Sie sind auf Seite 1von 157

Research Collection

Doctoral Thesis

Mean-Field and Quantum Interactions of Strongly Confined


Exciton-Polaritons

Author(s):
Fink, Thomas

Publication Date:
2018

Permanent Link:
https://doi.org/10.3929/ethz-b-000258681

Rights / License:
In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For more
information please consult the Terms of use.

ETH Library
DISS. ETH NO. 24969

Mean-Field and Quantum Interactions of Strongly


Confined Exciton-Polaritons

A thesis submitted to attain the degree of


DOCTOR OF SCIENCES of ETH ZURICH

(Dr. sc. ETH Zurich)

presented by

Thomas Fink

M.Sc. RWTH
RWTH Aachen University
born on 04.12.1989
citizen of Germany

accepted on the recommendation of

Prof. Dr. Ataç İmamoğlu Examiner


Prof. Dr. Jacqueline Bloch Co-examiner

2018
Abstract

In this dissertation, we report on quantum optics experiments in


a cavity quantum electrodynamics (QED) setting. In particular,
we study the interactions of quantum well (QW) exciton-polaritons
in hybrid dielectric-semiconductor microcavities. A carbon diox-
ide (CO2 ) laser-based fabrication technique is used to create zero-
dimensional polariton boxes in a highly flexible fiber cavity setup at
cryogenic temperatures. We verify that this approach creates strong
optical mode confinement with a confinement area of π × 1 µm2
effectively boosting polaritonic interactions. Using frequency- and
time-domain measurements, we also report long cavity lifetimes of
almost 40 ps, mainly limited by the semiconductor.
The large degree of tunability of the fiber cavity setup as well as
the favorable ratio of nonlinearity compared to the polariton decay
rate allows us to investigate different scenarios: Upon tuning the
experimental parameters, we can go from a regime with vanish-
ing polariton interactions and negligible quantum fluctuations to
one where quantum fluctuations are the dominant source of noise.
Photon correlation measurements are established as a tool to mea-
sure the Liouvillian eigenspectrum dictating the dynamics of the
driven-dissipative system over this whole range. Approaching the
thermodynamic limit with diverging particle number, we demon-
strate the critical slowing down of the system dynamics by more
than nine orders of magnitude compared to the intrinsic decay rate
as it undergoes a first-order dissipative phase transition (DPT).

I
Exploring the opposite regime, we show photon correlation
measurements under weak excitation. An upconversion-based
setup is employed to ensure a time resolution sufficient to
measure single-particle dynamics. The observed antibunching
of g (2) (0) = 0.93 ± 0.04 is a genuinely-quantum manifestation
of exciton-polaritons establishing the onset of single-particle
properties. We report an unexpected long-lived photon bunching
superposing these results which we attribute to a thermal bath of
localized excitons feeding back into the cavity.

II
Zusammenfassung

Diese Dissertation behandelt Quantenoptik-Experimente auf dem


Gebiet der Resonator-Quantenelektrodynamik. Insbesondere wer-
den die Wechselwirkungen von Exziton-Polaritonen in hybriden
Dielektrikum-Halbleiter-Mikroresonatoren untersucht. Mittels ei-
ner CO2 -Laser basierten Fabrikationstechnik stellen wir nulldi-
mensionale Polariton-Boxen in einem hochgradig flexiblen Faserre-
sonator bei kryogenischen Temperaturen her. Wir zeigen, dass
hierdurch die optische Mode räumlich stark auf einen Bereich
mit π × 1 µm2 Flächeninhalt eingeschränkt wird und dadurch die
Polariton-Wechselwirkungen verstärkt. Gleichzeitig beobachten wir
mittels frequenz- und zeitaufgelösten Messungen lange Photonenle-
benszeiten von fast 40 ps, welche massgeblich durch den Halbleiter
limitiert sind.
Die grosse Abstimmbarkeit des Faserresonators in Kombination
mit dem vorteilhaften Verhältnis von Nichtlinearität und Zerfalls-
rate der Polaritonen erlaubt uns die Untersuchung verschiedener
Regimes: Mittels geeigneter Wahl der experimentellen Parameter
können wir das System von einem Szenario mit verschwindender
Polariton-Wechselwirkung und vernachlässigbaren Quantenfluktua-
tionen, bis hin zu einem von Quantenfluktuationen dominierten
Zustand bringen. Wir führen Korrelationen von Photonen als In-
strument ein mit dem sich das Liouvill‘sche Eigenspektrum, welches
die Dynamik des getrieben-dissipativen Systems bestimmt, über
die komplette Spannbreite messen lässt. Nahe dem thermodynami-

III
schen Limit, charakterisiert durch eine divergierende Teilchenzahl,
demonstrieren wir die kritische Verlangsamung der Dynamik des
Systems um mehr als neun Grössenordnungen im Vergleich zur
intrinsischen Zerfallsrate während es einen Phasenübergang erster
Ordnung durchläuft.
Anschliessend zeigen wir Messungen der Photonenkorrelationen im
entgegengesetzten Regime mit geringer Anregung. Dabei benutzen
wir einen auf Frequenzkonversion baiserenden Versuchsaufbau um
eine hinreichende Zeitauflösung zur Messungen der Einteilchendy-
namik sicherzustellen. Das beobachtete Photonen Antibunching
von g (2) (0) = 0.93 ± 0.04 ist eine tatsächlich quantenphysikalische
Erscheinung der Exziton-Polaritonen und demonstriert Ansätze
von Einteilcheneigenschaften. Wir beobachten ein unerwartetes und
langlebiges Photonen Bunching welches diese Messungen überlagert.
Dieses schreiben wir einem thermischen Bad lokalisierter Exzito-
nenzustände zu, welche wiederum den Resonator anregen.

IV
Contents

Title I

Contents V

List of Abbreviations IX

1 Introduction 1

2 Cavity QED with Exciton-Polaritons in Microcavities 7


2.1 Exciton . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Microcavity . . . . . . . . . . . . . . . . . . . . . 12
2.3 Exciton-polaritons . . . . . . . . . . . . . . . . . 17
2.4 Nonlinearity . . . . . . . . . . . . . . . . . . . . . 20
2.5 Confinement . . . . . . . . . . . . . . . . . . . . . 23
2.6 Photon correlations . . . . . . . . . . . . . . . . . 25
2.7 Blockade . . . . . . . . . . . . . . . . . . . . . . . 28

3 Fiber Cavity 31
3.1 Plane-convex Fabry-Pérot cavity . . . . . . . . . 31
3.2 Fiber fabrication . . . . . . . . . . . . . . . . . . 34
3.3 Mode profile . . . . . . . . . . . . . . . . . . . . . 40
3.3.1 Longitudinal . . . . . . . . . . . . . . . . 40
3.3.2 Transversal . . . . . . . . . . . . . . . . . 42
3.3.3 Beam waist measurement . . . . . . . . . 44
3.4 Spectroscopy . . . . . . . . . . . . . . . . . . . . 46

V
Contents

3.5 Noise spectrum . . . . . . . . . . . . . . . . . . . 48


3.6 Strong coupling regime . . . . . . . . . . . . . . . 50
3.7 Conclusion and perspectives . . . . . . . . . . . . 52

4 Dissipative Phase Transition 55


4.1 Master equation of open quantum systems . . . . 55
4.2 Bistability . . . . . . . . . . . . . . . . . . . . . . 57
4.2.1 Mean-field description . . . . . . . . . . . 58
4.2.2 Quantum description . . . . . . . . . . . . 61
4.3 Dissipative quantum phase transition . . . . . . . 64
4.4 Measuring the Liouvillian eigenspectrum with cor-
relations . . . . . . . . . . . . . . . . . . . . . . . 66
4.5 Dissipative phase transition in the thermodynamic
limit . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.6 Polariton nonequilibrium dynamics in the quantum
regime . . . . . . . . . . . . . . . . . . . . . . . . 74
4.6.1 Liouvillian gap scaling . . . . . . . . . . . 76
4.7 First- vs. second-order coherence . . . . . . . . . 79
4.8 Conclusion and perspectives . . . . . . . . . . . . 80

5 Quantum Correlations of Exciton-Polaritons 83


5.1 Upconversion-based photon correlation measure-
ments . . . . . . . . . . . . . . . . . . . . . . . . 85
5.1.1 Setup . . . . . . . . . . . . . . . . . . . . 87
5.1.2 Characterization . . . . . . . . . . . . . . 89
5.2 Polariton correlation measurements . . . . . . . . 92
5.2.1 Same-time correlations . . . . . . . . . . . 95
5.2.2 Time-dependent correlations . . . . . . . 98
5.3 Difficulties in reproducing photon antibunching . 101
5.4 Conclusion and perspectives . . . . . . . . . . . . 102

6 Conclusion and Outlook 105

A Sample Characterization 109

Bibliography i

List of Figures xxvii

VI
Contents

List of Tables xxix

Acknowledgments xxxi

VII
List of Abbreviations

AFM atomic force microscope


APD avalanche photodiode
BEC Bose-Einstein condensate
CCD charge-coupled device
CW continuous wave
DBR distributed Bragg reflector
DPT dissipative phase transition
EOM electro-optic modulator
FDTD finite-difference time-domain
FIB focused ion beam
FPC fiber polarization controller
FWHM full width half maximum
HBT Hanbury Brown and Twiss
HH heavy hole
IRF instrument response function
LH light hole
MBE molecular beam epitaxy
PL photoluminescence
ppLN periodically poled lithium niobate
PSD power spectral density
QED quantum electrodynamics
QPM quasi phase matching
QW quantum well
RMS root mean square

IX
List of Abbreviations

SEM scanning electron microscope


SFG sum frequency generation
SO split-off band
SSPD superconducting single-photon detector
TCSPC time-correlated single photon counting
TEM transverse electromagnetic

AlAs aluminum arsenide


AlGaAs aluminum gallium arsenide
CO2 carbon dioxide
GaAs gallium arsenide
HeNe helium neon
InAs indium arsenide
InGaAs indium gallium arsenide
SiO2 silicon dioxide
Ta2 O5 tantalum pentoxide
ZnSe zinc selenide

X
Chapter 1

Introduction

The idea of quantum simulation was first conceived by Richard


Feynman in his seminal lecture Simulating physics with computers
[1] more than three decades ago, stating that

Nature isn’t classical, dammit, and if you want to make a simulation


of nature, you’d better make it quantum mechanical, and by golly
it’s a wonderful problem, because it doesn’t look so easy.
— Richard P. Feynman

This short excerpt already conveys the concept of quantum simula-


tion: In order to understand a complex physical system, one can
mimic its behavior in another, easily accessible and controllable
system, and test models and hypotheses there. While this reasoning
does not a priori exclude simulations of a quantum system by a
classical one, most notably computers, the exponential scaling be-
havior of the problem size dooms such approaches to be applicable
only to small model systems. Suppose we wish to simulate N spin-
1/2 particles. In order to express the full wave function of these
particles we need 2N coefficients. For N = 30, the space required to
store them is already comparable to the memory size of a personal
computer, and for N > 50 the simulation becomes intractable on
any state-of-the-art classical system. Similar arguments hold for
the exponentiation of the 2N by 2N matrices required to calculate

1
1. Introduction

2000

Publications per year


1500

1000

500

0
1980 1990 2000 2010
Year

Figure 11: Historical publication data. Number of publications containing the


term “quantum simulation” per year extracted from Google Scholar.

the time evolution. In 1996, Seth Lloyd showed that Feynman’s


conjecture is indeed true by proving that a quantum system, on the
other hand, can simulate any many-body Hamiltonian efficiently,
that is with polynomial growth of the required computation time as
the problem size increases [2]. Already today, the simulations of a
wealth of exciting phenomena – specifically those that rely on many-
body interactions – pose a formidable challenge to computation
power [3]. Among those, strongly correlated electrons in cuprous
oxides giving rise to high-temperature superconductivity stand out
due to their enormous potential for applications [4, 5]. There is
hence a paramount need to go beyond classical simulations and in-
voke quantum physics in order to simulate and understand complex
many-body quantum phenomena. Naturally, the quest to realize
such a quantum simulator has stimulated a tremendous amount
of research over the past decades reflected by the ever-increasing
number of publications on the topic illustrated in Fig. 11.
Akin to the DiVincenzo criteria for a quantum computer [6], Ignacio
Cirac and Peter Zoller [7] defined criteria for a quantum simulator
requiring that it (i) be made up of bosons and/or fermions, (ii) can
be initialized in a known quantum state, (iii) allows to engineer its
Hamiltonian, possibly involving a reservoir for open-system sim-
ulations, (iv) can be measured locally and/or globally, (v) allows
for verification of its result, e.g. by first benchmarking it with a
problem where the solution is known. Different physical systems

2
are competing to realize a scalable quantum simulation platform
that features these properties. Currently, ultracold atoms in optical
lattices and trapped ions are the most advanced technologies in
this regard, with a Hilbert space size on the order of 50 individ-
ually accessible sites on which a simulation can be performed [8].
First experiments on quantum dynamics [9] and fermionic Hubbard
models [10] entered regimes where the problem becomes intractable
with classical computers. However, the race is still open and also
other approaches, such as gate-defined quantum dots and supercon-
ducting circuits, feature desirable properties (see Refs. [3, 11] for
reviews on different experimental approaches).
Another less-prominent solid-state contender are exciton-polaritons.
These quasiparticles, formed by the strong coupling of QW excitons
to an optical cavity field, have undergone a remarkable journey since
their first observation by Claude Weisbuch et al. in 1992 [12]. Being
half-light, half-matter in nature, they combine highly desirable
properties from their photonic (extremely low mass, large degree of
coherence, ease of initialization, manipulation, and detection) and
excitonic (strong repulsive interactions) constituents. These initially
sparked interest in fundamental phenomena exploiting this unique
combination which gave rise to the concept of exciton-polariton
Bose-Einstein condensates (BECs) first predicted in 1996 [13] and
realized experimentally one decade later [14]. More experiments,
such as stimulated parametric scattering [15], polariton BECs up
to room temperature [16, 17], long-range spatial coherence with
Berezinskii-Kosterlitz-Thouless-type algebraic decay [14, 18, 19],
superfluidity [20, 21], and quantized vortices [22, 23], to name only
a few, followed suit.
Recently, however, advances in fabrication techniques stimulated
another novel and more application-geared direction of polariton re-
search. The ability to easily tailor the potential landscape of polari-
tons at will [24] enabled the creation of ultra-low threshold polariton
lasers as well as all-optical polariton transistors [25]. Combining
many such elements allowed for the realization of driven-dissipative
nonlinear photonic simulators [26]. These are particularly attractive
due to their all-optical initialization, manipulation, and readout
with readily available optical techniques, as well as scalability aris-

3
1. Introduction

ing from their solid-state nature. In fact, these experiments already


involved more than 103 lattice sites, and also fulfilled the last three
criteria for a quantum simulator. The main obstacle to enable
full-fledged quantum simulations with exciton-polariton lattices
is yet the lack of single-particle interactions. Whereas there is
indeed a finite repulsive interaction between two such particles, its
magnitude is typically so weak that it becomes observable only
for macroscopic populations, and most experiments to date can be
explained in a classical mean-field picture. Increasing polariton in-
teractions and observation of emerging quantum dynamics is hence
crucial to enable all-optical and scalable exciton-polariton quantum
simulators.
In this dissertation, we seek to fill this gap. To this end, we
restrict ourselves to the most fundamental building block of such a
quantum simulator, namely a single zero-dimensional polariton box.
We create experimental conditions in order to maximize polariton
interactions in a highly tunable fiber cavity setup. This allows to
study polariton interactions both in a mean-field regime, where
they give rise to a dissipative phase transition, and in the quantum
regime, where the onset of single-particle interactions is observed.

Scope of this dissertation


We review the theoretical framework of semiconductor QW excitons,
microcavities, and their mutual coupling creating exciton-polariton
quasiparticles in Chapter 2. Different techniques to confine these
polaritons are discussed as well as their interactions. Photon statis-
tics and the emergence of polariton blockade are introduced as
concepts to quantify this interaction.
Chapter 3 then introduces the fabrication technique used to create
strong polariton confinement and thus boost polariton-polariton
interactions. Characterization measurements reveal a mode area of
about 3 µm2 and cavity quality factors in excess of 80 000. Limita-
tions and extensions of this work to increase the latter are discussed.
In Chapter 4 we investigate the dynamics of the driven-dissipative
nonlinear polariton system. In particular, we introduce photon

4
correlations as an analysis tool to access the Liouvillian eigen-
spectrum of a system and use this technique to demonstrate the
nonlinearity-induced DPT that polaritons undergo under certain
driving conditions. For large populations, the closing of the en-
ergy gap as this transition is crossed leads to a critical slowing
down of the system dynamics by more than 10 orders of magnitude
compared to the intrinsic decay time. There, strongly suppressed
quantum fluctuations compete with classical noise sources.
Chapter 5 is devoted to the opposite regime and studies nonlin-
ear polariton interactions in the few-particle regime. To this end,
we introduce a novel technique to measure photon correlations
with picosecond time resolution based on sum frequency generation
(SFG). We use this technique to measure photon correlations of
polariton emission and demonstrate sub-Poissonian statistics with
g (2) (0) = 0.93 ± 0.04, heralding the onset of quantum behavior of
polaritons. Surprisingly, we identify a yet to be understood com-
peting mechanism leading to long-lived bunched photon statistics.
Having established the rich physics arising from the driven-
dissipative nonlinear polariton system down to the few-particle
regime, Chapter 6 puts our results into the context of other
experiments and presents concluding remarks.

5
Chapter 2

Cavity QED with


Exciton-Polaritons in
Microcavities

In this chapter we lay the theoretical groundwork required for


the concepts used in this dissertation. Sections 2.1 to 2.3
introduce excitons, microcavities, and the physics of their
mutual coupling to form exciton-polaritons. We will then
discuss the nonlinear properties of these quasiparticles in
Section 2.4, the effect of confinement in Section 2.5, and
combine these to describe polariton blockade in Section 2.7.
Section 2.6 then introduces photon correlations as a tool to
measure the blockade phenomenon.

2.1 Exciton
Semiconductors are materials with an energy gap between the
valence and conduction band. Since the Fermi energy lies in between
those bands, the valence band is completely filled with electrons, and
the conduction band is empty [27]. At equilibrium, the ground state
of the material is hence insulating. Upon excitation with sufficient
energy overcoming the energy gap, an electron from the valence

7
2. Cavity QED with Exciton-Polaritons in Microcavities

E
InAs (GaAs) at 300 K:
X-valley EΓ = 0.35 (1.42) eV
Γ-valley EL = 1.08 (1.71) eV
L-valley EX = 1.37 (1.90) eV
ESO = 0.41 (0.34) eV
EX
EΓ EL
[100] [111]
k
ESO HH
LH

SO

Figure 21: InAs/GaAs band structure. Schematic band structure of bulk


InAs / GaAs along the [111] and [100] momentum-space directions. The bandgap
for both is direct with a minimum at the Γ-point. The light- and heavy-hole bands
(LH and HH, respectively) are degenerate there and form the valence band. Due to
the large energy difference, the split-off band (SO) can be neglected. The figure
is reproduced from Ref. [28, 29]. The energy gap parameters at the Γ-, L-, and
X-Point as well as the split-off band energy are quoted at room temperature.

band can be promoted to the conduction band. It leaves behind a


hole in the valence band, which can be considered as its antiparticle
with negative mass and positive charge. These two particles can
move in their respective band and contribute to charge transport.
This excited state of the semiconductor is hence conductive. The
extremal values of the two bands, i.e. the wavevectors for which the
energy gap is minimal, classify semiconductors into two families:
If the minimum of the conduction band and the maximum of the
valence band coincide in wavevector, the semiconductor is called
direct. For different wavevectors, the transition is indirect and the
absorption of a photon invokes a phonon to compensate for the
momentum mismatch. Such materials, most notably silicon, are
poor light emitters.
The materials relevant for this dissertation are indium arsenide
(InAs) and gallium arsenide (GaAs), as well as combinations thereof.
They are direct III-V semiconductors with zinc blende crystal struc-
ture grown by molecular beam epitaxy (MBE). The band structure
of both InAs and GaAs is qualitatively similar and displayed along

8
2.1. Exciton

the X-, Γ-, and L-Point in Fig. 21. The lowest-energy conduction
band Bloch orbital has s-type symmetry with zero orbital angular
momentum and is hence twofold degenerate with the electron spin
Sz = ±1/2. The highest-energy valence band orbitals, on the other
hand, have a p-type symmetry. The so-called split-off band with
total angular momentum J = 1/2 is separated due to spin-orbit
coupling by 0.41 meV and 0.34 meV for InAs and GaAs, respectively
[28, 29]. We hence neglect it for the rest of the discussion. At the
Γ-point (k = 0), the J = 3/2 valence bands with Jz = ±3/2 (heavy
hole) and Jz = ±1/2 (light hole) have a fourfold degeneracy. Due
to their different effective masses, however, they split away from
that point.
Upon absorption of an optical photon with an energy larger than
the bandgap, the electron promoted to the conduction band and the
hole left behind in the valence band are mutually attracted due to
Coulomb interaction. They form two-particle states in close analogy
to those of hydrogen atoms, characterized by a renormalized Bohr
radius a∗B and binding energy Ry∗ . For GaAs, a∗B ≈ 10 nm, exceeding
its atomic counterpart by two orders of magnitude due to the low
effective mass (m∗e = 0.063 me , m∗hh = 0.51 me , and m∗lh = 0.082 me
for electron, heavy-, and light-hole, respectively, where me is the
free electron mass) and large dielectric constant ( = 12.9) of the
material [28]. Eventually, the electron and hole will recombine upon
emission of an optical photon with the corresponding energy. The
bound electron-hole pair is called exciton. Considering all possible
combinations between conduction and valence bands, excitons can
have total angular momenta of ±1 and ±2. Since the photon
angular momentum is ±1, only excitons with matching momenta
can couple directly to the circular polarization of light. They are
hence called bright excitons whereas excitons with total angular
momenta ±2 are referred to as dark.
Although InAs and GaAs exhibit qualitatively similar band struc-
tures, the bandgap energy as well as lattice constant illustrated
in Fig. 22 (a) are quite different. Ternary compounds indium gal-
lium arsenide (Inx Ga1−x As), where x denotes the molar fraction
of In, hence allow to tune the bandgap illustrated in Fig. 22 (a)
by choosing the appropriate composition. Moreover, InGaAs het-

9
2. Cavity QED with Exciton-Polaritons in Microcavities

GaAs x cb (b)
1.5 0 1 0.8

Wavelength (µm)
EGaAs
Bandgap (eV)
E
1.0 1.2 z
Inx Ga1-x As -
EInGaAs
0.5 2.4
(a) InAs +

5.8 6.0
vb
Lattice constant (Å)

Figure 22: Excitons in Inx Ga1-x As quantum wells. (a) Lattice constant and
bandgap of ternary Inx Ga1-x As compounds at cryogenic temperatures of 2 K [30].
The black arrow at x = 4 % indicates the indium mole fraction of the sample used in
this dissertation. (b) Schematic of quantum well excitons. A GaAs / InGaAs / GaAs
heterostructure locally lowers the bandgap creating two-dimensional confinement
in a layer perpendicular to the growth direction z. The resulting confined states
are illustrated by black dashed lines together with their wave functions. Electrons
and holes in this layer form bound states due to Coulomb interaction eventually
emitting a photon upon recombination. cb and vb denotes the conduction and
valence band, respectively.

erostructures consisting of layers with different compositions enable


bandgap engineering. In particular, sandwiching an InGaAs layer
in between bulk GaAs locally lowers the bandgap in that region.
If the thickness of the center layer is on the order of the exciton
Bohr radius, discretized excitonic energy states emerge in this so-
called QW structure and the exciton is quantum-confined in a
two-dimensional plane perpendicular to the growth direction. Fig-
ure 22 (b) illustrates this scenario schematically. In such structures,
the band structure depicted in Fig. 21 changes drastically and can
be calculated with k · p theory [31]: The spatial confinement creates
eigenstates consisting of superpositions of low momenta along the
growth direction. Since their confinement energy is determined by

~2 π 2
E= , (2.1)
2m∗ d2
where d is the quantum well thickness, the degeneracy of the heavy
and light hole states at the Γ-point is lifted due to their different
masses. Moreover, in contrast to the isotropic case of bulk material,

10
2.1. Exciton

the effective mass along the growth direction now deviates from the
one perpendicular to it. In fact, the heavy-hole state in the quantum
well has a lower effective mass perpendicular to the confinement
direction, and the light-hole state is heavier, which is referred to
as mass inversion [32]. These two bands must hence cross at finite
momenta giving rise to an anticrossing due to hybridization for
momenta perpendicular to the growth direction ∼ 1/a∗B . While this
is indeed the case for e.g. GaAs QWs in AlGaAs spacer layers, the
confinement potential in the compressively strained InGaAs / GaAs
configuration considered here is shallower and the light hole state
is only weakly confined, if at all. The only relevant valence band is
hence the heavy hole [32, 33].

