goldman, and P. R.


Selvin, Science 300, 2061 (2003).
[115] A. Yildiz, M. Tomishige, R. D. Vale, and P. R. Selvin, Science 303, 676
(2004).
[116] C. Kural, H. Kim, S. Syed, G. Goshima, V. I. Gelfand, and P. R. Selvin,
Science 308, 1469 (2005).
[117] J. M. Gerard, Single Quantum Dots: Fundamentals, Applications, and New
Concepts (Springer, Berlin, 2003), chap. 4, pp. 269314.
[118] A. Kiraz, P. Michler, C. Becher, B. Gayral, A. Imamoglu, L. Zhang, E. Hu,
W. V. Schoenfeld, and P. M. Petroff, Applied Physics Letters 78, 3932
(2001).
[119] E. Peter, P. Senellart, D. Martrou, A. Lematre, J. Hours, J. M. Gerard,
and J. Bloch, Physical Review Letters 95, 067401 (2005).
[120] B. Gayral, J. M. Gerard, A. Lematre, C. Dupuis, L. Manin, and J. L.
Pelouard, Applied Physics Letters 75, 1908 (1999).
[121] M. Pelton, J. Vuckovic, G. S. Solomon, A. Scherer, and Y. Yamamoto, IEEE
Journal of Quantum Electronics 38, 170 (2002).
[122] T. Rivera, J.-P. Debray, J. M. Gerard, B. Legrand, L. Manin-Ferlazzo, and
J. L. Oudar, Applied Physics Letters 74, 911 (1999).
[123] C. Wilmsen, H. Temkin, and L. A. Coldren, Vertical-Cavity SurfaceEmitting Lasers (Cambridge University Press, Cambridge, 1999).
[124] L. A. Graham, D. L. Huffaker, and D. G. Deppe, Applied Physics Letters
74, 2408 (1999).
[125] D. G. Deppe, L. A. Graham, and D. L. Huffaker, IEEE Journal of Quantum
Electronics 35, 1502 (1999).
[126] E. R. Hegblom, N. M. Margalit, A. Fiore, and L. A. Coldren, IEEE Journal
of Selected Topics in Quantum Electronics 5, 553 (1999).
[127] A. G. Fox and T. Li, Bell Systems Technical Journal 40, 453 (1961).
[128] S. Schneider, P. Borri, W. Langbein, U. Woggon, R. L. Sellin, D. Ouyang,
and D. Bimberg, IEEE Journal of Quantum Electronics 40, 1423 (2004).
248

BIBLIOGRAPHY

BIBLIOGRAPHY

[129] N. Gisin, G. Ribordy, W. Tittel, and H. Zbinden, Reviews of Modern Physics


74, 145 (2002).
[130] A. N. Boto, P. Kok, D. S. Abrams, S. L. Braunstein, C. P. Williams, and
J. P. Dowling, Physical Review Letters 85, 2733 (2000).
[131] B. Lounis and M. Orrit, Reports of Progress in Physics 68, 1129 (2005).
[132] A. J. Shields, Nature Photonics 1, 215 (2007).
[133] E. Waks, K. Inoue, C. Santori, D. Fattal, J. Vuckovic, G. S. Solomon, and
Y. Yamamoto, Nature (London) 420, 762 (2002).
[134] T.-H. Lee, P. Kumar, A. Mehta, K. Xu, R. M. Dickson, and M. D. Barnes,
Applied Physics Letters 85, 100 (2004).
[135] M. Pelton, C. Santori, J. Vuckovic, B. Zhang, G. S. Solomon, J. Plant, and
Y. Yamamoto, Physical Review Letters 89, 233602 (2002).
[136] C. Santori, D. Fattal, J. Vuckovic, G. S. Solomon, E. Waks, and Y. Yamamoto, Physical Review B 69, 205324 (2004).
[137] N. G. Stoltz, M. Rakher, S. Strauf, A. Badolato, D. D. Lofgreen, P. M.
Petroff, L. A. Coldren, and D. Bouwmeester, Applied Physics Letters 87,
031105 (2005).
[138] W.-H. Chang, W.-Y. Chen, H.-S. Chang, T.-P. Hsieh, J.-I. Chyi, and T.-M.
Hsu, Physical Review Letters 96, 117401 (2006).
[139] K. Takemoto, M. Takatsu, S. Hirose, N. Yokoyama, Y. Sakuma, T. Usuki,
T. Miyazawa, and Y. Arakawa, Journal of Applied Physics 101, 081720
(2007).
[140] V. L. Colvin, M. C. Schlamp, and A. P. Alivisatos, Nature (London) 370,
354 (1994).
[141] I. Gur, N. A. Fromer, M. L. Geier, and A. P. Alivisatos, Science 310, 462
(2005).
[142] J. Urayama, T. B. Norris, J. Singh, and P. Bhattacharya, Physical Review
Letters 86, 4930 (2001).