Care must be taken to find the appropriate indium content suitable


for the application. On the one hand, a large In fraction is desirable
to reduce absorption losses of photons emitted by quantum well
excitons through reabsorption in the surrounding bulk GaAs layers.
On the other hand, a large indium content introduces strain and
defects in the crystal growth due to its larger lattice constant. The
arising fluctuations in quantum well thickness causes inhomoge-
neous broadening of the excitonic emission spectrum. Since this
effect is typically much stronger than the absorption by the GaAs
layers, a low In content of x = 4 % has been found as a good com-
promise to create a QW structure in this dissertation (black arrow
in Fig. 22 (a)). In Appendix A photoluminescence and transmission
spectra of the QW emission of the used sample are presented.

In a second-quantized form, we treat the ground state of the semi-


conductor with filled valence and empty conduction band as a
quasivacuum. Excitons then emerge as quasiparticle excitations on
top of this state. Since their two constituents are fermions, excitons
obey bosonic statistics. Assuming a parabolic dispersion in the
vicinity of the Γ-point, the Hamiltonian describing the excitation
of excitons with in-plane wavevector k perpendicular to the growth
direction of the QW can be written as [34]

d2 k
ZZ
Ĥexc = ~ωexc,k b̂†k b̂k , (2.2)
(2π)2

11
2. Cavity QED with Exciton-Polaritons in Microcavities

where
~2 |k|2
~ωexc,k = ~ωexc,0 + . (2.3)
2mexc

The bosonic operators b̂k and b̂†k annihilate and create an excitonic
excitation in the QW with in-plane wavevector k. Note that in
Eq. (2.2) we have dropped the sum over the two different spin
configurations of the bright exciton states. We will keep this con-
vention for the remainder of this dissertation since we are not
interested in spin-resolved phenomena. The exciton effective mass
mexc = (1/m∗e + 1/m∗hh )−1 is determined by the inverse curvature
of the conduction and heavy-hole band in the dispersion relation.
Note that the bosonic treatment is only valid for low exciton densi-
ties. At higher densities, phase-space filling [35] sets in and Pauli
exclusion of the fermionic constituents starts to screen the Coulomb
attraction. At the so-called Mott transition [36], electrons and holes
disassociate into a plasma [35, 37–40]. For InGaAs, this transition
typically occurs for exciton densities on the order of 1011 cm−2 [41],
well above densities used in this dissertation.

2.2 Microcavity
Fabry-Pérot cavities consist of two reflective layers facing each other
in such a way that a standing wave pattern of photons reflected
back and forth between the mirrors forms. This condition is fulfilled
whenever the spacing between the mirrors L is chosen such that

λ
L=q with q = 1, 2, 3, ... (2.4)
2n

where λ is the desired vacuum wavelength of light and n is the


refractive index of the medium between the mirrors. In that case,
light interferes constructively with itself upon each roundtrip and
this resonance condition allows a build-up of large optical fields
inside the cavity. Two important figures of merit characterize the
properties of such cavities: Firstly, the cavity finesse F determines
the number of roundtrips the photon makes before eventually leaking
through one of the mirrors. It describes how good the mirrors are

12
2.2. Microcavity

and is connected to their reflectivities R by



π R
F≈ , (2.5)
1−R
assuming identical reflectivities larger than 0.5 [42–44]. More com-
monly, however, the quality factor [45]

2πnL
Q=− , (2.6)
λ ln R
is used. It is related to the time the photon spends inside the
cavity before eventually leaking out through one of the mirrors
τ = 1/γ = Q/ω, where τ and γ are the cavity lifetime and linewidth,
respectively. Note that Q depends linearly on the cavity length, i.e.
one can increase the quality factor by just increasing the spacing
between the mirrors without any changes to their reflectivities.
The finesse, on the other hand, depends exclusively on the mirror
properties provided scattering and absorption losses are negligible.
Secondly, the cavity mode volume

d r n(r)2 |E(r)|2
RRR 3
Veff = h i, (2.7)
max n(r)2 |E(r)|2

where n(r) and E(r) are the local refractive index and electric field
amplitude, respectively, denotes the physical volume over which the
cavity’s electric field is extended. As we will see later in this chapter,
for the phenomena discussed in this dissertation it is desirable to
minimize Veff and hence make the cavity as short as possible and
also reduce the lateral size of the optical mode. In the field of cavity
QED with excitons, typical cavity lengths are of the order of a few
optical wavelengths, hence the name microcavity.
Although two closely spaced metallic mirrors can in principle be
used to form such cavities, their reflectivities are typically limited
to below 99 % due to losses. Instead, in order to achieve long
cavity lifetimes even at short cavity lengths, distributed Bragg
reflectors (DBRs) are used. These thin film structures are made up
of alternating layers of different refractive indices n1 and n2 with

13
2. Cavity QED with Exciton-Polaritons in Microcavities

air DBR 1
1 (b)
2

Reflectivity
norm. |E|2

∆E

n
(a)
1
0 0
0 2 4 1.2 1.4 1.6
z (µm) Energy (eV)

Figure 23: Electric field and spectrum of a DBR. (a) DBR with 15 alternating
layers of Ta2 O5 and SiO2 . The blue line (left axis) is the electric field profile of
light impinging perpendicularly onto the DBR surface with a refractive index profile
depicted in orange (right axis). (b) Reflectivity spectrum of the DBR. A stopband
with high reflectivity of 319 meV / 186 nm width (indicated by black arrow) as well
as typical side lobes are visible.

thicknesses λ0 /(4n1 ) and λ0 /(4n2 ), where λ0 is the so-called center


wavelength for which the DBR is designed. For each such bilayer,
reflected light interferes constructively whereas destructive inter-
ference reduces the transmitted field amplitude. Stacking multiple
bilayers on top of each other forms efficient interferometric mirrors
with ultra-high reflectivities exceeding 99.999 %. We use transfer
matrix calculations [46, 47] to illustrate the working principle of a
DBR with a center wavelength of 850 nm made up of 15 bilayers
in Fig. 23. The layer materials are the commonly used dielectrics
Ta2 O5 and SiO2 with refractive indices 2.085 and 1.477, respec-
tively. The electric field amplitude displayed in Fig. 23 (a) rapidly
decays as the light field penetrates into the DBR layers due to
destructive interference. Moreover, the reflectivity spectrum shown
in Fig. 23 (b) exhibits a high reflectivity at and around the center
energy hc/850 nm = 1.46 eV. The bandwidth of high reflectivity is
called stopband and is determined by the refractive index contrast
 
∆λ ∆E 4 n2 − n1
= = arcsin . (2.8)
λ0 E0 π n2 + n1

The dielectric DBR illustrated in Fig. 23 has a stopband width


of 186 nm whereas monolithic semiconductor DBRs made of

14
2.2. Microcavity

GaAs / aluminum arsenide (AlAs), with refractive indices 3.54


and 2.93 at 850 nm, respectively, lead to typical widths of 102 nm.
In contrast to metallic surfaces, there is a sizable penetration of
the electric field into the DBR layers. Defining this length via
the phase light acquires upon reflection of the DBR leads to the
penetration depth at the center wavelength [43, 44, 48]
λ0 1
LDBR = with n1 > n2 . (2.9)
4 n1 − n2

Combining two AlAs / GaAs DBRs with a GaAs spacer layer with
a well-defined thickness determined by Eq. (2.4) in between con-
stitutes a microcavity structure illustrated in Fig. 24. The electric
field amplitude is high in the λ0 /nGaAs -thick spacer layer and de-
cays to both sides upon entering the DBR layers. The green line
highlights the antinode with maximum field amplitude where an
emitter for cavity QED experiments could be placed to ensure good
overlap with the optical mode. In the center of the stopband, the
microcavity features a dip in reflectivity corresponding to the cavity
resonance at which perfect transmission of incident light with the
corresponding wavelength is achieved. The width of this dip is the
cavity linewidth and allows to determine the quality factor. Ex-
perimentally, such GaAs-based monolithic microcavities can have
quality factors exceeding 105 despite their small mode volume [49].
It should be noted that although the spacer layer can be chosen
so short that it hosts only one optical wavelength, most of the
electric field is actually localized in the DBR layers due to the finite
penetration depth. Equation (2.4) hence needs to be modified to
account for this offset. The effective length of a cavity structure
can be defined as [43, 48]

dz 2n2 (z)|E(z)|2
R
Leff = , (2.10)
|E(zqw )|2
where zqw is the position of the emitter in cavity QED experiments.
Calculating this quantity for the structure displayed in Fig. 24 leads
to an effective optical cavity length of [48]
 
λ0
2
Leff = nGaAs 2LDBR + ≈ 10.2 µm (2.11)
nGaAs

15
2. Cavity QED with Exciton-Polaritons in Microcavities

DBR DBR 1
1 (b)

Reflectivity
norm. |E|2

n
2

(a)
0 1 0
0 1 2 1.2 1.4 1.6
z (µm) Energy (eV)

Figure 24: Electric field and spectrum of a microcavity. (a) DBR with 10
alternating layers of AlAs / GaAs sandwiching a GaAs spacer layer with thickness
λ0 /nGaAs . Analogous to Fig. 23, the electric field amplitude decays away from the
center upon entering the DBR layers. Due to the lower index contrast the field
penetrates deeper into the layers. For cavity QED experiments, an emitter would
be placed at the antinode with the highest field amplitude indicated by a green line
in the layer structure on top of the plot. (b) Reflectivity spectrum. The symmetric
DBR configuration enables the formation of a standing wave profile giving rise to the
resonance observed through a minimum in reflectivity where perfect transmittance
occurs. The corresponding energy is determined by the spacer layer thickness.

albeit its short physical length.


Whereas the confinement introduced by the cavity mirrors leads
to quantized longitudinal modes along the growth direction of
the mirror layers, there is a continuum of modes for the in-plane
wavevector k. The polariton dispersion for a cavity reads
~2 |k|2
q
~c
~ωq,k = kq2 + |k|2 ≈ ~ωq,0 + . (2.12)
neff 2mcav
Here, kq = qπneff /Leff is the quantized wave vector along the
mirror direction and k is the in-plane wavevector. In the last
step, we expanded the expression for small in-plane momenta
which yields a parabolic dispersion with an effective photon mass
mcav = ~neff kq /c = n2eff ~ωq,0 /c2 typically on the order of 10−5 me
for microcavities. In second quantized form, the Hamiltonian for the
two-dimensional cavity can be expressed as a harmonic oscillator
[50]
d2 k
ZZ
Ĥcav = ~ωq,k â†k âk . (2.13)
(2π)2

16
2.3. Exciton-polaritons

The bosonic operators âk and â†k destroy and create a cavity photon
with in-plane momentum k, respectively.

2.3 Exciton-polaritons
After introducing QW excitons as well as microcavity photons, we
now discuss their mutual interaction giving rise to the creation
of exciton-polaritons. We combine the individual cavity and QW
exciton Hamiltonians defined in Eqs. (2.2) and (2.13) and their
mutual coherent coupling yielding

d2 k 
ZZ
Ĥ = ~ωexc,k b̂†k b̂k + ~ωq,k â†k âk
(2π)2 (2.14)
h i
† †
+~ΩR b̂k âk + âk b̂k .

Here, the first line is the free evolution of excitons and cavity
photons, respectively, and the second line describes their coherent
interconversion with Rabi frequency ΩR [50]. Again, we do not
explicitly sum over the spin indices but note that the coupling
between exciton and photon is spin-conserving, i.e. excitons with a
given spin only couple to the corresponding circular polarization of
light. The coupling rate is given by
r
4πωcav E(zQW )
ΩR = f2D
, (2.15)
Leff Emax
where E(zQW )/Emax is the normalized electric field amplitude at
the QW position zQW . Generally, the oscillator strength is a dimen-
sionless quantity that describes the number of classical oscillators
at a certain frequency and is related to the absorption or emission
probability of an optical transition. For two-dimensional QW exci-
tons, it is proportional to the area of the sample and hence defined
per unit area [51]
Z 2
2
|huvb |.p|ucb i| |φ(0)| dz χe (z)χh (z) . (2.16)
2 2

f2D =
me ~ωexc
Here, huvb |.p|ucb i is the interband transition momentum element
between conduction and valence band Bloch orbitals ucb and uvb ,

17
2. Cavity QED with Exciton-Polaritons in Microcavities

2
1
(Ecav − Eexc )/~ΩR (a) Ecav
2
Eup |cx |2

(E − Ex )/~ΩR
Elp Eup 1
2 ~ΩR h↠âi 0
0 Eexc
−4
10 0
10−6
−2
10−8 (b) Elp
Ecav Eexc −1
−2 0 2 −5.0 −2.5 0.0 2.5 5.0
(E − Eexc )/~ΩR ~|k|neff /mcav c

Figure 25: Exciton-polaritons. (a) Emission spectrum as a function of the cavity-


exciton detuning. The bare cavity and exciton energies are denoted by white dashed
lines. An avoided crossing with a splitting of 2 ~ΩR emerges when the cavity photon
and exciton are in resonance. The two new eigenmodes, upper and lower polariton,
are labeled correspondingly. (b) Polariton dispersion as a function of normalized
in-plane wavevector k. The bare cavity and exciton dispersion are indicated by
black dashed lines. The latter appears almost flat due to its heavy mass compared
to the cavity photon. The line color denotes the exciton Hopfield coefficient |cx |2
of the polariton modes.

respectively. φ(re − rh ) is the in-plane exciton wavefunction and the


integral term is the overlap of the axial electron and hole envelope
functions (χe (z) and χh (z), respectively). Typically, the oscillator
−2
strength is f2D ∼ 5 × 10−4 Å for InGaAs QW excitons [52, 53].
It is related to the radiative decay rate γexc of a QW exciton by
[54]
π e2 ~
~γexc = f2D . (2.17)
neff 4π0 me c
When the Rabi frequency exceeds the individual decay rates of
excitons and photons, many coherent interconversions occur before
the excitation is lost either due to photon leakage out of the cavity
or decay of the exciton. The system is then in the strong coupling
regime and the Hamiltonian defined in Eq. (2.14) gives rises to two
new eigenstates with eigenenergies
ωexc (k) + ωcav (k)
EUP,LP (k) = ~
2
s (2.18)
ωcav (k) − ωexc (k) 2
 
2
± ~ ΩR + .
2

18
2.3. Exciton-polaritons

Figure 25 (a) shows these eigenvalues as a function of energy de-


tuning between exciton and cavity photon as well as the energy
dispersion of in-plane momenta k in Fig. 25 (b). The signature
of the strong coupling regime is the emergence of an anticrossing
when the exciton and the cavity photon are resonant. Both are then
no longer eigenstates of the system and the two new eigenmodes
are called upper and lower polariton according to their energy
with respect to the exciton. The Hamiltonian in Eq. (2.14) can be
diagonalized by the unitary transformation [55]
    
p̂lp,k −cc cx âk
= , (2.19)
p̂up,k cx cc b̂k

under which Eq. (2.14) can be rewritten in the new eigenbasis

d2 k 
Z 
Ĥ = E p̂†

lp,k lp,k lp,k + E †
up,k up,k up,k up,k .
p̂ p̂ p̂ (2.20)
(2π)3
Here, p̂lp,k and p̂up,k are lower- and upper-polariton operators,
respectively, and the so-called Hopfield coefficients
 
1 ∆ cx
|cx |2 = 1 + q , (2.21)
2 2
4Ω + ∆2
R cx

and |cc |2 = 1 − |cx |2 describe their exciton and cavity content,


respectively. ∆cx = Ecav − Eexc is the energy detuning between
cavity and exciton. For ∆cx = 0, |cc |2 = |cx |2 = 0.5 and both
lower and upper polariton are an equal superposition of exciton
and cavity photon. The effective mass and the decay rate of the
polaritons are the average of the exciton and photon parameters
weighted by the corresponding Hopfield coefficients. This leads to
an extremely low effective mass for the polaritons at |k| = 0 where
mlp = mup ∼ 10−4 mexc due to the small photon mass. Note that
for the polariton decay, only nonradiative exciton decay channels
are relevant since radiative decay just excites a cavity photon. For
purely homogeneous exciton line broadening, the polariton decay
rate is hence only determined by cavity losses γlp ' |cx |2 γcav and
γup ' |cc |2 γcav . In most semiconductor-based experiments, how-
ever, the bare exciton emission is dominated by inhomogeneous

19
2. Cavity QED with Exciton-Polaritons in Microcavities

broadening due to disorder, yet the resulting polariton linewidths


are routinely found to be much smaller. This is due to motional
narrowing which denotes the spatial averaging over the local exci-
tonic disorder potential due to the low effective polariton mass in
the strong coupling regime [56, 57]. Changing the energy detuning
between cavity and exciton modes provides a handle to change the
weights of the exciton and cavity Hopfield coefficients, and hence
how photon- or exciton-like a polariton branch is.
The Hamiltonian introduced in Eq. (2.14) is analogous to the Tavis-
Cummings model [58] describing the physics of many two-level
systems coupled to a common cavity mode. For identical couplings,
cooperative enhancement of the interaction with the cavity mode
arises and collective operators for the sea of two-level systems can be
defined denoting a single excitation shared by all two-level systems
[59]. Here, the exciton operators introduced in Eq. (2.14) are
analogous to such collective operators acting on the sea of excitons
in the semiconductor coupling to the single cavity mode.
Although we used a second-quantized formalism to describe the
coupling between excitons and cavity photons, the emergence of the
two polariton modes can be understood from an entirely classical
point of view by two coupled oscillators. In fact, the effect of the
exciton on the optics of the cavity photon can be modeled by a
classical dielectric polarizability
δ(z − zQW )f2D
χexc (z, k, ω) = . (2.22)
ωexc (k) − ω − iγexc /2
Unlike e.g. in the Jaynes-Cummings model describing the physics
of a single quantum emitter coupled to a cavity mode [60], the
spectrum of exciton-polaritons described by Eq. (2.14) is entirely
harmonic.

2.4 Nonlinearity
Classically, the nonlinearity of a medium may be described via its
polarizability
h i
P (t) = 0 χ(1) E(t) + χ(2) E 2 (t) + χ(3) E 3 (t) + ... . (2.23)

20
2.4. Nonlinearity

Here, χ(1) describes the linear response of the medium to an applied


electric field E, whereas the higher-order terms are its nonlinear
components. For many media, among which liquids, gases, amor-
phous solids, and many crystals, the second-order susceptibility χ(2)
vanishes due to their inversion symmetry, and the first nonzero non-
linear term is the third-order polarization with susceptibility χ(3) ,
also called Kerr nonlinearity [61]. Due to their centrosymmetric
zinc blende structure, GaAs and InGaAs used in this dissertation
are also of this type with χ(2) = 0 and rather large typical nonlinear
susceptibilities compared to other semiconductors on the order of
χGaAs = 10−18 m2 /V2 [61, 62].
(3)

At the microscopic level, the Coulomb interaction between two


excitons in a planar configuration gives rise to a scattering term
X Vxx,q †
Ĥint = b̂k+q b̂†k0 −q b̂k b̂k0 . (2.24)
0
2
k,k ,q

Typically, the momentum exchange wavevector q is much smaller


than the exciton Bohr radius and a two-body contact interaction
with Vxx,q ∼ Vxx,0 = Vxx is assumed [63]. The interaction term is
given by [64]
6e2 a∗B
Vxx ' , (2.25)
A
where A is the confinement area of excitons, i.e. a stronger con-
finement boosts exciton interactions. Polaritons also inherit the
nonlinear interaction from their excitonic part, i.e. their interaction
strength is scaled by the exciton content

Vlp,up = |cx,c |4 Vxx . (2.26)

Determining the single-particle interaction strength is experimen-


tally challenging and typically the collective nonlinearity of a macro-
scopic polariton population is used to infer the single-particle one.
In this way, the two-dimensional exciton-exciton interaction has
been estimated as Vxx 2D = V × A ' 30 µeVµm2 [65–67], in agree-
xx
ment with theoretical predictions [64]. Assuming cavity and exciton
to be on resonance, i.e. |cx |2 = 0.5, and polaritons to be confined in

21
2. Cavity QED with Exciton-Polaritons in Microcavities

disk with a radius of 1 µm, we estimate the polariton nonlinearity


to be
30 µeVµm2
 2
1
Vpol = × = 2.4 µeV. (2.27)
2 π × 1 µm2
For completeness, we also wish to mention results by Sun et
al. [68] where extrapolated exciton-exciton interaction strengths
of (1.74 ± 0.46) meVµm2 were reported, two orders of magnitude
larger than values measured in other experiments and theoretical
predictions. These results should yet be regarded with great caution
since the strong excitonic disorder in this experiment is likely to
falsify the estimation of the particle density on which the value for
the polariton interaction is based. Moreover, the findings rely on
the assumption that the particle population at the detection spot
consists exclusively of |k| = 0 polaritons. In particular, the presence
of a spurious dark excitonic population, which may contribute to
the observed signatures, is excluded. The nonresonant excitation
scheme in this experiment is yet insufficient to ensure this condition
and skepticism towards the drawn conclusions is hence due.
In addition to the direct exchange interaction defined in Eq. (2.24),
polaritons are subject to a second mechanism breaking the harmonic-
ity. For high densities, the fermionic nature of excitons discussed
in Section 2.1 saturates the exciton-photon coupling which can be
modeled by a term
X ~ΩR †
sat
Ĥpol =− âk+q b̂†k0 −q b̂k b̂k0 + h.c. (2.28)
0
n sat A
k,k ,q

Since the densities required for this mechanism to be sizable are


much higher than those relevant for this dissertation, we will neglect
this term.
The formation of two excitons into a biexciton state, equivalent
to a hydrogen molecule, enables a polariton nonlinearity which is
effectively similar to the Kerr-nonlinearity in Eq. (2.24) due to the
lower energy of the bound molecular state [69]. In ultracold atoms,
this Feshbach resonance is widely used to control atom interactions
over vast ranges [70]. In the semiconductor setting considered here,
however, the short lifetime of the bound exciton state leads to

22
2.5. Confinement

strong line broadening [71, 72] which makes it unfavorable from


an experimental point of view compared to the direct exchange
interaction.

2.5 Confinement
So far we have considered planar microcavity systems confined in
the longitudinal direction but with a continuum of in-plane wave vec-
tors k. When the polariton mode is also confined in the transverse
plane with a confinement energy exceeding its linewidth, discrete
momentum eigenstates emerge and the mode is characterized by
three integer numbers q, m, n. Here, q is the longitudinal quantiza-
tion number introduced in Eq. (2.4), and m, n define the transverse
electromagnetic (TEM) mode, e.g. Laguerre-Gaussian modes for
geometries with circular symmetry. Of course, the confinement can
also be restricted to a single in-plane dimension, quantizing e.g.
the momentum states in kx but maintaining a continuum in the ky
direction, thus effectively creating a one-dimensional polariton wire
[73]. In this dissertation, however, we are interested in polaritons
with three-dimensional confinement, also called polariton boxes
[48]. Moreover, we focus on the lowest momentum eigenstate (i.e.
TEM00 ) of the lower polariton mode since all higher-lying eigen-
states may be subject to additional nonradiative decay processes
effectively shortening their lifetime. If we neglect coupling between
the lower and upper polariton mode in Eq. (2.20) as well as between
different momentum eigenstates, the corresponding single-mode
lower-polariton Hamiltonian yields
U † †
Ĥ = Ep p̂† p̂ + p̂ p̂ p̂p̂. (2.29)
2
Here, Ep is the energy of the lowest momentum eigenstate of the
lower polariton mode, and U is the effective polariton interaction
strength defined in Eq. (2.26). Including a coherent drive with
amplitude F and frequency ωL , and going into a frame rotating
with ωL , we write down the Hamiltonian for the driven nonlinear
lower-polariton mode
U † †
Ĥ = −∆p̂† p̂ + p̂ p̂ p̂p̂ + F ∗ p̂† + F p̂. (2.30)
2

23
2. Cavity QED with Exciton-Polaritons in Microcavities

Here, ∆ = EL − Ep is the energy detuning between the drive


frequency and the polariton mode. Experimentally, the coherent
drive is realized by an excitation laser. Since the laser only couples
to the photonic part of the Hamiltonian, its amplitude has to
be scaled by the polariton’s photon content |cc |2 to obtain the
amplitude F in the polariton basis.
Experimentally, there are many different approaches to create in-
plane confinement for polaritons, both by acting on the excitonic as
well as the photonic part. We hence mention only a few and refer
to Ref. [24] for an exhaustive and recent review. The application of
compressive strain via a small metallic tip pressed onto a polariton
microcavity sample has been shown to locally lower the exciton
energy [16]. Although this technique is quite flexible and allows for
in-situ tuning of the trap, the generated confinement, on the order
of 10 µm to 100 µm, is rather weak and did not allow to observe
quantization of the polariton mode. Optical confinement relies on
injecting a large population of localized bare excitons in the sample,
effectively creating a repulsive potential for polaritons due to exciton-
exciton repulsion. This technique allowed to observe quantized
polariton states [73] and even creation of a polariton box [74] with
a donut-shaped laser beam. The approach is extremely flexible
and allows for almost arbitrary geometries by shaping the pump
laser beam. However, the macroscopic population of excitons also
diffuses and may partly relax to polaritons, which is an undesired
process in single-particle experiments relying on well-controlled and
coherent excitation of polaritons.
One of the most versatile techniques is the fabrication of integrated
two-dimensional exciton-polariton samples and subsequent etching
of a microcavity structure. The large refractive index contrast
between the semiconductor material used (typically GaAs with
n = 3.54) and the surrounding vacuum confines the optical mode
transversally [65]. Figure 26 (a) illustrates this showing the inten-
sity distribution of the fundamental optical mode in an exemplary
micropillar. Due to the ability to etch structures with diameters on
the µm scale, this technique in principle allows for strong in-plane
confinement and gives rise to discretized in-plane momentum eigen-
states split by several meV in Fig. 26 (b). However, the increasing

24
2.6. Photon correlations

1.54
(a) (b)
top DBR 1.52

Energy (eV)
spacer 1.50

1.48
bottom DBR
2 µm 1.46
substrate
−5 0 5
kk (µm-1 )

Figure 26: Micropillar cavity. (a) Intensity distribution of the lowest-energy TEM00
photonic mode in a micropillar cavity superimposed onto the material structure.
The top and bottom DBRs consist of 24 AlAs / GaAs bilayers sandwiching a λ/n-
thick spacer layer. The pillar diameters is 2 µm. The index contrast between the
microcavity material and the surrounding vacuum strongly confines the optical
mode. The colormap is logarithmic and spans two decades. The insets left of the
pillar display the in-plane intensity distribution at the center of the spacer layer for
the second- and third-lowest energy modes (bottom and top, respectively). (b)
Dispersion of the micropillar emission spectrum. The strong confinement leads to
quantized momentum eigenstates. The relative intensities of the eigenmodes are
due to different overlaps with the excitation profile of the simulation.

overlap of the optical mode and the surface of the semiconductor


material introduces scattering losses that scale with the degree of
confinement imposing a trade-off between strong confinement and
long polariton lifetime [49]. The strength of the etching technique
lies in the ability to fabricate lattices with hundreds of identical
and coupled polariton boxes with almost arbitrary geometries, and
remarkable experiments exploiting this to emulate one- and two-
dimensional Hamiltonians have been reported [26].

2.6 Photon correlations


The Heisenberg uncertainty principle requires that light exhibits pho-
ton number fluctuations. In addition to frequency, spatio-temporal
profile etc., a photonic state is hence also characterized by its
statistics. Typically, three different classes of photon statistics are
distinguished according to the variance of the state’s distribution
compared to a Poissonian distribution function: Poissonian, super-

25
2. Cavity QED with Exciton-Polaritons in Microcavities

Coherent Thermal Fock


1
p(n)

0
0 5 10 0 5 10 0 5 10
n n n

2
g (2) (t)

0
−5 0 5 −5 0 5 −5 0 5
t/τ t/τ t/τ

Figure 27: Photon statistics of different states. Top row: Photon number dis-
tribution for a coherent, thermal, and Fock state. The average particle number is
Tr ρ̂↠â = 2 in each case. The orange solid line is a Poissonian distribution illus-


trating that the thermal and Fock state have super- and sub-Poissonian variances,
respectively. The coherent state exactly follows this line. Bottom row: Illustra-
tion of corresponding second-order correlation functions. The super-Poissonian
(sub-Poissonian) statistics for the thermal (Fock) state lead to photon bunching
(antibunching). Coherent statistics is recovered after some timescale τ .

Poissonian, and sub-Poissonian. These terms denote whether the


number fluctuations of the state are equal to, larger to, or smaller,
respectively, than those occurring in an uncorrelated, coherent state.
We illustrate this in the top row of Fig. 27 where the photon number
distributionsof three
different states with identical average photon
numbers Tr ρ̂â â = 2 are shown. The orange line corresponds to

a Poissonian distribution and we observe that the coherent state


follows this curve exactly whereas the thermal and Fock state have
wider and narrower distributions, respectively. The photon statis-
tics are intimately linked to the second-order photon correlation

26
2.6. Photon correlations

function defined as [75]




â (t)↠(t + ∆t)â(t + ∆t)â(t)

(2)
g (t, t + ∆t) = † . (2.31)
hâ (t)â(t)ih↠(t + ∆t)â(t + ∆t)i

This quantity describes the likelihood of observing a photon at


time t + ∆t, conditioned on the detection of a photon at time t,
and normalized by the photon numbers at time t and t + ∆t. In
most cases, the absolute time t has no physical significance and
Eq. (2.31) is abbreviated to g (2) (∆t). The bottom row of Fig. 27
displays this quantity for the same states as above. When the
photons are uncorrelated, the conditional probabilities are equal to
the independent ones and we obtain g (2) (∆t) = 1 ∀ ∆t. This is the
case for a coherent state, typical for the output of a laser, where
photon number fluctuations are Poissonian and hence uncorrelated.
When the fluctuations are super-Poissonian, it is more likely to
detect two photons at the same time compared to the average. We
hence obtain g (2) (0) > 1 and the photons are said to be bunched.
A typical state for this case is a thermal state with g (2) (0) = 2.
For long delays, all correlations must vanish such that g (2) (∞) = 1.
The time scale over which the coherent-state value is recovered is
hence characteristic for the underlying photon correlations. For a
thermal state, it is given by the coherence time.
Invoking the Cauchy-Schwarz inequality, one can show that any
classical state must obey

g (2) (0) ≥ 1 (2.32a)


g (2)
(0) ≥ g (2)
(∆t) for ∆t 6= 0. (2.32b)

As expected, both the coherent and thermal state in Fig. 27 fall


into this category. The Fock state with sub-Poissonian statistics,
however, violates both criteria demonstrating that it is a true
quantum state. In this case, it is less likely to detect two photons at
the same time compared to the average, and the photons are said
to be antibunched. This signature is typical for quantum systems
which may only host a single excitation and hence emit only one
photon at a time. The observation of antibunching can be used to
prove the quantum nature of the underlying emitter and gives rise

27
2. Cavity QED with Exciton-Polaritons in Microcavities

to blockade phenomena described in the next section. Note that


in addition to antibunching, a value g (2) (0) < 0.5 signals that the
source is a single quantum emitter instead of an ensemble.
The second-order correlation function is the quantum-mechanical
equivalent of the classical intensity autocorrelation
hI(t)I(t + ∆t)i
(2.33)
(2)
gclass (t, t + ∆t) = .
hI(t)ihI(t + ∆t)i

Experimentally, g (2) (∆t) could be measured by recording the time


trace of incoming photons, e.g. with avalanche photodiodes (APDs).
These devices typically have a deadtime on the order of 10 ns after
a detection event during which they are insensitive to subsequent
photons. In order to still measure photon correlations for delays
below the deadtime, a Hanbury Brown and Twiss (HBT) setup [76,
77] is used. It relies on splitting the incoming light into two parts
with a beamsplitter and directing each onto a separate detector.
Measuring cross-correlations in between the two APDs then allows
to measure photon correlations with arbitrarily small delay, limited
only by the detector resolution. Whenever we measure photon
correlations with APDs in this dissertation we employ such a HBT
configuration.

2.7 Blockade
The polariton-polariton repulsion in Eq. (2.30) causes a blueshift
of the two-polariton manifold with respect to a harmonic ladder of
eigenstates. When the system is excited resonantly, this blueshift
effectively pushes the two-excitation manifold away from the drive.
For polariton interactions comparable to or exceeding the mode
linewidth, the overlap between the resonant drive and the two-
excitation manifold vanishes and the system enters the polariton
blockade regime predicted by Verger et al. [78]. Experimentally,
the photon correlation function introduced in the previous section
provides a good measure to observe such blockade phenomena [79].
Figure 28 illustrates this scenario schematically. In the blockade
regime, the polariton system is a truly quantum one since the oc-
cupation with a single excitation inhibits the presence of another.

28
2.7. Blockade

|ppi
U γ

|pi
γ
Ep

|0i

Figure 28: Polariton blockade. Schematic of a regime when the nonlinearity-


induced energy shift U of the two-excitation manifold is comparable to the mode
linewidth γ. The vanishing overlap of the resonant excitation with energy Ep and
the two-polariton manifold blocks the excitation of a second polariton.

Only after the first polariton decayed, the system may be re-excited
by another incoming photon. This generates a stream of single
photons emitted by the polariton system in a n = 1 Fock state. The
correlations of this photon stream hence yields a strong antibunch-
ing dip with a time scale over which coherent statistics are recovered
given by the polariton lifetime. The antibunching is strongest when
the excitation is slightly red-detuned compared to the polariton
mode. There, it excites the one-polariton manifold on the slope
of the Lorentzian mode, where the sensitivity to energy shifts is
strongest. Conversely, a blue-detuned excitation leads to strong
photon bunching due to a larger overlap with the two-polariton
manifold. In this scenario, the system is more likely to emit two
photons at once than a single one. Figure 29 (a) shows same-time
photon correlations g (2) (0) as a function of detuning in a more
experimentally realistic scenario where the nonlinearity is small
but non-negligible with respect to the mode linewidth. We indeed
observe a crossover from antibunched to bunched photon statistics
as the excitation energy is varied across the polariton mode. More-
over, the picture presented here is only
valid
for weak excitations
with F /γ  1 and mode populations â† â  1. For stronger exci-
tations, illustrated in Fig. 29 (b), the increasing mode population

29
2. Cavity QED with Exciton-Polaritons in Microcavities

1.05 0.99
(a) 101 (b)
g (2) (0)
100 0.98

g (2) (0)
h↠âi
1.00 10−1 0.97
−2
10
0.96
10−3
0.95 0.95
−5 0 5 10−2 10−1 100 101
∆/γ F /γ

Figure 29: g (2) (0) detuning- and power-dependence (a) Same-time photon
correlations for varying excitation laser detuning ∆ for U /γ = 0.05 and F /γ = 0.1.
The photon statistics change from bunching to antibunching when the excitation is
blue- and red-detuned, respectively. (b) Mode population (orange, left axis) and
same-time photon correlations (blue, right axis) for varying excitation strength,
U /γ = 0.05 and ∆/γ = −0.5. For F /γ  1, the population grows linearly with
power and the photon antibunching is constant. When the excitation is strong
compared to the decay rate, antibunching is reduced and a sublinear behavior of
the mode population appears.

reduces the antibunching towards coherent-state statistics.


Assuming a weak drive F /γ  1, we can truncate the Hilbert space
at the two-excitation manifold and neglect occupation of higher-
lying states. In this regime, Ferretti et al. [80] derived an analytical
expression for equal-time photon correlations in the steady state

1 + 4 (∆/~γ)2
g (2) (0) = . (2.34)
1 + 4 (U /2 − ∆)2 / (~γ)2

30
Chapter 3

Fiber Cavity

In this chapter we discuss the fabrication of the hybrid fiber-


based microcavity used for the experiments in this dissertation.
Sections 3.1 and 3.2 will be dedicated to the setup and fabrica-
tion of the so-called fiber cavity followed by measurements on
the mode profile, spectroscopic features, lifetime, mechanical
stability, and strong coupling regime in Sections 3.3 to 3.6,
respectively.

3.1 Plane-convex Fabry-Pérot cavity


In order to exploit the anharmonicity of exciton-polariton systems
described in Chapter 2, the two crucial requirements that a sys-
tem should fulfill are strong optical confinement and long cavity
lifetimes. Moreover, working in a transmission geometry would
be highly desirable since the photon correlation measurements we
seek to perform require extinguishing the excitation light. While
this can be achieved by e.g. polarization suppression, transmission
measurements automatically single out the signal light without any
reduction in detection efficiency. Comparing the different tech-
niques to confine polaritons discussed in Section 2.5 it becomes
clear that most of them lead to a trade-off between confinement
and cavity lifetime making them unsuitable for the goals of this

31
3. Fiber Cavity

dissertation. Only the approach using a plane-convex cavity ge-


ometry promises to create strong confinement without introducing
additional loss channels and thus degrading the cavity quality at
the same time. This type of open plane-convex cavities recently
received a lot of attention due to its small mode volume beneficial
for cavity QED experiments as well as high tunability and has been
successfully used with atoms [81–83], ions [84, 85], GaAs excitons
[48, 86], self-assembled quantum dots [87–89], nitrogen vacancy
centers [90–93], optomechanical systems [94–98], biophysics [99],
and the emerging material class of van der Waals heterostructures
[100–102]. Moreover, tunable tunnel-coupling of two such cavities
has been reported when two convex structures with spatial overlap
are fabricated [103].
We therefore use this type of hybrid cavity approach where the
confinement is created by the convex geometry of one of the mirrors.
At the heart of our setup is a cavity formed by an epitaxially
grown semiconductor AlAs / GaAs DBR and a dielectric tantalum
pentoxide (Ta2 O5 ) / silicon dioxide (SiO2 ) DBR fabricated onto
the facet of an optical fiber with a convex surface profile. For
hemispherical shapes with radius of curvature R, the beam waist of
the mode propagating in this plane-convex cavity is then given by
r
2
4 λ
w0 = Leff (R − Leff ), (3.1)
π2
where Leff is the total cavity length and λ is the wavelength of
the light. In order to create strong confinement, one thus seeks
to minimize R while maintaining the condition R ≥ Leff to ensure
stability of the mode. In Fig. 31 the setup including the cavity
structure as well as in- and outcoupling access is schematically
shown. The fiber cavity has been described in detail in Ref. [48,
104], where a previous generation of samples and fibers with weak
confinement were used.
On top of the semiconductor DBR, a spacer layer containing the
active material, i.e. the QW, is deposited. The fiber and sample
are each mounted on nanopositioners1 allowing for in situ relative
1
ANPx101/RES for the semiconductor sample and ANPz101/RES for the
fiber, both attocube systems AG

32
3.1. Plane-convex Fabry-Pérot cavity

to detector
Substrate

DBR

QW
FPC 4K
Mode
from source
DBR

Fiber

Figure 31: Fiber cavity setup. Light enters the dipstick immersed into liquid
helium through a single-mode fiber (blue lines). Fiber polarization controllers (FPCs)
are used to set the input polarization. The transmitted light (red lines) is then
collimated and leaves the dewar through a window allowing for free-space access.
It is collected in another single-mode fiber and directed towards the respective
detection device used. Zoom-in: Schematic of the cavity geometry. Light couples
from the fiber core to the cavity mode confined by the DBRs deposited on the fiber
facet and semiconductor substrate. The distance between fiber and sample, and
hence cavity energy, can be varied with nanopositioners controlling the position of
the fiber with sub-nanometer precision.

movement in all three dimensions. Note that the lens used for
collimating the transmitted light, however, can only be aligned at
room temperature and thermal contraction while cooling down the
system is likely to cause a shift of its focal position. The cavity is
then mounted at the bottom of a dip-stick filled with 25 mbar He
exchange gas and immersed into a dewar containing 100 L liquid
helium. In this way, the system is kept at a stable temperature
of 4.2 K for 4–5 weeks avoiding additional vibrations caused by
e.g. the pulse tube in closed-cycle configurations [105]. To further
reduce acoustic noise and decouple the cavity from vibrations in
the building, the dewar sits on a passive vibration isolation plate2
and the whole system is enclosed in a box cladded with acoustic
foam.
The optical single-mode fiber onto which the DBR is deposited
exits the dewar through a vacuum-tight Teflon feedthrough [106]
which allows to couple light to the system simply by connecting
the corresponding fiber. Manual fiber polarization controllers along
2
PTT600600, Thorlabs GmbH

33
3. Fiber Cavity

the path are used to set the polarization of the incoming light.
Note, however, that the mode overlap between the guided mode
propagating in the fiber core and the transverse profile of the cavity
mode (cf. the following sections) is small and hence leads to a
low incoupling efficiency. Light leaving the cavity through the
semiconductor mirror is collimated with an aspheric lens and leaves
the dewar through an anti-reflection coated window. The light
is then coupled into another single-mode fiber and sent to the
detection device.

3.2 Fiber fabrication


At the heart of the setup is the DBR-coated fiber facet whose
surface profile creates the strong photonic mode confinement. Next
to a few proof-of-principle experiments, e.g. transfer of convex
high-quality mirrors onto optical fibers [107], mainly two different
techniques have been established to machine the desired depressions
in a substrate which, upon subsequent coating with dielectric DBR
layers, creates the convex mirror surface: focused ion beam (FIB)
milling and CO2 laser ablation. The former relies on sputtering
small amounts of material of the substrate upon irradiation with
Ga+ or He+ ions accelerated to the keV scale [108]. It enables a
large degree of control over the precise surface geometry with in
principle arbitrarily small radii of curvature due to the nanometer to
sub-nanometer precision of the milling process. Moreover, arrays of
dimples with different shapes can be created on the same substrate
enabling in-situ tuning of the cavity parameters. On the other
hand, this technique is rather time-consuming (several hours per
milling session) and the surface roughness of the machined struc-
tures reported are relatively high with root mean square (RMS)
values of 0.7 nm [109]. The second technique, which is used in this
dissertation, relies on surface evaporation of substrate material
upon irradiation with a short CO2 laser pulse. While the spot size,
power, and exposure time determine the depth and shape of the
evaporated area, melting in a thin layer at the surface creates a
smoothing effect of the depression.
The fabrication process with optical fibers has been introduced by J.

34
3.2. Fiber fabrication

6x BW BW
CO2 Laser

dump

VIS Laser

Stage
FM
PR
Fiber

CCD

Figure 32: Fiber fabrication setup. The CO2 light passes two Brewster windows
(BW) that set the transmitted power via their relative angle. A beam expander (6x)
increases the diameter to reduce divergence. The phase-retarding mirror (PR) turns
the linear polarization (↔) of the beam into circular ( ). The CO2 light is then
focused with a ZnSe lens. The fiber sits on a stage that allows for three-dimensional
movement of the facet into the laser focus. A flip mirror (FM) directs visible light
onto the fiber facet. The reflected light is then interfered with another laser beam
in order to image the fiber facet topography with a CCD after the CO2 irradiation.

Reichel [110–113] in the context of atom BEC experiments [81]. In


these, long cavity lifetimes and hence cavity lengths of several 10 µm
with radii of curvature of several 100 µm creating weak confinement
have been used. In this dissertation, we build up on this technique
and optimize it in order to create small radii of curvature capable
of inducing strong photonic mode confinement. A sketch of our
fabrication setup is depicted in Fig. 32. Linearly polarized light from
a CO2 laser3 at 10.6 µm wavelength with 2.5 mm beam diameter
passes two water-cooled attenuators with Brewster windows. Upon
rotation of the second one, the transmitted power can be tuned.
Then, a 6x beam expander is used to increase the beam diameter
and hence decrease the final focal spot size. A larger diameter also
reduces the effect of divergence, which, due to the short Rayleigh
length of 46 cm for this wavelength at the initial diameter, would
otherwise lead to sizable changes of the beam profile upon reaching
the imaging optics. The beam is then directed via two mirrors, out
3
v30, Synrad Inc.