249

BIBLIOGRAPHY

BIBLIOGRAPHY

[143] J. M. Smith, P. A. Dalgarno, R. J. Warburton, A. O. Govorov, K. Karrai,


B. D. Gerardot, and P. M. Petroff, Physical Review Letters 94, 197402
(2005).
[144] P. O. Holtz, E. S. Moskalenko, M. Larsson, K. F. Karlsson, W. V. Schoenfeld, and P. M. Petroff, in Optical Materials in Defence Systems Technology
III. Edited by Grote, James G.; Kajzar, Francois; Lindgren, Mikael. Proceedings of the SPIE, Volume 6401, pp. 64010I (2006). (2006), vol. 6401 of
Presented at the Society of Photo-Optical Instrumentation Engineers (SPIE)
Conference.
[145] E. Waks, C. Santori, and Y. Yamamoto, Physical Review A 66, 042315
(2002).
[146] D. C. Unitt, A. J. Bennett, P. Atkinson, D. A. Ritchie, and A. J. Shields,
Physical Review B 72, 033318 (2005).
[147] A. Zrenner, E. Beham, S. Stufler, F. Findeis, M. Bichler, and G. Abstreiter,
Nature (London) 418, 612 (2002).
[148] R. Raussendorf and H. J. Briegel, Physical Review Letters 86, 5188 (2001).
[149] M. A. Nielsen, Physical Review Letters 93, 040503 (2004).
[150] T. B. Pittman, B. C. Jacobs, and J. D. Franson, Optics Communications
246, 545 (2005).
[151] J. McKeever, A. Boca, A. D. Boozer, R. Miller, J. R. Buck, A. Kuzmich,
and H. J. Kimble, Science 303, 1992 (2004).
[152] A. J. Bennett, D. C. Unitt, P. See, A. J. Shields, P. Atkinson, K. Cooper,
and D. A. Ritchie, Applied Physics Letters 86, 181102 (2005).
[153] D. J. P. Ellis, A. J. Bennett, A. J. Shields, P. Atkinson, and D. A. Ritchie,
Applied Physics Letters 90, 233514 (2007).
[154] C. Bockler, S. Reitzenstein, C. Kistner, R. Debusmann, A. Loffler, T. Kida,
S. Hofling, A. Forchel, L. Grenouillet, J. Claudon, et al., Applied Physics
Letters 92, 091107 (2008).
[155] D. J. P. Ellis, A. J. Bennett, S. J. Dewhurst, C. A. Nicoll, D. A. Ritchie,
and A. J. Shields, New Journal of Physics 10, 043035 (2008).

250

BIBLIOGRAPHY

BIBLIOGRAPHY

[156] S. Strauf, N. G. Stoltz, M. T. Rakher, L. A. Coldren, P. M. Petroff, and


D. Bouwmeester, Nature Photonics 1, 704 (2007).
[157] J. M. Raimond, M. Brune, and S. Haroche, Reviews of Modern Physics 73,
565 (2001).
[158] M. Nielsen and I. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, 2000).
[159] S. J. van Enk, J. I. Cirac, and P. Zoller, Physical Review Letters 78, 4293
(1997).
[160] J. I. Cirac, P. Zoller, H. J. Kimble, and H. Mabuchi, Physical Review Letters
78, 3221 (1997).
[161] W. Yao, R. Liu, and L. Sham, Physical Review Letters 92, 217402 (2004).
[162] W. Yao, R. Liu, and L. Sham, Physical Review Letters 95, 030504 (2005).
[163] Y. L. Lim, A. Beige, and L. C. Kwek, Physical Review Letters 95, 030505
(2005).
[164] C. Simon, Y.-M. Niquet, X. Caillet, J. Eymery, J.-P. Poizat, and J.-M.
Gerard, Physical Review B 75, 081302 (2007).
[165] H. J. Krenner, E. C. Clark, T. Nakaoka, M. Bichler, C. Scheurer, G. Abstreiter, and J. J. Finley, Physical Review Letters 97, 076403 (2006).
[166] P. Chen, C. Piermarocchi, and L. J. Sham, Physical Review Letters 87,
067401 (2001).
[167] C. Kammerer, G. Cassabois, C. Voisin, M. Perrin, C. Delalande, P. Roussignol, and J. M. Gerard, Applied Physics Letters 81, 2737 (2002).
[168] A. V. Khaetskii, D. Loss, and L. Glazman, Physical Review Letters 88,
186802 (2002).
[169] I. A. Merkulov, A. L. Efros, and M. Rosen, Physical Review B 65, 205309
(2002).
[170] J. R. Petta, A. C. Johnson, J. M. Taylor, E. A. Laird, A. Yacoby, M. D.
Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard, Science 309, 2180
(2005).