35
3. Fiber Cavity

of which one also induces a phase shift to turn the polarization from
linear to circular reducing the astigmatism of the produced dimples.
An aberration-corrected zinc selenide lens with 25.4 mm focal length
is used to focus the beam onto a spot size with nominally 23 µm
diameter. To create the dimple on the fiber, we move the facet of a
cleaved optical fiber into the focal plane and use the short rise and
fall time of the laser to create a pulse by current modulation. The
power and exposure time have to be chosen such that two undesired
effects are avoided: For short pulses and high powers, resolidification
of the irradiated layer occurs faster than the smoothing process
and the resulting surface roughness would result in high scattering
losses in the cavity. In the opposite regime, when long pulses and
low powers are used, the molten region extends deeply into the fiber
volume and the material contracts resulting in a bulge around the
depression [113]. We find that powers of typically 1.0 W to 2.0 W
and exposure times between 1 ms to 2 ms lead to the desired shape
of a Gaussian depression closely following the intensity profile of
the CO2 laser focus and described by
 2
x + y2

z(x, y) = t · exp − , (3.2)
σ2
where t is the depth of the dimple and 2σ = D its diameter. In the
center, Eq. (3.2) can be expanded to second order and approximated
by a hemisphere with radius of curvature

D2
R= , (3.3)
8t
which determines the confinement created in the cavity according
to Eq. (3.1). This is schematically shown in Fig. 33 along with a
scanning electron microscope image of a machined fiber facet.
In order to analyze the shape of the created dimple in situ we
employ an interferometry-based technique: Upon insertion of a
glass plate with a flip mount indicated in Fig. 32, two collimated
visible laser beams, where one of them reflects off the fiber facet,
enter a microscope objective and are imaged with a charge-coupled
device (CCD) camera. The different path lengths generated by the
profile of the fiber facet induce interference fringes in the recorded

36
3.2. Fiber fabrication

(a) (b)
t

R
D

Figure 33: Dimple profile on fiber facet. (a) Schematic illustrating the Gaussian
depression on the fiber facet with depth t and diameter D = 2σ. In the center,
the dimple can be approximated by a sphere with radius R. (b) SEM image of a
fabricated fiber with large radius of curvature dimple before coating taken from
Ref. [112].

π/2
0.6
(a) (b)
20 20

15 15
y (µm)

y (µm)

z (µm)
Phase

10 10

5 5

0 -π/2 0 0.0
0 10 20 0 10 20
x (µm) x (µm)

Figure 34: Interferometric dimple profile measurement. Observing interference


fringes allows to extract the phase (a) and to reconstruct the surface profile (b) of
the dimple fabricated onto the fiber facet.

image which allow for reconstruction of the surface shape. Figure 34


shows an exemplary phase profile obtained in this way along with
the reconstructed height profile.
The height data is then fitted with two-dimensional Gaussian profiles
to extract the depth and diameter of the dimple. In Fig. 35 (a)
we show this procedure resulting in a shape that is in excellent
agreement with Gaussian fits (black dashed lines). The extracted
width is 7.4 µm, depth 500 nm, and resulting radius of curvature

37
3. Fiber Cavity

0
100
7.4 µm
75

R (µm)
−200
z (nm)

500 nm
50
−400
(a) 25 (b)

0 10 20 −0.2 0.0 0.2


x, y (µm) ∆x/D

Figure 35: Dimple profile analysis. (a) Linecuts through the apex of the dimple
along orthogonal axes (dashed lines in Fig. 34 with corresponding colors) together
with Gaussian fits (black dashed lines). A plane has been fitted to the height profile
in Fig. 34 (b) and subtracted from the data here. (b) Local radius of curvature
around the center
√ of the dimple. At the inflection point of the Gaussian profile at
∆x/D = 1/ 8 (black dashed lines), R diverges and light is no longer confined.

13.9 µm in the center. We also obtain a small deviation of the


dimple diameters along orthogonal axes (blue and orange lines in
Fig. 34 (b) and Fig. 35 (a)) of 0.8 %. This is likely due to a residual
astigmatism of the focal position or a non-perpendicular alignment
of the fiber facet and the CO2 beam. Among different dimples,
the ellipticity ranges from close to zero up to a few per cent. In
Fig. 35 (b) the local radius of curvature of the dimples is displayed.
Away from its center, where the spherical approximation becomes
invalid, the radius increases due to the Gaussian shape and diverges
at the inflection point. This lack of confinement results in so-called
clipping losses degrading cavity lifetimes of modes with diameters
comparable to the dimple size and sets in for cavity lengths long
before the stability criterion is reached [112, 114]. Due to the small
feature diameters similar to the size of the fiber core (5 µm), the
CCD camera is also used to align the fiber core position and the
CO2 laser focus by detecting light coupled into the other end of the
fiber and emitted from the facet.

Table 31 compares typical parameters of different CO2 -based (Paris,


Munich, Basel, and this dissertation) as well as FIB-milled (Oxford)
dimples. The latter four show similar values and CO2 machining

38
3.2. Fiber fabrication

Setup D (µm) R (µm) t (µm)

Paris 20–60 35–200 0.5–3.0


Munich &4 &5 0.01–3.00
Basel 6–10 & 10 0.1–1.5
Zurich 8–13 7–30 0.2–1.4
Oxford (FIB) &3 & 5.6 0.2–0.7

Table 31: Overview of dimple parameters. Different dimple geometries achieved


with CO2 laser-based setups in Paris [113], Munich [115], Basel [116], this disserta-
tion, and FIB-based fabrication in Oxford [86, 109].

leads to comparable features sizes as FIB milling. However, only


FIB milling allows to create truly spherical shapes and thus avoids
clipping losses.
Another parameter affecting the cavity performance are scattering
losses
4πσsc 2
 
S≈ , (3.4)
λ
where σsc is the RMS surface roughness of the dimple [117]. In
Fig. 36 (a) we show atomic force microscope measurements of a
dimple with 898 nm depth and 13.9 µm diameter. To obtain σsc we
fit a fourth-order polynomial to the data in the center of the dimple
(white dashed box) and calculate the RMS of the residuals. We
obtain σsc = 0.18 nm leading to 7 ppm scattering losses at 850 nm
in good agreement with values reported in other CO2 -based setups
[112, 116]. The smoothing effect of the laser irradiation can be
observed at the edge of the fabricated dimple in Fig. 36 (b). After
removing the overall shape by again subtracting a two-dimensional
fourth-order polynomial fit, the residuals clearly show an increase of
variation towards the area where the CO2 beam did not evaporate
and melt the material.
After fabrication of the dimple, a reflective DBR coating is deposited
on the fiber facet to create the convex cavity mirror. Due to the
demand for high reflectivity, a batch of typically 20–100 processed
fibers is coated commercially4 with a Ta2 O5 / SiO2 DBR coating
4
Laseroptik GmbH

39
3. Fiber Cavity

25 12
(a) 1.5
24 (b)
20
23

∆z (nm)
15
y (µm)

y (µm)
z (µm)
10 22

5 21

0 0.0 0
0 10 20 16 18 20
x (µm) x (µm)

Figure 36: Dimple AFM measurement. (a) Surface profile of a machined fiber
measured with atomic force microscopy. In the center (white dashed box), the RMS
surface roughness after subtraction of a fourth-order polynomial is 0.18 nm. (b)
At the edge of the dimple (zoom-in of the red dashed box in (a)), the smoothing
effect caused by the CO2 laser irradiation is clearly visible in the residuals of the fit.

with 17.5 bilayers centered at 835 nm to allow for experiments with


InGaAs as well as GaAs QW excitons (850 nm and 814 nm emission
wavelengths at liquid helium temperature, respectively). To reduce
contaminations, the fibers are coated in an ultra-high vacuum.
This excludes the usage of standard optical fibers due to excessive
outgassing of the acrylic coating typically surrounding the fiber
cladding. Instead, we use fibers coated with a thin copper film5 .
The nominal transmission losses of the coating are T = 22 ppm at
850 nm with expected absorption losses below A = 10 ppm. The
nominal reflectivity of the coating is hence

R = 1 − T − A − S ≈ 99.996 %. (3.5)

3.3 Mode profile


3.3.1 Longitudinal
To gain insights into the mode profile and ultimately the confine-
ment of the photonic mode we start out by simulating the cavity
structure consisting of the semiconductor sample, air gap, and
fiber DBR, with transfer matrix calculation similar to Section 2.2.
5
Cu800, IVG Fiber Ltd

40
3.3. Mode profile

sc DBR air d DBR 1


1 (b)

Reflectivity
norm. |E|2

n
2

(a)
0 1 0
0 1 2 3 1.2 1.4 1.6
z (µm) Energy (eV)

Figure 37: Electric field and spectrum of a fiber cavity. The cavity structure
consists of the semiconductor GaAs / AlAs DBR with a λ/nGaAs -thick spacer layer
on top. After an air gap of variable length, the Ta2 O5 / SiO2 DBR forms the second
mirror of the cavity. The number of DBR pairs is 13 and 5.5, respectively. The
blue solid line (left axis) represents the electric field amplitude of the structure
with refractive index profile in orange (right axis). The green line at the antinode
at 1.84 µm denotes the position of the QW. The components of the structure are
schematically shown above the plot. Sc DBR and d DBR denote the semiconductor
and dielectric DBR, respectively. (b) The reflectance spectrum features a resonance
dip similar to that of a microcavity but with reduced depth due to the asymmetric
mirror configuration.

This approach provides a good description even for systems with


transverse confinement such as our fiber cavity if the paraxial ap-
proximation is valid. Figure 37 (a) shows that the electric field is
mostly localized in the spacer layer and air gap between the sample
and fiber surface. To ensure optimal coupling between the exciton
and photonic mode, the spacer layer is chosen to be λ/nGaAs thick.
This configuration minimizes the cavity length and allows to have
a field antinode in the center of the spacer layer where the QW
(green line at z = 1.84 µm) is located. Although the physical length
of the fiber cavity may well exceed that of a monolithic microcavity
due to the additional air gap, the low penetration depth into the
dielectric top DBR leads to an effective cavity length of only

 
λ
Leff = n2GaAs LDBR
AlAs/GaAs + + L + LDBR
Ta2 O5 /SiO2
nGaAs (3.6)
≈ L + 6.9 µm

41
3. Fiber Cavity

at 850 nm where L is the size of the air gap in between sample and
fiber DBR. For small gaps, Leff is shorter than that of monolithic
devices calculated by Eq. (2.11) [48] and we hence expect that the
polariton normal mode splitting in our system is equal to or ex-
ceeds that of pure semiconductor structures. Moreover, the electric
field node at the fiber DBR reduces the sensitivity to scatterers
accidentally deposited there during handling and operation of the
system. The reflectivity spectrum in Fig. 37 (b) shows a resonance
dip similar to that of a monolithic system. Even on resonance
there is yet a finite reflection and hence imperfect transmission.
This is due to the asymmetric configuration of the system with
non-identical mirror reflectivities leading to a reduced incoupling
efficiency when the system is excited with resonant light.

3.3.2 Transversal
Transfer matrix calculations accurately describe the longitudinal
electric field profile for two-dimensional systems. However, we
are particularly interested in the transverse mode profile to assess
the polariton confinement and therefore employ finite-difference
time-domain (FDTD) simulations [118] with a commercial-grade
simulator6 . This technique not only allows to investigate the dy-
namics and in particular lifetime of the cavity, but also to use
the actual geometry of the fiber DBR measured via interferometry
in the previous section. We use exemplary dimple parameters of
t = 850 nm and D = 10 µm together with the number of bilayers
of the structure used experimentally (35 bilayers of AlAs / GaAs
and 17.5 bilayers Ta2 O5 / SiO2 ) and simulate the electric field decay
as well as transverse mode profile upon narrow-band excitation of
the fundamental cavity mode. The quality factor is displayed in
Fig. 38 (a) and well exceeds 106 for short cavity lengths. During the
total simulation time of 12 ps chosen here, the electric field has not
yet decayed significantly in these cases. The employed technique
to estimate Q by demodulating the electric field dynamics and
fitting an exponential decay to the resulting envelope hence leads
to strong variations. Note that absorption and scattering losses
6
FDTD Solutions, Lumerical Inc.

42
3.3. Mode profile

2 4
(a) (c)
0 2000
3 |E|2
Q (106 )

1
2

y (µm)
0 1
(b) L
3
0
w0 (µm)

2
−1
1

0 −2
0 5 10 −5 0 5
L (µm) x (µm)

Figure 38: FDTD simulations. Numerically obtained cavity quality factor (a)
and beam waist (b) for the fundamental TEM(0,0) mode of a fiber with dimple
depth 850 nm and diameter 10 µm. The orange line is a fit to the data. For the
data points highlighted by green stars, the intensity profile is depicted in (c) along
with the surface of the fiber DBR (top dashed white line) as well as the interfaces
air-spacer layer and spacer layer-bottom DBR (bottom dashed white lines).

are neglected here and will further degrade the maximum quality
factor. When the effective cavity length Leff ≈ L + 6.9 µm becomes
comparable to the radius of curvature of the dimple R = 14.7 µm,
the cavity lifetime drops quickly by about an order of magnitude
and decreases upon further increase of the cavity length. This
regime corresponds to an instable Fabry-Pérot cavity where the
light rays in a geometric picture are reflected outside of the cavity
mode. In principle, we seek to minimize the radius of curvature of
the structure in order to maximize the transverse mode confinement.
However, it is imperative that it still exceeds the effective cavity
length to prevent the emergence of losses due to instability. This
sets a lower bound on the achievable confinement in our cavity
geometry, which is mainly limited by the finite cavity length due to
penetration into the semiconductor DBR.

43
3. Fiber Cavity

In Fig. 38 (b), we show the beam waist at the QW position obtained


by fitting a Gaussian function to the mode profiles exemplarily
shown in Fig. 38 (c) for L = 1.05 µm. We fit the simulation results
with Eq. (3.1) taking into account the penetration depth into the
DBRs as well as the dimple depth. The radius of curvature obtained
by the fit yields RFDTD = 17.4 µm which exceeds the geometrically
determined one
σ2
Rgeom = = 14.7 µm. (3.7)
2t
This reflects the effect of the non-spherical shape of the dimple which
leads to an effectively larger radius of curvature felt by the mode and
reduces the confinement. Using the electric field distribution one
can also calculate the effective mode volume of the cavity according
to Eq. (2.7) which for the mode displayed in Fig. 38 (c) evaluates
to 1.4 λ3 .
Figure 39 (a) shows the in-plane momentum dispersion of a cav-
ity with the same dimple geometry as in Fig. 38. We clearly see
discretized momentum eigenstates due to the strong in-plane con-
finement with a large splitting between the lowest two eigenstates
of 9.5 meV. In Fig. 39 (b) linecuts display the momentum space
distribution of the four lowest-lying states. The numerical aper-
ture of the fundamental mode, defined via a decay to 1/e2 of the
maximum intensity, is 0.16.

3.3.3 Beam waist measurement


We now wish to verify the strong confinement obtained in the FDTD
simulations experimentally. To this end, we employ a technique
presented in Ref. [116] for a similar purpose: Thin gold patterns with
25 nm thickness are fabricated onto a quartz substrate. Due to the
high resolution achievable with electron beam lift-off lithography,
the edges of these patterns are assumed to be sharp on optical
wavelength scales. We then scan the cavity mode laterally across the
structures while monitoring the mode linewidth under broadband
excitation in transmission. The effective reflectivity for the cavity
mode is then given by

Reff = αRgold + (1 − α)Rquartz (3.8)

44
3.3. Mode profile

1
(a) (b)
1.50 0 0.5 1
|E| (a.u.)
Energy (eV)

|E|2 (a.u.)
1.48

1.46

0
−4 −2 0 2 4 −4 −2 0 2 4
kk (µm-1 ) kk (µm-1 )

Figure 39: Numerical aperture. (a) Simulated cavity dispersion for a low-Q cavity
with 15 DBR bilayers for the sample and fiber DBR. The dimple geometry is identical
to Fig. 38. Multiple discrete states corresponding to different transverse modes are
visible. The strong variations in field amplitude are due to the initial conditions of
the simulation which favor some modes due to larger overlap. The energy splitting
between the two lowest-lying states is 9.5 meV. (b) Momentum-space distribution
of the cavity eigenmodes (linecuts at colored arrows in (a)). Each curve has been
normalized to its area. The numerical aperture of the fundamental mode is 0.16.

with Rgold and Rquartz being the reflectivities of the thin gold layers
and the quartz substrate, respectively. The coefficient

" √ !#
1 2x
α(x) = 1 + erf , (3.9)
2 w0

where erf is the error function, accounts for the overlap of the
Gaussian mode with this edge determined by the mode beam waist
w0 . For example, the cavity finesse drops quickly as the edge is
crossed for a strongly confined mode whereas a spatially extended
decay is expected for a large beam waist. A schematic of this
experiment and its results are depicted in Fig. 310 (a) and (b),
respectively. When the decay of cavity finesse is fitted with the
model in Eq. (3.8) we obtain w0 = 1.01 µm in close agreement with
FDTD simulations for the fiber used in this experiment yielding
w0,FDTD = 1.09 µm for short cavity lengths.

45
3. Fiber Cavity

30
(a) (b)
Fiber
20

Finesse
10
25 nm gold
quartz 0
0 5 10
2w0 x (µm)

Figure 310: Beam waist measurement. (a) Schematic of the experiment. The
fiber is scanned across the edge of thin gold stripes deposited on a quartz substrate
while monitoring the cavity finesse. (b) Finesse measurement as a function of
position. The orange line is a fit to the data in blue dots.

3.4 Spectroscopy
To assess the performance of the hybrid cavities we employ energy-
and time-resolved measurement techniques to determine its quality
factor. In Fig. 311 (a) we show the transmission of a tunable
continuous wave (CW) laser whose wavelength is scanned across
the fundamental TEM(0,0) mode resonance with low input power.
Since here we use the semiconductor sample with a QW inside, the
cavity length is chosen such that the cavity energy is tuned far
away from the exciton resonance in order to probe the bare cavity
properties. In Fig. 311 (a) we observe two Lorentzian peaks with
full width half maximum (FWHM) linewidths 23 µeV and 19 µeV,
respectively, and a spacing of 180 µeV. The double peak structure is
caused by the breaking of azimuthal symmetry of the mode due to
the slightly elliptic dimple shape giving rise to two non-degenerate
linearly polarized modes. The small difference in linewidth might
stem from the larger mode diameter of the lower-energy mode which
in turn leads to a small increase in clipping losses. Both modes can
be individually excited and detected with high selectivity due to
their orthogonal polarization.

When the cavity mode is excited with laser pulses short compared
to its lifetime, time-resolved observation of the decay allows to

46
3.4. Spectroscopy

3
(a) 180 µeV 2 (b)
Counts (106 )

Counts (104 )
2
25 ps
19 µeV 1
1

0 0
1.4572 1.4574 100 150 200
E (eV) t (ps)

Figure 311: Cavity spectroscopy. (a) Blue dots: Resonant scan of the excitation
energy across the cavity mode at 9.2 nW input power. Blue line: Double Lorentzian
fit to the data. The extracted linewidths (FWHM) for the two modes are 23 µeV and
19 µeV, respectively, with a mode splitting of 180 µeV. (b) Blue dots: Ring-down
measurement of the cavity with circularly polarized excitation. Blue shaded area:
Exponentially decaying oscillatory fit to the data with decay time 37 ps and period
25 ps.

extract the lifetime. We send short circularly polarized pulses from


a mode-locked Ti:sapphire laser7 with a repetition rate of 76.3 MHz
and pulse durations of 3 ps to 5 ps to excite the cavity. After leaving
it, the light is directed to a streak camera8 that allows for measuring
dynamics with a resolution of a few ps. In Fig. 311 (b) we show the
cavity decay time trace. Since the spectral width of the excitation
pulse is broader than the splitting of the two linearly polarized
modes, the excitation with circular polarization is no longer an
eigenbasis of the system and leads to a beating between the two
modes. We extract a beating period of 25 ps and decay time of
37 ps from an exponentially decaying oscillatory fit to the data.

While the cavity lifetime extracted from the temporal measurement


is directly linked to the homogeneous linewidth, the response of
the spectral measurement could be inhomogeneously broadened e.g.
due to cavity length fluctuations. This is indeed reflected by the
small discrepancy between the two measurement techniques leading
to temporally and spectrally measured quality factors of 80 300 ± 80
7
MIRA 900, Coherent Inc.
8
C5680, Hamamatsu Photonics K. K.

47
3. Fiber Cavity

and 78 327 ± 23, respectively. The spectral line is broadened by


about 10 % of its homogeneous linewidth due to acoustic noise
changing the cavity length. The finesse values obtained from these
experiments lie between 4 000 to 5 000 and are lower than the dielec-
tric coating would suggest according to Eq. (3.5). We thus perform
control experiments with an empty cavity with identical dielectric
DBRs both on the fiber and a planar glass substrate. In this case,
the extracted cavity lifetimes and finesse values of 450 ps and 50 000,
respectively, exceed those with semiconductor DBR by about an
order of magnitude [119], and demonstrate that the limiting element
for the cavity lifetime are losses in the semiconductor DBR. Note
that this configuration with vastly different mirror reflectivities also
reduces the incoupling efficiency as indicated in Fig. 37.
Figure 312 displays the cavity emission profile imaged with a CCD
camera when the CW excitation is kept at a fixed wavelength
and the cavity length is swept. A multitude of modes with differ-
ent transverse profiles successively appears whenever their energy
matches that of the excitation. These Hermite-Gaussian modes
reflect the breaking of azimuthal symmetry along two orthogonal
axes in analogy to the polarization splitting of the fundamental
mode. For the TEM(2,2) and also following higher-order modes
(not shown), however, the emission profile deviates from the ex-
pected one. This is due to a more complex interplay of mode mixing
caused by increasing clipping losses for modes with larger spatial
extent that lead to increasing linewidths. Once these linewidths well
exceed the energy of the polarization splitting, we expect Laguerre-
Gaussian mode profiles since the breaking of azimuthal symmetry
is washed out by the energy uncertainty.

3.5 Noise spectrum


Due to its open geometry, the two mirrors of the fiber cavity may
fluctuate relative to one another, effectively causing a changing
cavity length and leading to a small amount of inhomogeneous
broadening already observed in the previous section. In order
to characterize this contribution in more detail, we excite the
cavity mode on resonance and with a laser detuning of ∆/γ =

48
3.5. Noise spectrum

TEM(0,0) TEM(1,0) TEM(2,0)

TEM(0,1) TEM(1,1) TEM(2,1)

TEM(0,2) TEM(1,2) TEM(2,2)

Figure 312: Transverse modes. The energy of the excitation is kept constant
while sweeping the cavity length and imaging the mode onto a CCD. A multitude
of Hermite-Gaussian modes is visible.

−1/2, respectively, while monitoring the transmitted photon counts


with a sampling rate of 1 kHz. On resonance, the detected signal
is only second-order sensitive to acoustic noise corresponding to
spectral shifts of the mode, but linearly sensitive to e.g. laser
and electric noise. On the slope, both the acoustic and other
noise sources dependence is linear such that comparing the two
cases allows to extract the acoustic contribution. Measuring far
off resonance indicates the baseline of the noise spectroscopy due
to e.g. electric noise of the detector [116]. This is displayed in
Fig. 313. Comparing the orange and green curve (on the slope
and on resonance, respectively) we identify multiple resonances
below 250 Hz with a total RMS acoustic noise amplitude of 2.9 pm
obtained by subtracting the integrals of the two curves. Relating
this value to the mode linewidth is in good agreement with the

49
3. Fiber Cavity

∆/γ
10−1 -1

PSD (pm2 /Hz)


-0.5
0
10−3

10−5
0 100 200 300 400 500
Frequency (Hz)

Figure 313: Cavity noise spectrum. Power spectral density (PSD) of the cavity
noise with excitation laser detuning far off resonance (blue line), ∆/γ = -0.5 (orange
line), and ∆/γ = 0 (green line). The amplitude is calibrated by the slope of the
resonance at the linear part of the Lorentzian spectrum and indicates spectral shifts
of the mode. Several acoustic noise resonances are located in the frequency range
below 250 Hz.

inhomogeneous broadening identified by comparing to the lifetime


measurement in the previous section.

3.6 Strong coupling regime


After a careful characterization of the cavity properties we now
wish to study its coupling to excitons located in the InGaAs QW.
To this end, we excite the system with broadband CW light and
monitor the transmission spectrum with a grating spectrometer9
while sweeping the cavity length and hence tuning the cavity en-
ergy. The results for a sweep range of about 2 µm are displayed in
Fig. 314 (a). Three sets of TEM modes with different longitudinal
mode numbers are visible whose energy changes as a function of
nanopositioner steps. The relative intensity in each set is highest
for the fundamental TEM(0,0) mode and decreases with increasing
transverse mode number, with even mode numbers generally slightly
brighter. This can be understood by invoking the overlap integral
of the respective mode with the Gaussian profile for excitation and
detection each provided by coupling out of and into a single-mode
9
SpectraPro 2750, Princeton Instruments Inc.

50
3.6. Strong coupling regime

E (eV)
1.38 1.40 1.42 1.44 1.46 1.48 1.50 1.52
(a)
100

75
Steps

50 105 Counts (s-1 )

25 103

0 101

60 (b) EX
Voltage (V)

40
105
Counts (s-1 )

20 103

101 EC
0
1.460 1.465 1.470 1.475 1.480 1.485
E (eV)

Figure 314: Cavity transmission spectra. The cavity length is swept via stepping
(a) or changing the dc voltage (b) of the nanopositioner. Note the small discrepancy
in energy due to the different wavelength calibrations of the spectrometer gratings
used. In (a), three sets of transverse modes strongly couple to different excitations in
the system (colored arrows, see main text). (b) Close-up around 1.47 eV highlighting
the strong coupling between the cavity and quantum well exciton with 3.50 meV
normal mode splitting. The white dashed lines denote the bare exciton and cavity
energy.

fiber. For a system without any relative displacement between


dimple and fiber core we would expect the emission from odd mode
numbers to be strongly suppressed due to their mode profile being
orthogonal to the excitation. Moreover, several windows featuring
anticrossings indicating the strong coupling of the cavity mode to
another mode can be identified. The magnitude of the couplings
increases for increasing step
√ size corresponding to decreasing cavity
length and reflecting the Leff dependence of the coupling predicted
by Eq. (2.15). Around 1.41 eV (orange arrow), a large anticrossing
of the confined modes and another resonance with a fixed energy

51
3. Fiber Cavity

occurs. This splitting, on the order of several meV, is likely due


to mixing of the 0D cavity modes with other optical modes of the
structure whose energy does not depend on the cavity length. Since
the corresponding wavelength is still well within the stopband, we
exclude coupling to the leaky modes of the DBR [120]. Instead, we
associate this resonance with the 2D cavity mode in the spacer layer,
confined by the semiconductor DBR and the large refractive index
contrast between GaAs and air [121]. At 1.47 eV (green arrow), an
anticrossing arises from the coherent coupling between the cavity
mode and the collective excitation of QW excitons. Figure 314 (b)
displays a close-up of that region for the shortest cavity length
achievable, i.e. the fiber is in physical contact with the semicon-
ductor sample. Here, we use the application of a dc voltage to the
piezo crystal in the nanopositioner to change the cavity energy. The
extracted width of the normal mode splitting is 3.50 meV. Note
that this is slightly larger than typical normal mode splittings of
monolithic microcavity structures [43] due to the reduced DBR
penetration depth described in Section 2.2. At 1.50 eV (red arrow),
the mode exhibits a small anticrossing that might be attributed
to coupling to 13 C impurities deposited during the growth of the
sample. The last anticrossing at 1.51 eV (purple arrow) stems from
strong coupling of the cavity mode to excitations of bulk GaAs exci-
tons in the spacer layer. The splitting of the fundamental TEM(0,0)
mode and the next TEM(1,0) / TEM(0,1) modes of 11.8 meV again
demonstrates the strong confinement generated by the fiber cavity
in good agreement with the numerical simulations in Fig. 39.

3.7 Conclusion and perspectives


In summary, we have seen how CO2 laser-based fabrication of a
fiber cavity allows to create cavities capable of strongly confining
the photonic mode while at the same time exhibiting long cavity
lifetimes. The approach thus overcomes the trade-off between
these properties typical for other confinement techniques. Due
to the open geometry, the cavity is subject to a small amount
of acoustic noise inhomogeneously broadening the mode. At the
same time, this provides a handle to tune the cavity energy in situ.

52
3.7. Conclusion and perspectives

Together with the usage of a transmission geometry, the fiber cavity


setup fulfills all three requirements for a system suited to observe
quantum-correlated polariton emission that we started out with in
this chapter.
The main direction for further improvement of the cavity perfor-
mance is the prolongation of the cavity lifetime. The effective cavity
length, mainly determined by the finite penetration depth of the
optical mode into the DBRs for short cavity lengths, sets a lower
bound on the dimple’s radius of curvature. Since the cavity already
operates close to the stability regime with R & L, and due to the
weak dependence of the beam waist on the radius of curvature
in Eq. (3.1), we do not expect a significant improvement of the
confinement. The measured cavity lifetimes, however, seem to lag
behind considering the reflectivities possible with dielectric DBRs.
While polariton lifetimes of several hundred picoseconds have been
reported in semiconductor microcavities [68, 122], comparing the
lifetimes measured here with those where the GaAs / AlAs DBR
is replaced by a dielectric one clearly suggests that the main lim-
itation is the semiconductor DBR. Note that all dielectric DBRs
used here stem from the same coating run and exhibit a similar
behavior among different fibers. In earlier experiments, the fiber
cavity was used with coatings centered at 900 nm [48, 104]. There,
the cavity lifetime exceeded the one observed here, suggesting that
longer wavelengths are beneficial and e.g. reduce absorption by
the GaAs spacer layer or GaAs / AlAs DBR. In order to shift the
QW emission to longer wavelengths accordingly, the indium content
in these samples was higher. This in turn lead to an increase in
lattice strain between the InGaAs layer and surrounding GaAs.
The resulting strong inhomogeneous broadening of the polariton
emission for a sizable exciton content hence makes operating at
longer wavelengths not a promising strategy.
Ideally, the QW should be placed in between two dielectric DBRs,
replacing the semiconductor one altogether. In Ref. [123], epitaxial
lift-off was introduced as a technique to achieve this: The QW
inside of the spacer layer is epitaxially grown on top of a sacrificial
layer. Upon etching this layer, the thin spacer layer membrane is
transferred onto a dielectric DBR to which it bonds via van der

53
3. Fiber Cavity

Waals forces. Although this technique requires a few additional


processing steps, we expect it to significantly improve the polariton
lifetime by using the high-quality dielectric coatings to their full
potential.

54
Chapter 4

Dissipative Phase Transition

This chapter discusses the physics of a dissipative phase


transition that occurs in a driven polariton system. After
introducing the theoretical description and the phenomenon
of optical bistability, we present measurements in two regimes
– close to and far away from thermodynamic limit – in
Sections 4.5 and 4.6. The main findings of this chapter are
published in Ref. [124].

4.1 Master equation of open quantum systems


For systems at equilibrium and neglecting any kind of dissipation
to its environment, the evolution of such a so-called closed system
is completely described by the corresponding Hamiltonian. For
example, losses can be safely neglected in the analysis of atomic
spectra or calculation of harmonic oscillator eigenstates. In quan-
tum optics, however, losses and specifically dissipation from lossy
cavities play a central role already in one of the most fundamental
phenomenon of quantum optics, which is that of lasing action. In
fact, dissipation can give rise to interesting and novel phenomena
such as inhibiting particle losses and inducing strong correlations
in a cold atom gas [125, 126] or nonequilibrium phase transitions
[127]. This emphasizes the necessity to go beyond the Hamiltonian

55
4. Dissipative Phase Transition

closed-system description and incorporate losses into the analysis.


We do so by considering the coupling of the system under study to
an infinite number of environment modes. In this scenario, informa-
tion can be dissipated from the system to the environment, yet the
reciprocal process is highly unlikely. Tracing out the reservoir de-
grees of freedom then discards the correlations between system and
reservoir making the dynamics non-unitary. Three approximations
are commonly made to simplify the corresponding Liouville-von
Neumann equation:
• Born approximation: Correlations between the system and
bath degrees of freedom are neglected, which means that the
total density operator can be written as a tensor product of
the two at all times.
• Markov approximation: A memoryless reservoir is as-
sumed. This means that temporal correlations in the bath
degrees of freedom are neglected, i.e. its density operator at
time t does not depend on the prior evolution.
• Secular approximation: Similar to the rotating wave ap-
proximation, fast oscillating terms (i.e. faster than the optical
transition frequency) in the master equation are neglected
since they will average out on timescales of interest.
Applying the Born–, Markov–, and secular approximation, we
obtain the following master equation in Lindblad form to describe
the dynamics of a driven-dissipative system [128]
∂ ρ̂
= L̂ρ̂. (4.1)
∂t
Here, ρ̂ is the system’s density matrix and
ih i
L̂ = − Ĥ, ρ̂
~
γ (nth + 1)  
(4.2)
+ 2âρ̂↠− ↠âρ̂ − ρ̂↠â
2
γnth  † 
+ 2â ρ̂â − â↠ρ̂ − ρ̂ââ†
2
is the Liouvillian superoperator describing its coherent dynamics
according to the corresponding Hamiltonian (first line) as well as

56
4.2. Bistability

dissipation with rate γ (second and third line). ↠and â are the
system’s raising and lowering operators, respectively. The thermal
occupation number nth accounts for the finite temperature of the
bath and associated thermal excitation and deexcitation processes.
For exciton-polaritons considered in this dissertation, however, the
thermal energy of kB × 4.2 K ≈ 0.4 meV is orders of magnitude
smaller than the optical transition energy of hc/850 nm ≈ 1.46 eV
such that thermal excitations are negligible and we can assume
a zero temperature bath with nth = 0 unless specified otherwise.
Note that for e.g. circuit QED systems with typically nth . 0.1
(assuming a transition frequency of 5 GHz at a temperature of
0.1 K) this condition is not automatically fulfilled due to their
lower transition frequencies in the microwave range and thermal
excitations should be taken into account. Steady state properties
can be obtained from Eqs. (4.1) and (4.2) by setting

L̂ρ̂ss = 0 (4.3)

and solving for the steady state density matrix ρ̂ss .

4.2 Bistability
Bistability describes the coexistence of two stable states under the
same external driving conditions. The general Hamiltonian for a
system exhibiting this phenomenon is
U † †
Ĥ = EP ↠â + â â ââ + F ∗ eitEL /~ ↠+ F e−itEL /~ â, (4.4)
2
where EP,L are the mode and excitation laser energy, respectively, U
denotes the nonlinearity, F is the excitation amplitude, and â and â†
are the system’s lowering and raising operators, respectively. Going
into a frame rotating with the excitation frequency ωL = EL /~
leads to
U
Ĥ = −∆↠â + ↠↠ââ + F ∗ ↠+ F â, (4.5)
2
where ∆ = EL − EP denotes the energy detuning between the
mode and the pump. Note that this is identical to the Hamiltonian
introduced in Eq. (2.30) to describe the exciton-polariton system

57
4. Dissipative Phase Transition

in a lower polariton basis with effective nonlinearity U weighted by


the exciton Hopfield coefficient. Bistability occurs when the system
is excited with a finite energy detuning ∆ > 0 (∆ < 0) for repulsive
(attractive) interactions: As the drive strength is increased, the
resulting nonlinearity-induced and population-dependent energy
shift pushes the mode energy closer to that of the excitation, effec-
tively decreasing the detuning. This in turn increases the overlap
of the excitation energy with the mode, allowing for the injection
of an even higher population. For a sufficiently strong critical drive
strength, this self-accelerating process leads to a locking of the
mode energy to that of the excitation accompanied by a sharp jump
in the mode population. When the drive strength is decreased, the
mode will at first remain locked to the excitation energy and only
drop to restore the finite detuning at lower powers. This leads to
the observation of a hysteresis cycle when the excitation power is
ramped up and down and constitutes a smoking gun experiment for
bistability. In Fig. 41 (a), this behavior is illustrated by the blue
solid lines. Bistability is a very general feature exhibited by driven-
dissipative nonlinear systems and has been observed experimentally
over the past four decades in a plethora of different physical systems
such as sodium-filled cavities [129], suspended mirrors [130], silica
microspheres [131], exciton-polaritons [132], integrated silicon chips
[133], photonic crystals [134], dynamic nuclear spin polarization of
self-assembled quantum dots [135], and surface plasmon-polaritons
[136] to name only a few. From an application point of view, bistable
systems can be used as switching elements or transistors with fast
optical response and low energy consumption in all-optical circuits
[137–140].

4.2.1 Mean-field description


In order to treat the often highly nontrivial dynamics arising from
the nonlinear term in Eq. (4.4), a standard approach is the mean-
field approximation. It is based on replacing quantum operators by
their respective mean-field value

â(r, t) → hâ(r, t)i = α(r, t), (4.6)

58
4.2. Bistability

20 (a) 5 (b)

15 3
Population

Bunching

Im [α]
0 W (α)
10
2 0.2
5 0.1
−5 0.0
0 1
0 2 4 6 −5 0 5
F /γ Re [α]

Figure 41: Principle of bistability. (a) Simulation of the cavity population |α|2
in the bistable regime with a mean-field treatment (blue lines, left axis) and
h↠âi in a quantum description (orange line, left axis). Nonlinearity and detuning
are U /γ = 0.2 and ∆/γ = 3, respectively. For a range of excitation strengths
1.9 < F /γ < 4.6, the semiclassical approach has three solutions out of which two
are stable and one metastable (dashed line). Experimentally, the population jumps
along the vertical blue arrows to the upper branch when the high-power turning
point at F /γ ∼ 4.6 is reached from below, and back to the lower branch once the
excitation crosses the lower turning point at F /γ ∼ 1.9 from above, effectively
establishing a hysteresis cycle. The quantum solution (orange) has only one unique
solution with a nonlinear jump of the mode population. The second-order correlation
function g (2) (0) (green, right axis) displays strong super-Poissonian statistics around
this nonlinear threshold. (b) Wigner phase space distribution in the bistable regime
at F /γ = 3.4 (black cross in (a)) illustrating the bimodality of the steady state.

where α = Tr {ρ̂â} is the classical field variable. This substitution


neglects quantum fluctuations and non-commutivity between oper-
ators. It is hence only exact for terms with at most two operators
such as e.g. the free cavity evolution as well as coherent pumping
and losses. For any nonlinear term, however, an approximation is
made
D E
↠(r, t)â(r, t)â(r, t) ≈ |hâ(r, t)i|2 hâ(r, t)i. (4.7)

Note that this approximation is always implicitly made when the


nonlinearity of a system is expressed via a nonlinear susceptibility
χ(n) of the classical field. Applying the mean-field approximation to
the exciton-polariton system in a polariton basis including coherent
drive and losses then leads to the driven-dissipative Gross-Pitaevskii
equation also referred to as nonlinear Schrödinger equation [141,

59
4. Dissipative Phase Transition

142]
h i
i∂t α(r, t) = ∆ − U |α(r, t)|2 α(r, t)
γ (4.8)
− i α(r, t) + F (r, t).
2
The first line describes the (relative) energy of the mode as well as
the nonlinear response of the system proportional to the particle
density np = |α|2 . Since it is of third order in the field operator, the
interaction between polaritons is often referred to as a Kerr-type
nonlinearity. The second line of Eq. (4.8) introduces losses from
the system with rate γ and a coherent drive with amplitude F .
Multiplying with the complex conjugate leads to
  γ 2 
np (∆ − U np ) + 2
− |F |2 = 0. (4.9)
2
It can be easily shown that this equation has three positive, real
solutions for np when √
3
∆> γ, (4.10)
2

i.e. when the excitation laser frequency ωL exceeds ωp + 23 γ.
One of the solutions, however, is unstable. These solutions are
illustrated by the blue lines in Fig. 41 (a). In this bistable regime,
three solutions can be observed out of which two are stable (solid
lines, top and bottom)
√ whereas the third (dashed line, middle) is
not. When ∆/γ < 3/2, Eq. (4.9) admits only a single, stable
solution and the system is in the monostable regime.
The data displayed in Fig. 42 applies these concepts to our exciton-
polariton fiber cavity system and is obtained by experimentally
ramping the excitation power up and down at different detunings
while monitoring the transmitted photons. Three different regimes
are observable:
• ∆ < 0 : When the excitation is red-detuned with respect to
the mode (red curve), we obtain a unique sublinear power
dependence after an initially linear onset heralding the optical
limiter regime. Here, the injection of further excitations is
partly suppressed due to the increasing blueshift of the mode
with each additional polariton.

60
4.2. Bistability

(a) 1.0 (b)


Counts (104 ) 2

Counts (106 )
1 0.5

0 0.0
−5.0 −2.5 0.0 2.5 5.0 2 4 6
∆/γ Power (mW)

Figure 42: Nonlinear power dependence. (a) Polariton resonance with Lorentzian
fit (blue dots and blue solid line, respectively). (b) Input power ramps (first up,
then down) for laser detunings indicated by the colored arrows in (a). The orange,
green, and red curve demonstrate the bistable, monostable, and optical delimiter
regime, respectively.


• 0 < ∆/γ < 3/2 : For small positive detunings (green curve)
the transmitted photon numbers for increasing and decreasing
powers are identical. The nonlinear increase at a threshold
power indicates the monostable regime where the mode locks
to the pump energy but without exhibiting hysteresis.

• ∆/γ > 3/2 : For larger blue detunings (orange curve), a
hysteresis cycle is observed and the photon numbers for in-
creasing and decreasing excitation power only coincide outside
of this window. The system is in the bistable regime and the
hysteresis cycle spans the two stable solutions.

4.2.2 Quantum description


In a quantum description, the electric field observables are treated
as true quantum mechanical operators with nonzero commutators.
The bistable problem given by Eqs. (4.1) and (4.4) then only admits
a unique solution in stark contrast to the mean-field treatment [143,
144]. This apparent contradiction can be reconciled by invoking
the role of quantum fluctuations. In the mean-field treatment,
these were neglected resulting in an infinite lifetime of the two
stable solutions. In a quantum treatment, however, fluctuations can
trigger switching between the two branches and the steady state

61
4. Dissipative Phase Transition

response of the system illustrated as orange solid line in Fig. 41


(a) is an average weighted by the residence time in each of the
two branches. This can also be illustrated by the Wigner function
representation of the steady state
Z
2 −2|α|2 n o ∗
d2 γ Tr ρ̂ss eγâ −γ â eγ α−γα . (4.11)
† ∗ ∗
W (α) = e
π
Figure 41 (b) displays this phase space distribution at the nonlinear
threshold in the bistable regime. It features two well-separated
contributions underlining the bimodality of the steady state [145,
146].
The quantum fluctuation-induced switching behavior of the driven-
dissipative nonlinear system naturally raises the question what
the relevant timescales for these processes are. Analogous to the
eigenspectrum of a Hamiltonian, we write an eigenvalue equation
for the Liouvillian superoperator

Lρ = λρ ρ. (4.12)

Here we dropped the hats denoting operators and the Liouvillian


superoperator is recast as a matrix acting on the density matrix
vector ρ with eigenvalue λρ . Since dissipation leads to non-unitary
dynamics of the system, i.e. an initially pure state may be mapped
onto a mixed state under the action of the Liouvillian in Eqs. (4.1)
and (4.2), the eigenvalues λρ are complex in contrast to the real
eigenvalues of a closed-system Hamiltonian. The real part of the
Liouvillian eigenvalues corresponds to a rate and obeys Re [λρ ] ≤ 0,
ensuring that all higher-lying excited eigenstates eventually decay
towards the steady state with eigenvalue Re [λss ] = 0. The imagi-
nary part, on the other hand, expresses the energy associated with
the respective excitation. As an example, Fig. 43 shows the real and
imaginary parts of the Liouvillian eigenspectrum for the nonlinear
Hamiltonian described in Eq. (4.5) with ∆/γ = 1.5 and U /γ = 0.05,
i.e. in the mean-field bistable regime, as a function of drive strength.
As expected, we find one eigenstate with zero real and imaginary
part for all drive strengths which corresponds to the system’s steady
state ρ̂ss . For low powers, the Liouvillian excited state-decay rates
appear as a ladder of twofold degenerate pairs with a spacing of γ/2.

62
4.2. Bistability

−Re [λρ /γ], norm. h↠âi


1.0
1

Im [λρ /γ]
0.5 0

−1
(a) (b)
0.0
0 2 4 0 2 4
F /γ F /γ

Figure 43: Liouvillian eigenspectrum. (a) Real part of the five lowest-lying
Liouvillian eigenvalues of the driven-dissipative nonlinear system with ∆/γ = 1.5
and U /γ = 0.05 (blue and orange solid lines). The values for a linear system are
denoted by black dashed lines. The gap between the steady state with eigenvalue
zero and first excited-state eigenvalue is reduced around the nonlinear threshold
(orange line) corresponding to a vanishing decay rate. Note that also the second
excited-state eigenvalue is slightly reduced. The green dashed line denotes the
normalized population. (b) Imaginary part of the steady state (blue line) and first
excited state (orange line).