251

BIBLIOGRAPHY

BIBLIOGRAPHY

[171] M. Kroutvar, Y. Ducommun, D. Heiss, M. Bichler, D. Schuh, G. Abstreiter,


and J. J. Finley, Nature (London) 432, 81 (2004).
[172] G. A. Narvaez, G. Bester, and A. Zunger, Physical Review B 72, 245318
(2005).
[173] J. T. Shen and S. Fan, Optics Letters 30, 2001 (2005).
[174] E. Waks and J. Vuckovic, Physical Review Letters 96, 153601 (2006).
[175] E. Waks and J. Vuckovic, Physical Review A 73, 041803 (2006).
[176] A. Auff`eves-Garnier, C. Simon, J.-M. Gerard, and J.-P. Poizat, Physical
Review A 75, 053823 (2007).
[177] M. V. G. Dutt, J. Cheng, B. Li, X. Xu, X. Li, P. R. Berman, D. G. Steel,
A. S. Bracker, D. Gammon, S. E. Economou, et al., Physical Review Letters
94, 227403 (2005).
[178] S. J. van Enk, J. McKeever, H. J. Kimble, and J. Ye, Physical Review A
64, 013407 (2001).
[179] J. McKeever, J. R. Buck, A. D. Boozer, A. Kuzmich, H.-C. Nagerl, D. M.
Stamper-Kurn, and H. J. Kimble, Physical Review Letters 90, 133602
(2003).
[180] F. Hofbauer, S. Grimminger, J. Angele, G. Bohm, R. Meyer, M. C. Amann,
and J. J. Finley, Applied Physics Letters 91, 201111 (2007).
[181] A. Mizutani, N. Hatori, N. Nishiyama, F. Koyama, and K. Iga, IEEE Photonics Technology Letters 10, 633 (1998).
[182] B. D. Gerardot, S. Seidl, P. A. Dalgarno, R. J. Warburton, M. Kroner,
K. Karrai, A. Badolato, and P. M. Petroff, Applied Physics Letters 90,
221106 (2007).

[183] M. Geller, A. Marent, T. Nowozin, D. Bimberg, N. Akcay, and N. Oncan,


Applied Physics Letters 92, 092108 (2008).
[184] A. Imamoglu, H. Schmidt, G. Woods, and M. Deutsch, Physical Review
Letters 79, 1467 (1997).
[185] K. M. Birnbaum, A. Boca, R. Miller, A. D. Boozer, T. E. Northup, and
H. J. Kimble, Nature (London) 436, 87 (2005).
252

BIBLIOGRAPHY

BIBLIOGRAPHY

[186] B. Dayan, A. S. Parkins, T. Aoki, E. P. Ostby, K. J. Vahala, and H. J.


Kimble, Science 319, 1062 (2008).
[187] M. Bayer, G. Ortner, O. Stern, A. Kuther, A. A. Gorbunov, A. Forchel,
P. Hawrylak, S. Fafard, K. Hinzer, T. L. Reinecke, et al., Physical Review
B 65, 195315 (2002).
[188] J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, New York,
1994), revised ed.
[189] M. P. Van Exter, W. A. Hamel, J. P. Woerdman, and B. R. P. Zeijlmans,
IEEE Journal of Quantum Electronics 28, 1470 (2002).
[190] G. P. Agrawal and N. K. Dutta, Long-Wavelength Semiconductor Lasers
(Van Nostrand Reinhold, New York, 1986).
[191] G. P. Agrawal and G. R. Gray, Applied Physics Letters 59, 399 (1991).
[192] P. A. Jansson, Deconvolution: with applications in spectroscopy (Academic
Press, New York, 1984).

253

You might also like