The trace of the steady state ρ̂ss is conserved with Tr {ρ̂ss } = 1 since
the trace of the dissipative, non-unitary part of Eq. (4.2) acting
on the density operator is zero. Note, however, that the other
eigenstates should not be considered as true density operators since
their traces vanish. Yet, if the system is perturbed out of its steady
state, the resulting density operator may be written as a mixture
of the steady state and higher-lying Liouvillian excited states. In
particular, the admixture of the latter components does not violate
the normalization due to their vanishing trace. This is observed
around the nonlinear threshold, i.e. where the mode population
(green dashed line) exhibits a superlinear kink. Here, the behavior
of Liouvillian eigenspectrum changes drastically compared to low
excitation strengths: The gap between the steady state and first
excited state highlighted in orange is strongly reduced and the
imaginary parts become degenerate rendering the first excited state
a soft mode [147, 148]. For a vanishing gap, the steady state density
matrix is a mixture of two eigenvectors ρ̂0,1 with individual weights

63
4. Dissipative Phase Transition

p and 1 − p, respectively

ρ̂ss = pρ̂0 + (1 − p)ρ̂1 . (4.13)

This is the regime where the quantum and mean-field description


coincide: The mixture of the steady state composed of a low- and a
high-intensity contribution already observed in the bimodal phase
space distribution mirrors the coexistence of the two stable branches
in the mean-field description. Due to the strong reduction of the
Liouvillian gap, the associated timescales diverge. They correspond
to the switching times or lifetimes of the two stable branches in the
mean-field description and can exceed the intrinsic decay rate by
many orders of magnitude [145, 149, 150]. The excitation strength
can thus be swept non-adiabatically, i.e. sufficiently fast to prevent
switching between the two states, which is a necessary criterion
to observe the hysteresis cycle in Fig. 42 and the reason for the
successful description of bistability by mean-field theory.

4.3 Dissipative quantum phase transition


The closing of the Liouvillian gap is closely related to the closing of
the Hamiltonian gap in a closed system typically heralding a phase
transition. We hence briefly discuss quantum phase transitions
to then make the connection to the driven-dissipative Kerr-type
nonlinear polariton system considered here. Generally speaking,
phase transitions describe a sudden and qualitative change of system
properties when an external parameter is varied. In classical systems,
this parameter is the temperature and thermal fluctuations lead
to breaking of a symmetry when the phase transition is crossed,
accompanied by a change in the structural, electronic, or magnetic
properties of a solid state system. Quantum phase transitions, on
the other hand, occur even at zero temperature upon changing of
a parameter in the Hamiltonian. The transition is triggered by
quantum fluctuations and does not rely on thermal fluctuations
[151]. Quantum phase transitions are a highly active field of research
in condensed matter physics due to their associated intriguing many-
body properties in e.g. rare-earth magnetic insulators [152], heavy-
fermion compounds [153, 154], two-dimensional electron gases [155,

64
4.3. Dissipative quantum phase transition

(a)

∆(g)
O(g)

(b)

1 1
g/gc g/gc

Figure 44: Continuous phase transition. At the critical point g = gc , the order
parameter (a) describing the symmetry breaking is non-analytic and the energy gap
of the system (b) vanishes.

156], and high-temperature superconductors [4, 5, 157] – the latter


in particular also being of great interest for applications.

Phase transitions are generally divided into first-order and contin-


uous phase transitions. In the former, the two phases can coexist
at the transition point, e.g. ice and water at 0 ◦C. When a system
undergoes a continuous phase transition, however, the two phases
do not occur at the same time. This can be expressed via an order
parameter O(g) which describes the symmetry that is broken when
the system undergoes the transition. As illustrated in Fig. 44 (a),
the order parameter vanishes in one of the phases and is non-zero
in the other with a singular behavior at the critical point gc . As an
example, the order parameter in the Mott insulator to superfluid
phase transition is coherence, which is finite in the superfluid phase
but vanishes in the insulating phase [158]. Moreover, also the ex-
citation gap between the ground state and the first excited state
vanishes at the critical point illustrated schematically in Fig. 44 (b).
Continuous phase transitions exhibit a divergence of correlation
length and time as the critical point is approached leading to a
power-law behavior of all observables with universal exponents in
its vicinity [151, 159].

In dissipative systems, the role of the vanishing energy gap between


ground and excited is taken over by the gap between the Liouvillian

65
4. Dissipative Phase Transition

steady state and higher-lying eigenvalues −Re [λρ ]. When this gap
vanishes, the excited mode softens and the associated timescales
diverge. The relaxation to the steady state is then critically slowed
down giving rise to a DPT [160, 161]. The bistable system char-
acterized by quantum fluctuation-induced switching between the
two branches with vanishing rates at a critical drive considered
in this chapter is closely related to early experiments on the dy-
namics of two-mode lasers. There, random switching events with
diverging switching times between a low- and a high-intensity state
have been found [162, 163]. While increasing the pump strength,
the probability of the high-intensity state increased while that of
the low-intensity state decreased implying a discontinuous jump of
the most probable light intensity. These results have hence been
interpreted as a first-order quantum phase transition with the light
intensity as order parameter [164]. The analogy to systems exhibit-
ing bistability is now straightforward. Hence, these systems have
also been predicted to exhibit critical slowing down as a function
of the drive strength signaling a first-order DPT [161, 165].

4.4 Measuring the Liouvillian eigenspectrum


with correlations
To observe phase transitions in driven-dissipative systems, differ-
ent techniques have been employed. Measuring the phase space
distribution in circuit QED experiments allowed to observe the
bimodality of the steady state when the system undergoes the pho-
ton blockade breakdown phase transition [166]. Residence time
measurements have been used to analyze the timescale over which
the system returns to the steady state upon a perturbation provided
the dynamics are slow enough to be resolved experimentally [67,
167]. Here a different technique based on correlations of individual
photons is introduced. To illustrate the principle, let us consider
the case where the nonlinear system is in the bistable regime and
a continuous drive is applied, and suppose it starts out in the
low-intensity state. Fluctuations of quantum or classical origin
can then trigger switching to the high-intensity state leading to a
jump in the transmitted light intensity, which eventually returns

66
4.4. Measuring the Liouvillian eigenspectrum with correlations

to its steady state value. The intermittent light bursts give rise to
strong super-Poissonian photon statistics illustrated in Fig. 41 (a).
The timescale over which the bunching persists is determined by
the inverse Liouvillian gap, i.e. the relaxation time of the system
back to its steady state [168]. In a quantum mechanical picture,
the collapse operator â, i.e. detecting a photon dissipated from
the cavity, projects the system out of its steady state ρ̂ss into a
mixture of Liouvillian eigenstates. The second photon detection,
conditioned on the first, then measures the relaxation dynamics
back to ρ̂ss . In this sense, photon correlations provide an analysis
tool to study the Liouvillian spectrum and measure the dissipative
counterpart of Hamiltonian quench dynamics [169, 170]. Although
we will apply this measurement technique to a driven-dissipative
nonlinear polariton system, it constitutes a general tool to study
the Liouvillian eigenspectrum of driven-dissipative systems and a
few comments on its applicability are due. If we consider a general
collapse operator ĉ acting on the system, we require that it exhibits
super-Poissonian statistics in order to measure its timescale and
extract the Liouvillian gap. Note that, although in our case ĉ = â,
the technique is not restricted to optical photon detection. In order
to obtain bunching, we require that the density matrix the system
is projected into after the collapse, ρ̂p , and the steady state differ
in the expectation value of the observable. Formally, this can be
written as n o n o
Tr ĉ† ĉρ̂p > Tr ĉ† ĉρ̂ss . (4.14)
Although it is not clear for which range of systems this is valid
warranting further research, Eq. (4.14) is generally fulfilled at and
around phase transitions where the divergence of fluctuations alters
the statistics of the associated observable opening up the application
to a wide range of physical systems.
Figure 45 illustrates the correspondence between photon bunching
and reduction of the Liouvillian gap, i.e. slowing down of dynamics,
for a range of excitation detunings and nonlinearities. The bunch-
ing amplitude is peaked in the region where the Liouvillian gap is
smallest. Moreover, it increases strongly with increasing detuning
which can be understood from the increased contrast between the
low- and high-intensity state populations for larger detunings in a

67
4. Dissipative Phase Transition

∆/γ

1.5
2.0

−Re [λρ /γ]


1.0
g (2) (0)

1.5
0.5
1.0
0.0
1.5
2.0

−Re [λρ /γ]


1.0
g (2) (0)

1.5
0.5
1.0
0.0
U /γ

1.5
2.0

−Re [λρ /γ]


1.0
g (2) (0)

1.5
0.5
1.0
0.0
1.5
2.0

−Re [λρ /γ]


1.0
g (2) (0)

1.5
0.5
1.0
0.0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
F /γ F /γ F /γ

Figure 45: Photon bunching and Liouvillian eigenspectrum. Simulated second-


order photon correlation function (blue solid lines, left axis) and decay rates of the
five lowest-lying Liouvillian eigenvalues (orange solid lines, right axis) as a function
of pump strength. The detuning is ∆/γ = 0.5, 1.0, and 1.5 (left to right) and
nonlinearity U /γ = 0.05, 0.1, 0.2, and 0.5 (top to bottom). The blue dashed lines
denote coherent photon statistics and the orange dashed lines correspond to the
eigenvalues of a linear system. The Hilbert space has been truncated at the N = 90
manifold.

68
4.5. Dissipative phase transition in the thermodynamic limit

mean-field description. The degree to which the Liouvillian gap is


reduced depends on the number of excitations. For large nonlineari-
ties, the number of particles required to trigger a nonlinear response
is small, and quantum fluctuations are sizable corresponding to
fast switching timescales and correspondingly large Liouvillian gaps.
For small nonlinearities or large detunings, the number of particles
required to create a blueshift compensating the finite detuning is
large and quantum fluctuations play a negligible role in comparison.
The lifetime of the Liouvillian excited state is hence long since large
fluctuations, required to induce switching in between states, are
rare. Consequently, the Liouvillian gap decreases with increasing
detuning and increases for larger nonlinearities. In the thermody-
namic limit, i.e. for a diverging number of excitations, we expect
the gap to completely vanish

∆ → 0 for N → ∞. (4.15)

In this limit, quantum fluctuations vanish and the quantum descrip-


tion consequently converges to the mean-field solution where the
lifetimes of the two stable branches is infinite without any switching
in between them.

4.5 Dissipative phase transition in the thermo-


dynamic limit
We apply the photon correlation-based measurement technique de-
scribed in the previous section to the nonlinear polariton system.
The cavity detuning is chosen such that the polariton nonlinearity
is small due to a small exciton content. The polariton mode is
driven continuously with a detuning of ∆/γ = 1.5 as indicated by
the black arrow in Fig. 46 (a), i.e. in the regime where hysteresis
is predicted by a mean-field description. We then sweep the input
power and record individual photon events measured with an APD-
based HBT setup at each power allowing to extract both time traces
as well as the correlation functions from the data. Figure 46 (b)
displays the intensity probability distribution as a function of input
power obtained after temporally sampling the photon counts with
a bin width of 1 ms. Between 750 µW and 1 200 µW, we find that

69
4. Dissipative Phase Transition

1
2 (a) (b)

norm. Intensity
Counts (103 )

p(I)
0 0
−4 −2 0 2 4 0.8 1.0 1.2
∆/γ Power (mW)

Figure 46: Polariton intensity distribution. (a) Blue dots: Resonant scan of the
excitation energy across the polariton mode at 100 nW input power. Blue line:
Lorentzian fit to the data. The black arrow denotes the detuning chosen for the
correlation experiments. (b) Transmitted photon number distribution. Between
0.75 mW and 1.20 mW the distribution is double-peaked indicating the bimodality
of the steady state. Note the nonlinear color mapping used to highlight the small
contributions in between these states stemming from switching events faster than
the binning time. The colored arrows indicate powers where the data in Fig. 47 has
been taken.

the distribution is double-peaked underlining the bimodal charac-


ter. Moreover, the temporal bin width sets a lower bound on the
switching dynamics, i.e. observing bimodality here implies that the
lifetimes of the two states is on the order of 1 ms or longer. Yet,
there are also finite contributions of intensities in between the two
states which are due to switching events occurring within a single
time bin.

We now analyze the dynamics of the polariton system. In Fig. 47 (a)


we show normalized time traces for the excitation powers indicated
by arrows colored correspondingly in Fig. 46 (b) with a time bin
width of 100 ms. We clearly see that the system randomly switches
between two states with low and high intensity, respectively. More-
over, the occupation probability depends on the excitation power,
i.e. we mostly observe the low-intensity state for low powers with
only occasional jumps to the high-intensity state and vice versa for
high excitation powers. This difference in occupation reflects the
weights of the two states constituting the steady state density opera-
tor in Eq. (4.13). The strong variation in emitted photons when the

70
4.5. Dissipative phase transition in the thermodynamic limit

4
(a)

norm. Counts
2

0
0 20 40 60 80
Time (s)

(b)
3
g (2) (t)

1
10−8 10−6 10−4 10−2 100
Delay (s)

Figure 47: Time traces and photon correlations. (a) Photon count time traces
binned into 100 ms windows for the powers indicated by the arrows in Fig. 46 (b)
with corresponding color. (b) Correlation function for the same data. Below 10 ms
the single-photon correlation function is shown. For longer delays, the classical
autocorrelation function with 10 µs binning is used.

system jumps randomly between the two states is reflected by the


second-order correlation function displayed in Fig. 47 (b). Since the
calculation of single photon correlation events is computationally
demanding due to its quadratic scaling with both photon count rate
and delay time, we restrict the calculation as defined by Eq. (2.31)
to delays below 10 ms. For longer delays we exploit the convergence
of the correlation function to the classical autocorrelation given
by Eq. (2.33). The polariton emission exhibits super-Poissonian
photon statistics. The bunching amplitude depends on the exci-
tation power and is strongest around the center of the bistable
window where the strongest switching dynamics occurs. Note that
it may well exceed that of thermal light emission with g (2) (0) = 2.
Outside of the bistable window, when the system is exclusively
in either the low- or the high-intensity state, we retrieve coherent

71
4. Dissipative Phase Transition

photon statistics with g (2) (t) = 1. The bunching signal displays


a twofold decay as a function of time delay between two photon
events: Next to a strongly power-dependent decay with a timescale
that varies over several orders of magnitude upon small variations
of the excitation power, there is another decay on a hundreds of
milliseconds scale that seems to be unaffected by the excitation
power. Note that both timescales exceed the intrinsic polariton
lifetime of about 30 ps by many orders of magnitude in good agree-
ment with the diverging timescales expected for the critical slowing
down at the DPT predicted in Section 4.3. The appearance of two
decay times as well as the independence of the longer one on the
excitation power is yet unexpected from the discussion in Section 4.4
since only the timescale associated to the Liouvillian gap should
occur. Upon closer examination of the Liouvillian eigenspectrum
simulations in Fig. 45 closest to the experimental parameters (i.e.
detuning ∆/γ = 1.5 and smallest nonlinearity in the top right
panel), one notices that in addition to the vanishing gap spanned
by the steady state and lowest-lying excited state, also the second
excited state eigenvalue is slightly reduced with respect to a linear
system. Note that in this simulation, the nonlinearity U /γ = 0.05
is still significant in order to truncate the Hilbert space at excitation
manifolds where the required computation time is not excessively
long. We thus expect that for a system with small nonlinearities,
a sizable reduction also for higher-lying eigenstates occurs and its
density matrix upon a perturbation generally is a superposition of
multiple Liouvillian eigenstates with different decay rates back to
the steady state. While this picture explains the observation of the
twofold decay of the photon bunching signature, the independence
of the longest one on the excitation power requires an extension of
the model. Since not only quantum fluctuations cause switching
in between the two states we invoke the effect of classical noise
originating e.g. from cavity fluctuations. In order to incorporate
dephasing caused by this type of noise, we include an additional

72
4.5. Dissipative phase transition in the thermodynamic limit

κ/γ = 10−4 κ/γ = 10−3 κ/γ = 10−2 κ/γ = 10−1


0
−Re [λρ /γ] 10

10−2

10−4

10−6
0 2 4 6 0 2 4 6 0 2 4 6 0 2 4 6
F /γ F /γ F /γ F /γ

Figure 48: Liouvillian eigenspectrum with dephasing. Real parts of the lowest-
lying Liouvillian excited state eigenvalues including dephasing with rates κ/γ = 10−4 ,
10−3 , 10−2 , and 10−1 (orange solid lines, left to right). The blue solid lines depict
the case without dephasing (κ = 0). Detuning and nonlinearity are ∆/γ = 1.5 and
U /γ = 2.5 × 10−2 , respectively.

dissipative term into the Liouvillian introduced in Eq. (4.2)

ih i
L̂ = − Ĥ, ρ̂
~
γ
(4.16)

+ 2âρ̂↠− ↠âρ̂ − ρ̂↠â
2
κ † 
+ 2â âρ̂â↠− â↠↠âρ̂ − ρ̂â↠↠â .
2

The dissipator ↠â with rate κ conserves the mode population but
changes the energy of the resonance mimicking dephasing due to
spectral fluctuations of the cavity. Figure 48 illustrates how this
reduces the closing of the Liouvillian gap, effectively pinning it to
the dephasing rate in the regime where min {−Re [λρ (κ = 0)]} 
κ  γ. The cavity noise spectrum in Fig. 313 exhibits pronounced
resonances in the range . 100 Hz in good agreement with the
millisecond timescales associated with the power-independent decay
of bunching. We thus conclude that in the absence of dephasing, the
Liouvillian gap, and hence the longest-lived bunching decay, would
indeed be strongly power-dependent and well exceed the second
scale. However, in the presence of classical dephasing noise present
in the semi-integrated fiber system the dynamics are pinned to
the associated timescale independent of how strongly the system is

73
4. Dissipative Phase Transition

driven. The bunching decay observed in Fig. 47 (b) should hence be


interpreted as a lower bound for the DPT-induced critical slowing
down. Note, however, that already this exceeds the intrinsic lifetime
of the polaritons by more than nine orders of magnitude indicating
the magnitude of fluctuations required to induce these switching
events [171].

4.6 Polariton nonequilibrium dynamics in the


quantum regime
After observing the DPT associated with vanishing quantum fluc-
tuations, we wish to investigate the opposite regime, i.e. stronger
nonlinearities associated with smaller mode population and more
pronounced quantum fluctuations. To increase the nonlinearity,
the cavity detuning is chosen such that the lower polariton exciton
Hopfield coefficient is about 40 % corresponding to U /γ ∼ 0.03. We
then measure the power-dependent second-order correlation func-
tion for the detunings indicated by the colored crosses in Fig. 49 (a).
Again strong, super-thermal photon bunching is observed for powers
where the mode population exhibits a nonlinear dependence. In
contrast to the previous section, however, the timescale after which
coherent statistics is retrieved is only on the order of a few nanosec-
onds. These fast dynamics do not allow us to experimentally ramp
the excitation power in a non-adiabatic fashion and we√do not ob-
serve a hysteresis cycle even for laser detunings ∆/γ > 3/2. Due
to the pronounced quantum fluctuations in this regime, mean-field
theory fails to describe the experimental results.
We analyze the scaling behavior of the bunching amplitude more
carefully. In order to extract the amplitude and timescale each
data set is fitted with an exponentially decaying bunching signal.
However, the timescale of this decay is similar to the time resolution
of the APDs used for photon detection1 . Their instrument response
function (IRF) is independently measured with short laser pulses
yielding a FWHM of only 64 ps but with a long-lived diffusion tail
due to carriers generated in the layers below the field region [172].
1
id100-20, ID Quantique SA

74
4.6. Polariton nonequilibrium dynamics in the quantum regime

1.0
(a) (b)
Counts (105 ) 6

Counts (103 )
4
0.5

0.0 0
−4 −2 0 2 4 200 400
∆/γ Power (µW)

600 (c) g (2) (t)


0 1 2
Power (µW)

400

200

−6 −4 −2 0 2 4 6
t (ns)

Figure 49: Polariton correlation function. (a) Blue dots: Resonant scan of
the excitation energy across the polariton mode at 100 nW input power. Blue
line: Lorentzian fit to the data. The crosses denote the detunings chosen for the
correlation experiments. (b) Detected photons as a function of pump power at
∆/γ = 0.83 (red cross in (a)). The different lines correspond to ramping the power
up and down repeatedly. (c) Power-dependent second-order correlation function
for the same detuning. For powers where a nonlinearity in the mode population is
observed, superthermal photon statistics occur.

In order to extract the parameters, the influence of the IRF thus


has to be taken into account and the fit function reads
 
g (2) (t) = 1 + IRF(t) ∗ A · e−|t|/τ , (4.17)

where ∗ denotes the convolution


Z ∞
(f ∗ g) (t) = f (τ )g(t − τ ) dτ. (4.18)
−∞

The extracted amplitude and timescale for the detunings indicated


by crosses in Fig. 49 (a) are displayed in Fig. 410 (a) and (b), respec-

75
4. Dissipative Phase Transition

3.0 (a) ∆/γ = 1.0 (b)


0.8 3
2.5
0.6
g (2) (0)

τ (ns)
2
2.0 0.4

1.5 0.2 1

1.0 0
0 200 400 0 200 400
Power (µW) Power (µW)

Figure 410: Photon bunching and timescale. Parameters extracted from fitting
Eq. (4.17) to the power-dependent photon correlation data for detunings ∆/γ
indicated by colored text (correspondingly colored crosses in Fig. 49 (a)). In (b)
the extracted timescales are only displayed for powers where g (2) (0) > 1.05 due to
the large uncertainty on the extracted τ for weak bunching signals.

tively, with corresponding line colors. Both of them increase with


increasing laser detuning and the observed timescales range from a
few hundred picoseconds up to a few nanoseconds. Qualitatively,
the behavior is identical for all detunings and there is in particular
no distinction
√ above and below the mean-field bistability threshold
∆/γ = 3/2, demonstrating a regime where the mean-field descrip-
tion is invalid. This also implies that dynamical hysteresis may
occur for all positive detunings provided the excitation power is
ramped non-adiabatically [173].

4.6.1 Liouvillian gap scaling


Although the phase transition observed in the bistable polari-
ton system is of first order without divergence at the critical
point and associated universality, we follow Ref. [165] and ana-
lyze the scaling behavior of the bunching decay around the critical
drive strength. We define this as longest timescale observed, i.e.
min {−Re [λρ (F )]} = −Re [λρ (Fc )]. Figure 411 displays the scaling
of the extracted decay times τ around this√ point where the power
and excitation amplitude are related by P = F . The finite values
at the critical point are a clear indicator for the absence of a con-
tinuous phase transition. Moreover, there is excellent qualitative

76
4.6. Polariton nonequilibrium dynamics in the quantum regime

101 101
(a) ∆/γ = 1.0 (b)

−γ/Re [λρ ]
100
τ (ns)

∆/γ = 0.2
−1
10 100
−100 −10−1 0 10−1 100 −100 −10−1 0 10−1 100
F /Fc − 1
p
P /Pc − 1

Figure 411: Bunching decay around the critical point. (a) Detuning-dependent
bunching timescales and (b) simulated inverse of the three lowest-lying Liouvillian
excited state eigenvalues as a function of excitation amplitude. The detunings in
(b) are identical to (a) with corresponding color and U /γ = 0.2.

agreement between the observed bunching times and the inverse


of the numerically obtained Liouvillian eigenvalues, i.e. associated
timescales. This demonstrates the predicted suitability of pho-
ton correlations for the investigation of Liouvillian eigenspectra
of driven-dissipative systems. Quantitatively, there is yet a dis-
crepancy: At the critical point, the dynamics of the experimental
data for, e.g., the largest detuning are slowed down by roughly
two orders of magnitude in units of the intrinsic decay time. For
the simulations, however, this factor is slightly below one order of
magnitude. This is due to the rather large nonlinearity chosen in
the Liouvillian eigenspectrum simulations to ensure a manageable
computation time and we expect also quantitative agreement with
the data when this is reduced appropriately. In particular, this anal-
ysis provides a handle to determine the single-particle nonlinearity
U , in contrast to the mean-field value h↠âi × U that is accessible
with other techniques.

In Ref. [165], the scaling of the Liouvillian gap as the critical drive
is approached has been introduced as the operational criterion to
access the number of excitations, i.e. determining how far the
system is from the thermodynamic limit. It has been shown that
the slowing down of dynamics becomes sharper towards this limit
and converges to a singularity. In Fig. 412 (a) we display power-law

77
4. Dissipative Phase Transition

(a) pol
exp
data
100 100
τ (ns)

τ (ns)
(b)
10−1 10−1
−0.6 −0.4 −0.2 0.0 −0.6 −0.4 −0.2 0.0
p p
P /Pc − 1 P /Pc − 1

Figure 412: Bunching scaling. (a) Bunching decay time below the critical point
(points) and power-law fits (dashed lines). (b) Bunching decay time for ∆/γ = 0.6
together with power-law and exponential fit (dashed and dash-dotted line, respec-
tively).

fits to the data presented in Fig. 411 (a) as the critical point is
approached. However, we note that also an exponential fit function
provides a reasonable agreement and the power-law fit only serves
as a means to assess the steepness of the data. In particular, the
choice of a fit function is not motivated by theoretical predictions
as for continuous phase transitions.

Figure 413 shows the analysis of the detuning-dependent correlation


function. The critical power for which the longest-lived bunching
is observed scales linearly with the excitation detuning. Since
the slowest dynamics are observed at drive strengths where the
nonlinear response of the system is triggered, the linear scaling
reflects the compensation of a larger energy detuning by an increased
number of excitations. The exponent extracted from the power-
law fits expresses how steep the scaling of the extracted bunching
timescale is as the critical point is approached. We find a quadratic
dependence on the excitation detuning. In Ref. [165] the degree to
which the system is in the thermodynamic limit has been set by
reducing the nonlinearity Ũ = U /N √ while compensating for this
by an increased drive strength F̃ = N F . Here, N is a scaling
parameter with N = 1 corresponding to the quantum regime and
N → ∞ is the thermodynamic limit. In the experiments described
here, the detuning plays a similar role as this scaling parameter.

78
4.7. First- vs. second-order coherence

(a) (b)
400
2
Power (µW)

Exponent
300
1
200

100 0
0.25 0.50 0.75 1.00 0.0 0.5 1.0
∆/γ ∆/γ

Figure 413: Critical point and exponent scaling. (a) Blue dots: Detuning-
dependence of the critical power Pc for which the longest bunching decay time is
observed. Orange dashed line: Linear fit to the data. (b) Blue dots: Extracted
exponent of the power-law fits in Fig. 412. Orange dashed line: Quadratic fit to
the data.

Whereas the system nonlinearity remains unchanged, increasing


the detuning requires an increased drive strength to trigger the
reduction of the Liouvillian gap. In this sense, the system is pushed
closer to the thermodynamic limit as the detuning is increased. The
increasing power-law exponent for larger ∆ reflects this behavior
demonstrating how photon correlation spectroscopy assesses the
’quantumness’ of a system.

4.7 First- vs. second-order coherence


The analysis in this chapter is entirely based on the second-order
coherence function, i.e. photon correlations. One may ask whether
the same information can also be retrieved by analyzing the first-
order coherence since nonlinearity and quantum fluctuations may
also reduce the photon coherence. In Fig. 414 we compare time-
and power-dependent g (2) and g (1) functions demonstrating that
this is indeed the case: For excitation strengths where photon
correlation measurements exhibit super-Poissonian statistics, also
the first-order coherence is reduced. Moreover, the linecuts for
different detunings show that the timescale over which this reduction
occurs are identical to that of the bunching decay. In this sense,
the information about the Liouvillian gap that can be extracted

79
4. Dissipative Phase Transition

3 1.3
(a) g (2) (t) (c)
0.6 1.0 1.4
2 1.2

g (2) (t)
F /γ

1 1.1

0 1.0
∆/γ = 0.2
(b) g (1) (t)
0.7 1.0 0.4
2 0.9
0.6

g (1) (t)
F /γ

1 0.8 0.8
(d) 1.0
0 0.7 0
0 10 20 30 10 1
10
t×γ t×γ

Figure 414: Comparison of first- and second-order coherence. Time- and


power-dependent first- and second-order coherence ((a) and (b), respectively). The
detuning is ∆/γ = 1. (c) and (d) show the coherence functions at the excitation
power for which the longest-lived bunching signature is observed. (dashed lines in
(a) and (b)). The detunings are ∆/γ = 0.2, 0.4, 0.6, 0.8, and 1.0 indicated by
colored text. The nonlinearity is U /γ = 0.2.

from the two different coherence functions are identical. From an


experimental point of view, measuring first-order coherence may
yet be much more challenging. In order to access the second-scale
dynamics reported in Section 4.5, optical path length differences on
the order of 108 m are required. While this is in principle achievable,
the sensitivity of first-order coherence to phase noise imposes highly
challenging requirements on the excitation laser bandwidth and
overall phase stability. Photon correlations, on the other hand,
are insensitive to this making them experimentally much easier to
access.

4.8 Conclusion and perspectives


In this chapter, photon correlation measurements have been used
to demonstrate the dissipative phase transition occurring in Kerr-
type nonlinear exciton polariton systems. Although the presented

80
4.8. Conclusion and perspectives

measurements are of a proof of principle nature, the ability to access


both regimes, close to thermodynamic limit and with pronounced
quantum fluctuations, demonstrates the versatility of the technique
to study driven-dissipative systems. In particular, the detuning-
dependent scaling behavior could be linked to the role of quantum
fluctuations over the mean excitation number. The observed closing
of the Liouvillian gap and associated non-adiabatic response at the
critical point can be interpreted in analogy to the Kibble-Zurek
mechanism in quenched quantum phase transitions [170, 174–176].
Our technique complements that of dynamical optical hysteresis
which is based on ramping the excitation power in a non-adiabatic
way and mapping out the resulting hysteresis area. Similar to
photon correlations, this measurement allows to extract information
on the Liouvillian spectrum [67, 177].
In general, the dependence on microscopic parameters and lack of
universality makes the class of first-order phase transitions such
as the one considered here less interesting compared to continuous
ones. Yet, there are also puzzling observations for this type of
transition [178] and the limited attention in the past warrants
further experiments on this less-understood class from a theoretical
point of view.
We expect that the presented technique can be readily applied in
other experimental settings. In particular, we wish to highlight the
second-order phase transition occurring in planar exciton-polariton
microcavities under parametric scattering [15, 179, 180]. Exciting
the lower-polariton branch at the inflection point of its dispersion
while measuring correlations of polariton emission from the |k| = 0
momentum state should allow for observing a critical slowing down
of dynamics similar to the one observed here but with true, algebraic
divergence as a function of drive strength. This optical parametric
oscillator scenario has been predicted to feature a rich phase diagram
upon tuning the cavity-exciton detuning as well as the excitation
strength. In the low-power regime, it can be mapped onto the so-far
experimentally elusive 2D compact anisotropic Kardar-Parisi-Zhang
phase, whereas the phase in the high-power regime is characterized
by a dynamical XY universality class. Experimentally, those are
accessible via temporal and spatial first- and second-order coherence

81
4. Dissipative Phase Transition

functions in analogy to the analysis presented here [181]. Moreover,


complex types of order particular to the nonequilibrium nature are
expected [182, 183] promoting exciton-polariton systems as an ideal
testbed for experiments on nonequilibrium phenomena.
The extension of the photon correlation technique to other phys-
ical systems, e.g. circuit QED or Rydberg atoms, where similar
signatures of the Liouvillian eigenspectrum have been observed
[166, 167, 171, 184], is straightforward. Moreover, the phases of
driven-dissipative many-body systems have recently garnered a lot
of interest [185–191] and we expect that photon correlations provide
a useful tool for experiments in this emerging field.

82
Chapter 5

Quantum Correlations of
Exciton-Polaritons

In this chapter we discuss measurements towards the obser-


vation of quantum-correlated polarity. Section 5.1 describes
the upconversion-based photon correlation technique used to
achieve the required time resolution1 . In Section 5.2 we
apply this technique to investigate the statistics of polariton
emission indicating the onset of non-classicality.

Since the first observation of strong coupling between quantum


well excitons and cavity photons by Weisbuch et al. [12], the field
of exciton-polaritons opened up a plethora of experiments. The
favorable properties of those particles, such as strong nonlinear
interactions, low effective masses, and all-optical access, manipu-
lation, and measurement, make them a suitable platform to study
fundamental properties. Among those, Bose-Einstein condensation
at elevated and even room temperatures [14, 16, 17, 19, 193], su-
perfluidity [20, 21], dark solitons [194–196], and the optical spin
hall effect [197, 198] are the most notable ones. The ability to
engineer the potential landscape and create polaritonic lattices
1
The assembly of the detection setup as well as characterization measure-
ments presented in Section 5.1.2 have been performed by J. Chun Tat Ngai and
A. Delteil under the supervision of A. Delteil [192].

83
5. Quantum Correlations of Exciton-Polaritons

by fabrication of an initially two-dimensional system [24] enabled


experiments where other physical systems are simulated by an ar-
ray of coupled exciton-polariton sites [199]. However, to this date,
most of these experiments can yet be described by a classical mean-
field description introduced in Section 4.2.1, and demonstrations
of the single-particle quantum character remains elusive. In order
to use exciton-polariton lattices for simulations of more complex
systems intractable with classical computers, interactions at the
single quantum level are crucial. For example, the occupation of
a single site in a one-dimensional Bose-Hubbard model of coupled
exciton-polariton microcavities should inhibit the occupation of the
neighboring site. Signatures indicating the quantum character of
exciton-polaritons are thus highly sought after.

Cuevas et al. [200] went a step into the polariton quantum realm. In
their experiment, they demonstrated that a pair of entangled pho-
tons, where one of them is converted into a polariton excitation and
eventually back into a photon upon cavity decay, maintained their
entanglement. Whereas entanglement is indeed a truly quantum-
mechanical property, this experiment does not address the question
whether the interaction of polaritons, crucial for a quantum simu-
lator, also exhibit quantum behavior. Rather, the quantumness is
imposed by the source with which the polariton is excited instead
of being intrinsic to the polaritons themselves. In fact, the findings
observed in this experiment are also expected from a bare cavity,
which can be described as a classical harmonic oscillator without
any interactions. The results hence only show that polaritons are
not detrimental to the quantumness of the source by e.g. reducing
the quantum coherence via pure dephasing.

Turning this concept around, the signature unambiguously demon-


strating the quantum character of polaritons is the observation of
polariton blockade introduced in Section 2.7. In this case, the exci-
tation is a classical, coherent drive, yet the emitted light exhibits
non-classical statistics stemming from polariton quantum dynamics.
Observation of photon antibunching heralds strong interactions
between polariton excitations, and indicate the suitability of the
system as a building block for a polariton-based quantum simulator.

84
5.1. Upconversion-based photon correlation measurements

5.1 Upconversion-based photon correlation


measurements

detection efficiency (%)


Device time resolution (ps) duty cycle
at 850 nm

fast APD2 5 55 –
slow APD3 45 300 –
SSPD4 . 85 25 –
streak camera 0.26 4 6 × 10−6
upconversion ∼ 65 5 4 × 10−4

Table 51: Photon detection parameters. Specifications of different technologies


to measure photon correlations. The indicated time resolution is the FWHM of the
IRF. Note that APDs have a long-lived diffusion tail which further reduces their
resolution in contrast to the Gaussian IRF of SSPDs. Although APDs and SSPDs
do not have a duty cycle during which they are sensitive, the maximum count rate
they can measure is typically on the order of 10 MHz due to their dead time after a
photon detection event.

Experimentally, the short lifetime of polaritons, on the order of


10 ps, poses a challenge to measuring photon correlations with the
time resolution required to resolve the dynamics over which Pois-
sonian statistics are recovered. Table 51 summarizes parameters
of different techniques commonly used to this end as well as the
upconversion technique employed here. Among the single-photon
detection devices, superconducting single-photon detectors (SSPDs)
offer both the highest quantum efficiency and best time resolu-
tion. Yet, the latter still exceeds the lifetime of most polariton
systems making it unsuitable to resolve their dynamics. In earlier
experiments during this dissertation, a streak camera-based photon
correlation setup has hence been developed. This technique relies
on the conversion of incoming photons to electrons which are then

2
ID100, ID Quantique SA
3
SPCM-AQRH, PerkinElmer Inc.
4
SSPD, Scontel
5
The detection efficiency is composed of the coupling efficiency into the
nonlinear crystal, the upconversion efficiency, and the quantum efficiency of the
APDs (about 40 %, 80 %, and 20 %, respectively.)

85
5. Quantum Correlations of Exciton-Polaritons

deflected by a fast voltage ramp. Recording the impact position of


the electrons on a fluorescent screen then allows to extract informa-
tion about the arrival time and to calculate the photon correlation
function [201–203]. The main limitation of this technique is its small
duty cycle: In our case, the screen can host 60 770 ps-long time
traces, and readout of the positions takes 8 ms, mainly limited by
the millisecond-scale decay of the screen fluorescence. This leads to
a small duty cycle of only 5.8 × 10−6 of the acquisition time during
which the streak camera records the incoming photons. Together
with the low quantum efficiency of 0.26 %, the acquisition times
required for this technique to obtain a sufficiently small uncertainty
on the photon correlations turned out to be too long: Assuming
that every photon emitted from a polariton system with 10 ps life-
time is sent to the device, the coincidence rate is only 0.5 s-1 with a
5 ps-wide binning window. Even in this unrealistic scenario of unity
collection efficiency, the required acquisition time in order to obtain
an uncertainty of 0.02 on the g (2) value already exceeds 3.5 days.
While this does not a priori invalidate streak camera-based corre-
lation measurements, we therefore employ a different technique to
measure photon correlations with picosecond time resolution based
on ultrafast gating of our signal. It is based on the SFG of the
signal to be measured and a strong pump pulse. The principle
of SFG is schematically shown in Fig. 51 (a). Two photons with
wavelengths λS and λP , denoting signal and pump, respectively,
are superimposed in a medium with strong χ(2) nonlinearity which
allows them to be upconverted to a single photon with shorter
wavelength λSFG . Energy conservation requires that

1 1 1
+ = . (5.1)
λS λP λSFG

The finite phase mismatch between the pump, signal, and upcon-
verted light would normally lead to an oscillatory SFG amplitude as
they propagate through the nonlinear crystal, indicated by the blue
line in Fig. 51 (b). In order to overcome this, we use a so-called
quasi phase matching (QPM) technique that relies on a periodic
inversion of the nonlinearity in the χ(2) medium [204, 205]. In this
way, a large build-up of SFG amplitude similar to perfect phase

86
5.1. Upconversion-based photon correlation measurements

(a) (b)

SFG intensity (a.u.)


λS

λSFG
∆k 6= 0
∆k = 0
λP
QPM

0 1 2
χ(2) medium z/(2π/∆k)

Figure 51: Sum frequency generation. (a) Schematic of sum frequency genera-
tion. In a medium with strong χ(2) nonlinearity a pump and a signal photon with
wavelengths λp and λS , respectively, are converted into a sum frequency photon
with wavelength λSFG . (b) SFG intensity with perfect phase matching and finite
phase mismatch (orange and blue line, respectively) as a function of crystal length.
Due to the periodic inversion of the nonlinear coefficient in the medium, indicated
by grey shading and arrows, QPM allows for a build-up of SFG intensity as the
signal passes the crystal even in the presence of a finite phase mismatch.

matching can be reached as the light passes the crystal (compare the
orange and green line in Fig. 51 (b) denoting true phase matching
and QPM, respectively).
The principle to achieve a high time resolution in photon correlation
measurements relies on using a pulsed pump laser. In this way,
the SFG component is only generated at times where the pump
pulse and the incoming signal overlap temporally, i.e. the signal
is sampled with a sampling window given by the length of the
pump pulse [206]. The technique introduced here has also been
used to transduce photons from the near- and mid-infrared to
the visible range enabling usage of highly efficient photodetection
devices available for this range while maintaining their quantum
properties [207–209].

5.1.1 Setup
The SFG-based photon correlation setup is schematically shown in
Fig. 52. The signal and pump pulses are superimposed in a nonlinear
crystal with strong χ(2) nonlinearity. We use a periodically poled

87
5. Quantum Correlations of Exciton-Polaritons

Signal Upconversion Correlation


DM CF
Laser
ppLN WG
EOM
t2 − t1
dump

Pump pulse
TCSPC

4K
FPC

MIRA

Figure 52: Schematic of upconversion-based g (2) setup. The four main parts
are separated by colored boxes. Green box: The signal to be measured (green
pulses) is generated by exciting polaritons with EOM-modulated pulses of a few
hundred picosecond pulse length. The transmitted signal is collected and sent to
the measurement setup. Red box: The pump pulses at 950 nm wavelength for sum
frequency generation (red pulses) are generated by a mode-locked laser (MIRA).
The emitted light is split into two paths with a variable relative delay in between
before recombination creating two consecutive pump pulses with a controllable
delay t2 − t1 . The clock cycle of the pulsed laser is also used to trigger the EOM
modulating the polariton excitation laser. Grey box: Both signal and pump pulses
are combined with a dichroic mirror (DM) and coupled into the nonlinear crystal
waveguide (ppLN WG) creating sum-frequency emission at 446 nm (blue pulses)
at times where the two overlap. The unconverted signal light as well as pump
and second harmonic emission are dumped with a color filter (CF) transmitting
only the sum frequency components. Blue box: The two sum-frequency pulses
enter a conventional HBT setup with two APDs. Photon events are recorded
with time-correlated single photon counting (TCSPC) electronics and coincidences
processed in real-time.

lithium niobate waveguide. The sum-frequency emission then enters


a conventional APD-based HBT setup. Note that the used APDs
may well have a low time resolution since the time resolution of the
setup is set by the length of the pump pulses probing the signal light
rather than time-resolved detection. The delay between consecutive
pump pulses then sets the delay for which correlations are probed.
In this configuration, recorded photon coincidences may not only
stem from a single upconverted photon created in each pump pulse.

88
5.1. Upconversion-based photon correlation measurements

Instead, we also record photon pairs due to two photons generated


in each of the two pulses individually. For a given pump pulse
separation ∆t the measured coincidences C are thus
C(∆t) = G(2) (0) + G(2) (∆t). (5.2)
In the further analysis, we implicitly take this renormalization into
account unless stated otherwise. On top of correlations created
by the two pump pulses with delay ∆t, we also record photon
coincidences from events corresponding to pulses from different
periods, i.e. different cycles of the pulsed laser. The repetition
period of Trep = 13.1 ns is orders of magnitude longer than the
lifetime of the polaritons and these coincidences hence exhibit
uncorrelated statistics. We exploit this to normalize the same-
period correlations with the uncorrelated coincidences from the
next 10 periods. The upconversion efficiency is highly polarization-
dependent and chromatic, restricting the signal wavelengths for
which large SFG amplitudes are obtained to a spectral window
of about 1.7 meV FWHM6 . Photon correlations measured with
this setup will hence automatically single out the lower polariton
mode spectrally (note the normal mode splitting of about 3.5 meV).
Moreover, photons with misaligned polarization are not upconverted
and hence not detected.

5.1.2 Characterization
In order to verify the ability to measure photon correlations with
picosecond time resolution with the upconversion-based setup we
conduct test measurements and characterize its performance. We
first measure the signal of a bare CW laser. As expected, we obtain
a value of g (2) (∆t) = 1.00 ± 0.01 for all delays accessible with our
system. While this demonstrates that no spurious non-Poissonian
statistics are induced by the correlation setup, we wish to also test
it with time-varying signals to characterize its dynamical response.
Figure 53 (a) displays the second-order correlation function using
a CW laser modulated with an electro-optic modulator (EOM)
6
The conversion bandwidth is determined by a convolution of the pump
pulse spectrum (bandwidth 263 µeV in our case) and the conversion efficiency
which saturates around the center of the pulse spectrum

89
5. Quantum Correlations of Exciton-Polaritons

driven sinusoidally at 2 GHz as the signal light. Note that the EOM
is operated at the minimum of the sinusoidal response function
leading to an effective signal modulation at 4 GHz. The measured
photon correlations clearly reproduce the modulation as a function
of photon delay with a visibility of (38.2 ± 0.2) % and a period
matching that of the signal. This visibility is close to the maximum
value of 50 % achievable for a sinusoidally modulated signal and
the deviation is due to the limited bandwidth of the EOM reducing
the modulation depth. In Fig. 53 (b) we display photon correlation
events from different periods indicated in the legend. The index
q denotes that the second photon of a pair was recorded q laser
periods later, i.e. with a delay of q × Trep + ∆t. Both the phase
and amplitude of the oscillations of different-period correlations
vary as a function of period index q due to the incommensurate
frequencies of the signal modulation and pump pulse repetition
rate. Comparing the observed oscillations of photon correlations to
measurements taken with SSPDs on the same signal in Fig. 53 (c)
reveals a visibility reduced to (26.4 ± 0.8) % for the latter indicating
that the SFG-based correlation technique has a significantly higher
resolution than SSPDs.
To quantify this further, a pulsed laser source similar to the pump
laser is used as signal light. Both signal and pump pulses run inde-
pendently without any synchronization between them. Figure 54 (a)
displays the resulting photon correlation function. Around zero
time delay, strong photon bunching reflects the pulsed excitation
used here. The FWHM width of (7.1 ± 0.2) ps obtained from a
Gaussian fit allows to estimate the time resolution of the setup with
s
7.1 ps 2

∆tIRF = √ − (2.5 ps)2 = (4.3 ± 0.2) ps. (5.3)
2

The division by 2 takes into account that in a g (2) measurement
the pulse is convolved with itself, and the second term under the
square root denotes the pulse length of 2.5 ps for both signal and
pump pulses. Note that Eq. (5.3) assumes a Gaussian IRF as well
as Gaussian pulses and hence only serves as an estimate for the
time resolution. In Fig. 54 (b) coincidences of photon pairs from
different pump laser periods are shown. The bunching peak observed

90
5.1. Upconversion-based photon correlation measurements

30
g (2) (0) + g (2) (∆t)

28
26 249 ps

24
22 (a)

−200 −100 0 100 200 300


∆t (ns)

q 200
g (2) (0) + g (2) (∆t)

25 1
2
G(2) (∆t)

3 150
20
4

15 100
(b) (c)

−200 0 200 0 500 1000 1500


∆t (ns) ∆t (ps)

Figure 53: Photon correlations of an EOM-modulated CW laser. The EOM is


driven at 2 GHz close to a minimum of its response function, effectively modulating
the CW laser at 4 GHz. (a) Blue solid line: Photon correlations as a function of delay.
The g (2) function reproduces the modulation of the input signal with a visibility
of (38.2 ± 0.2) %. The blue shaded area is the RMS error and the solid orange
line is a sinusoidal fit to the data. (b) Correlations of photons from consecutive
laser cycle periods indicated in the legend. The incommensurate frequencies of
the signal modulation and pump laser lead to a varying phase and amplitude as a
function of period index q. (c) Photon correlations measured with SSPDs on the
same signal. Again, they reproduce the modulation of the input signal but with a
reduced visibility of (26.4 ± 0.8) % compared (a).

91
5. Quantum Correlations of Exciton-Polaritons

800 20 ps
(a) (b) q
g (2) (0) + g (2) (∆t)

g (2) (0) + g (2) (∆t)


2500 1
600
2
7.1 ps 400 3
2000 4
200
1500
0
−20 0 20 −20 0 20
∆t (ps) ∆t (ps)

Figure 54: Resolution of upconversion setup. (a) Blue dots: Photon correlation
function using a pulsed laser with 2.5 ps pulse length (FWHM) as signal light. The
blue shaded area denotes the uncertainty (one standard deviation). The orange
solid line is a Gaussian fit to the data with an extracted width of (7.1 ± 0.2) ps. (b)
Different-period photon correlations. Due to the small mismatch of the pump and
signal laser repetition rate, the bunched signal in (a) splits up into two peaks with
increasing splitting for increasing period difference q.

in Fig. 54 (a) splits into two peaks with half the coincidences
each symmetrically located around zero time delay. The splitting
increases for increasing pump laser pulse period difference q and
occurs due to the small difference in repetition period of the two
lasers used for signal and pump on the order of 10 ps.

5.2 Polariton correlation measurements


Having verified the correct operation of the upconversion-based
setup with picosecond time resolution, we apply it to measure
photon correlations. We use the emission from resonantly excited
polaritons as signal light in order to observe sub-Poissonian statistics
heralding single-particle polariton interactions. Following Eq. (2.34),
we choose a red-detuned CW laser excitation with ∆/γ = −0.5
where the largest degree of antibunching is expected. Yet, even in
this optimal case we exclusively observe bunched photon emission
with a bunching amplitude of g (2) (0) = 1.05 to 1.10 which decays on
a timescale of 100 ps to 200 ps for excitation powers between 5 µW
and 50 µW. This observation agrees with earlier APD-based exper-
iments. There, the time resolution was not sufficient to observe the

92
5.2. Polariton correlation measurements

dynamics of the expected antibunching signature but the longer-


lived bunching decay could be resolved. Hence, both conditions for
classical light in Eq. (2.32) are not violated. Whereas for vanishing
ratios U /γ we would expect any antibunching signature to con-
verge towards the statistics of coherent emission with g (2) (0) = 1,
the appearance of photon bunching is surprising for negative laser
detunings. The observed timescale is also much longer than the
measured polariton lifetime and cannot be explained by the simple
dynamics predicted by Eq. (2.30). A possible reason for the occur-
rence of bunching is scattering of polaritons into a bath of localized
excitons. Those states emerge due to excitonic disorder7 and po-
laritons will form out of coupling of a certain superposition of these
to the cavity mode, yet leaving a large number of localized exciton
states weakly coupled to the polaritons. Due to their localization,
these excitons are typically longer-lived than polaritons and their
incoherent excitation might leave them in a thermal state resulting
in long-lived photon bunching upon eventual decay via the cavity.
Streak camera-based lifetime measurements presented in Fig. 55
seem to confirm this picture: For a sizable exciton content, the po-
lariton lifetime is reduced from (38.0 ± 0.1) ps to (17.2 ± 0.1) ps as
compared to the bare cavity. Moreover, a second, slower decay with
a decay time of (88 ± 6) ps appears which might be attributed to de-
cay of the longer-lived localized excitonic states. Muñoz-Matutano
et al. [210] reported a similar reduction of the polariton lifetime
accompanied by biexponential decay. They identified a negatively
charged trion resonance stemming from residual charges of carbon
impurities as a possible explanation. This resonance introduces
strong energy-dependent scattering around the lower polariton res-
onance into localized excitonic states since the trion binding energy
of 1.0 meV to 1.5 meV is on the order of the Rabi splitting [211].
Another reason for the shortening of the polariton lifetime could
be biexciton decay [71, 72] as well as disorder-induced momentum
scattering of localized polaritons into guided mode polaritons with
momentum ωc < |k| . nω c [44].

Even though the emergence of long-lived photon bunching, possibly


7
For systems with translational symmetry, all exciton states with in-plane
momenta |k| > ωc are dark.

93
5. Quantum Correlations of Exciton-Polaritons

101
|cx |2
10 0 0.0 38 ps

norm. Counts
0.5
10−1 17 ps

10−2 88 ps

10−3
600 700 800 900 1000 1100
t (ps)

Figure 55: Biexponential polariton decay. Ring-down measurements of the


polariton population with negligible (blue solid line) and roughly 50 % (orange solid
line) exciton content. The data is obtained by exciting the mode with a short laser
pulse and measuring the signal decay with the streak camera. For 50 % exciton
content, the lifetime is shortened and a biexponential decay sets in. The dashed lines
denote an exponential (red) and biexponential (green) fit to the data. Extracted
decay times are indicated next to the respective curve. The modulation of the
photon-like signal decay is due to beating of the two linearly polarized modes (cf.
Fig. 311 (b)). The data and fit for |cx |2 = 0 are offset by an order of magnitude
for clarity.

caused by localized excitonic states, is an unwanted process, it is not


a priori detrimental to the observation of single-particle polariton
interactions. In particular, the interplay of the two mechanisms
might lead to shorter-lived antibunched photon statistics on top
of a long-lived bunching background. In order to investigate the
dynamical properties of the photon statistics in more detail and
possibly avoid the build-up of a localized exciton population, we
switch to a quasi-CW excitation. The emission of the CW laser used
to excite the polaritons is modulated with an EOM synchronized
to the repetition rate of the pump laser (green box in Fig. 52). In
this way, we excite polaritons resonantly with an excitation pulse
length down to 250 ps. The pulse length chosen still exceeds the
polariton lifetime by about an order of magnitude. Conversely, the
bandwidth of the excitation pulse is negligible compared to the
polariton linewidth and the detuning of the excitation is thus still
well-defined.

94
5.2. Polariton correlation measurements

5.2.1 Same-time correlations


Figure 56 (a) shows the same-time correlation function g (2) (t, t)
along the excitation pulse (grey dashed line) again for a detuning
of ∆/γ = −0.5. Note that for same-time correlation measurements
the static arm of the two pump pulses in Fig. 52 is blocked. We
observe a build-up of photon bunching on a time scale of 100 ps
to 150 ps. We also measure the time- and detuning-dependent
emitted photon number in Fig. 56 (b,c) and (d,e) for low- and high-
power excitation, respectively. In these measurements, a dynamical
blueshift of the polariton mode for the higher-power excitation
is revealed. We identify two possible reasons, or a combination
thereof, to qualitatively explain the build-up of photon bunching
as a function of time.
• In this dynamical setting, the reservoir of long-lived localized
excitons feeds back into the cavity on a timescale given by
their lifetime. This implies that, provided they are in a
thermal state, the associated photon bunching only sets in
after this time and the photon statistics are governed by the
bare polaritons prior to that.
• The dynamical blueshift observed in Fig. 56 (b-e) implies that
a detuning of e.g. ∆/γ = −0.5, which at the onset of the pulse
would be optimal to enable the observation of antibunching,
is effectively increased at later times. Note that the shift is
on the order of the linewidth. Assuming a constant, bunched
background in agreement with earlier CW measurements,
this implies that at early times this background is maximally
compensated for by polariton antibunching. As time passes,
however, the effective detuning leads to a reduced degree
of antibunching such that the resulting statistics are overall
super-Poissonian converging towards the background offset.
Clearly, more research is needed to pin down the effect of polariton
scattering into the localized exciton reservoir. In particular, the
possible interplay between polariton antibunching and occurrence of
reservoir bunching, and their respective timescales warrants further
investigations, and we only attempt a qualitative description of the
observed phenomena here.

95
5. Quantum Correlations of Exciton-Polaritons

1
(a)

norm. Counts
1.5
g (2) (t, t)

1.0
0
−200 −150 −100 −50 0 50 100 150 200
t (ps)
1.0
400 (b) (c)
0.8
200

Counts (103 )
0.6
t (ps)

0.4
−200

100 nW 0.2
−400
400 (d) (e)
6
200

Counts (103 )
4
t (ps)

−200 2

900 nW
−400 0
−25 0 25 −250 0 250
∆E (µeV) t (ps)

Figure 56: Same-time correlations. (a) Blue dots (left axis): g (2) (t, t) as a
function of time relative to the excitation pulse. The blue shaded area denotes the
uncertainty (one standard deviation). Grey dashed line (right axis): Normalized
photon counts corresponding to the population of the polariton mode. The average
power is 700 nW corresponding to a pulse peak power of 35 µW. (b,d) Detuning-
and time-dependent photon counts. The respective average powers are indicated
in the bottom right. In the high-power case, the mode blueshifts as a function of
time. (c,e) Linecuts along the dashed and correspondingly colored lines in (b,d)
illustrating the different dynamical responses for different laser detunings.

96
5.2. Polariton correlation measurements

1.2

g (2) (t, t)
1.1

1.0

−1.5 −1.0 −0.5 0.0 0.5 1.0 1.5


∆/γ

Figure 57: g (2) (t, t) detuning dependence. Blue dots: Same-time correlations
as a function of normalized laser detuning. The blue shaded area denotes the
uncertainty (one standard deviation). Orange line: Fit to the data based on a model
including a constant bunching background (see main text). Extracted fit values
are U /γ = 0.08 ± 0.02 and offset 0.11 ± 0.01 (green line). The cavity-exciton
detuning is chosen such that |cx |2 = 0.5.

We now wish to study the detuning dependence of the same-time


photon correlations in this dynamical setting. Figure 57 displays
g (2) (t, t) as a function of normalized laser detuning ∆/γ. As ex-
pected, the correlation function increases as the laser energy crosses
the polariton mode from a red- to a blue-detuned scenario. However,
we do not observe any sub-Poissonian behavior even for ∆/γ = −0.5
but instead the data seems to vary around a constant bunching
background. We fit the data with a model that takes into account
the expected detuning behavior according to Eq. (2.34) as well as a
constant bunching offset (orange curve in Fig. 57). The extracted
fit values are U /γ = 0.08 ± 0.02 and 0.11 ± 0.01 for the bunching
background. For these parameters, we achieve a good agreement
between the data and the model further underlining the inter-
play between the expected polariton statistics behavior and overall
bunched and detuning-independent reservoir contribution. Taking
into account the polariton lifetime measurement in Fig. 55, the
extracted nonlinearity agrees well with an estimate of U = 2.4 µeV
based on an exciton-exciton interaction strength of 30 µeVµm2 [66,
67], a confinement area of π × 1 µm2 (see Section 3.3.3) and an
exciton content of |cx |2 = 0.5.

97
5. Quantum Correlations of Exciton-Polaritons

5.2.2 Time-dependent correlations


We now exploit the possibility to resolve the dynamics of polari-
ton correlations with our setup and measure the time-dependent
second-order correlation function g (2) (t1 , t2 ). The absolute time
with respect to the center of the pulse is chosen to be t1 = −70 ps (cf.
Fig. 56 (a)), on the rising edge of the pulse and well before the onset
of photon bunching build-up, yet sufficiently close to the maximum
of the pulse where the coincidence rate is high. We henceforth drop
the absolute time t1 and write g (2) (∆t) where ∆t = t2 − t1 . In
Fig. 58 (a) we show the photon counts of the SFG pulse measured
with the two APD channels of the HBT setup. Note how the input
pulse of only a few picoseconds length is temporally broadened due
to the convolution with the APD IRFs. To reduce the contribution
of dark counts and coincidences due to residual room light, we only
consider detection events with arrival times in a narrow window
denoted by the grey shaded area. We also use the events outside
of this window to check for correct mode-locking and stability of
the pump laser. Figure 58 (b) displays the photon correlation
function g̃ (2) (∆t) for same- and different-period photon pairs (blue
and orange curves, respectively), normalized by the photon counts8 .
Due to interference of the two pump pulses at short delays and
other technical reasons, even the detection of two uncorrelated
photons each stemming from different pump laser cycles does not
yield photon correlations centered around the uncorrelated-state
value g̃ (2) (∆t) = 1. We thus normalize the same-period correlations
(blue solid line) with the average of different-period correlations
(orange lines) to obtain the time-dependent second-order correlation
function g (2) (∆t) displayed in Fig. 58 (c). We observe an increase
of correlations around ∆t = 0. Moreover, we also measure sub-
Poissonian photon statistics beyond the measurement uncertainty
with g (2) (0) = 0.93 ± 0.04. This antibunching signals a violation of

8
This quantity is obtained by g̃ (2) (∆t) = N coin (∆t)
(Ncts /2)2
· T acq · frep , where
Ncoin (∆t) and Ncts are the number of photon pairs for a given delay and total
number of photons, respectively. The factor 2 in the denominator accounts for
the HBT configuration where half of the total photons impinge on each APD.
Tacq is the total acquisition time of the measurement and frep the repetition
rate of the pump laser.

98
5.2. Polariton correlation measurements

Channel 1.2 (b)


400 1 q=0

g̃ (2) (∆t)
Counts
2 1.0
200

(a) 0.8 q>0


0
0 5 10 −50 0 50
t (ns) ∆t (ps)

1.4
g (2) (∆t)

1.2

1.0
(c)
0.8
−60 −40 −20 0 20 40 60
∆t (ps)

4
h↠âi

2 (d)
1.00
(e)
g (2) (∆t)

g (2) 0.98
t2 × γ

20 1.15
1.00 0.96
(f)
0.85
10
10 15 20 25 −10 −5 0 5 10
t1 × γ ∆t × γ

Figure 58: Sub-Poissonian photon statistics. (a) Coincidence counts of the


two APD channels. The SFG input pulse of a few picoseconds length is smeared
out in time to hundreds of picoseconds due to convolution with the APD IRF.
To reduce the influence of dark counts, we only consider coincidences in a time
window indicated by grey shading. (b) Photon correlations for same-period (blue)
and different-period (orange) coincidences. (c) Second-order correlation function
probed around the leading edge of the excitation pulse. The average input power is
700 nW corresponding to 35 µW peak power at 262 ps pulse length. The excitation
laser detuning is ∆/γ = −0.5. Photon antibunching with g (2) (0) = 0.93 ± 0.04 is
observed. (d) Simulated mode population h↠âi for a Gaussian input pulse with
pulse length FWHM 10/γ and peak amplitude F0 /γ = 2. The excitation power
F 2 (t)/γ 2 is displayed as a grey dashed line (same scale). Other parameters are
U /γ = 0.1 and ∆/γ = −0.5. (e) Simulated second-order correlation function as a
function of photon detection times t1 and t2 . (f) Linecuts along dashed colored
lines in (e) showing g (2) (∆t) at fixed t1 where ∆t = t2 − t1 .
99
5. Quantum Correlations of Exciton-Polaritons

both classicality criteria in Eq. (2.32) and constitutes a demonstra-


tion of weakly sub-Poissonian statistics and the quantum character
of exciton-polaritons. Two more features of the correlation function
are surprising. Firstly, also here the statistics seem to converge
towards an overall bunched background at larger delays. While
this is unexpected from a simple Kerr-type nonlinear model which
should converge to Poissonian statistics at large delays, it agrees
well with the data presented earlier in this chapter. Secondly, there
seems to be an asymmetry with respect to the delay time observed
through oscillatory dynamics for positive ∆t. Whereas a steady-
state scenario should exhibit symmetry with respect to zero delay
time, i.e. g (2) (∆t) = g (2) (−∆t) (e.g. in the case of CW excitation),
the pulsed excitation scheme used here breaks this. To understand
the dynamics in more detail, we numerically simulate the evolution
of the system under excitation with a Gaussian pulse. Figure 58 (d)
shows the mode population as a function of time. Since the excita-
tion pulse is longer than the lifetime (we chose its FWHM as 10/γ),
the polariton number closely follows the shape of the excitation
pulse (grey dashed line) with a small deviation on the trailing edge
where the mode population decays exponentially. Figure 58 (e)
presents correlations as a function of the two photon detection times
t1 and t2 , respectively. We notice oscillations between sub- and
super-Poissonian statistics around the antibunched same-time cor-
relations g (2) (t, t). These are similar to relicts of Rabi oscillations:
Due to the moderate excitation power, the mode population is of
order unity for which the effective polariton nonlinearity becomes
significant and alters the dynamics. We confirm that for weaker
excitation powers the oscillations are absent and only the antibunch-
ing at g (2) (t, t) is observed recovering towards uncorrelated photon
emission on a timescale given by the lifetime. As expected, we also
retrieve Poissonian statistics for all times in this dynamical setting
when the nonlinearity is set to zero. In Fig. 58 (f) the correlation
function is shown for a fixed detection time of one of the photons
with respect to the excitation pulse (dashed lines in Fig. 58 (e)), i.e.
we probe photon correlations with respect to a fixed time during
the dynamically evolving population, analogous to the experiment.
In agreement with the data presented in Fig. 58 (c), we notice a
lack of symmetry with respect to ∆t = 0 since the system is not

100
5.3. Difficulties in reproducing photon antibunching

in a steady-state. The dynamics of the numerical simulations also


exhibit oscillations between sub- and super-Poissonian statistics
for positive delays, qualitatively similar to those of the measured
photon correlation function. Although we do not wish to make a
quantitative comparison, our results may hence be interpreted by
a scenario in which the moderate mode population leads to more
complex nonlinearity-induced dynamics on top of the simple anti-
bunching for equal times. Clearly, the simulations do not predict
the experimentally observed convergence towards bunched statistics
at larger delays since this yet to be understood phenomenon is
not included in the simple driven-dissipative Kerr-type nonlinear
polariton model.

5.3 Difficulties in reproducing photon anti-


bunching
Although the presented data demonstrates non-classical photon
statistics beyond the measurement uncertainty, we wish to men-
tion difficulties occurring during data acquisition. In particular,
we observe stochastic and abrupt changes of the polariton mode
wavelength on the timescale of hours, with amplitudes on the order
of the mode linewidth. During the measurement, we periodically
scan the resonance to track its center wavelength and compen-
sate for any long-term drifts, e.g. due to creeping of the piezo
crystal-based nanopositioner that sets the cavity length. In this
way, the desired laser detuning is maintained for the duration of
the measurement, which might be up to several days. Yet, the
presence of mode jumps seems to be detrimental for the observation
of sub-Poissonian photon statistics, and any non-classical signature
vanishes once they occur. In the data shown in Fig. 58 these mode
jumps were absent and the center wavelength, and hence also the
excitation laser wavelength, remained constant throughout the in-
tegration time. We have not yet identified the physical origin of
the jumps in mode energy and can only speculate about possible
reasons, e.g. long-term fluctuations of the charge configuration in
the vicinity of the quantum well. From a measurement perspective,
it is highly desirable to understand what causes these changes in

101
5. Quantum Correlations of Exciton-Polaritons

mode energy and how they affect the photon statistics in order to
either avoid them or develop a measurement strategy to keep the
system in the state that preserves the desired polariton statistics
instead of relying on post-selection.

5.4 Conclusion and perspectives


The measurements presented in this chapter constitute a manifes-
tation of quantum polariton dynamics at the single-particle level.
Based on the extracted parameters, our system is in between the
weakly interacting mean-field and the blockade regime characterized
by nonlinearities comparable to or in excess of the intrinsic decay
rate [78]. Reaching the latter requires increasing the polariton life-
time by an order of magnitude, or, alternatively, reducing the mode
waist by a factor of 3. A sizable increase in polariton confinement
is unlikely with the current fiber cavity setup and would require e.g.
coupling excitons to other types of cavities such as photonic crystal
defects [212]. Significantly increasing the cavity lifetime, however,
seems within reach with small changes of the sample design (cf.
the discussion in Chapter 3). In addition, the exciton-exciton in-
teraction could be increased by different mechanisms. Coupling
exciton-polaritons to a two-dimensional electron gas introduces
additional interactions due to the delocalized many-body state of
the Fermi sea of electrons [213–215]. A different approach relies
on strong dipolar interactions of polar excitons where a permanent
dipole is induced via an electric field [216], or where the electron and
hole wavefunctions are delocalized in two different QWs [217, 218].
We expect that these techniques, in particular in combination with
an increase of the cavity lifetime, lead to a significant enhancement
of the quantum correlations.
Our experiments hence serve as a precursor into the direction of
strongly quantum-correlated polaritons with potential applications
as bright single-photon sources [78], and single-photon all-optical
quantum gates. Coupling many such polariton boxes with strong
interactions paves the way for polariton-based driven-dissipative
quantum simulators and research on more fundamental phenomena
such as photon crystallization [219, 220] or the quantum phases

102
5.4. Conclusion and perspectives

of XY models [221] and insulator-superfluid transitions [185, 222].


From a more technical point of view, identifying the origin of the
mode switching, or at least developing a measurement strategy to
suppress them, is crucial since it currently limits the reproducibility
of our results and hinders a systematic analysis of the parameter
space.
Our results are closely related to experiments by Muñoz-Matutano
et al. [210] where a pulsed excitation scheme has been used to
demonstrate a degree of antibunching similar to our experiments.
Analogous to the data presented here, a reduction of the polari-
ton lifetime and emergence of biexponential population decay have
been reported, and their explanation by elastic scattering into a
long-lived reservoir of localized excitonic states [211] seems plau-
sible also for our experiment. They also seem to observe a trend
towards a bunching offset for increasing exciton content. Yet, we
cannot confirm whether the overall long-lived bunched background
at large photon delays occurs in their experiment as well since their
measurement technique only allows for measurement of same-time
correlations g (2) (0). Hence, future research needs to elucidate the
origin of the bunching background.

103
Chapter 6

Conclusion and Outlook

In this dissertation, we studied the interactions of microcavity


exciton-polaritons. CO2 -laser based fabrication of a fiber cavity
setup allowed us to create zero-dimensional polariton boxes with
strong in-plane confinement. Utilizing photon correlations as an
analysis tool, we investigated polariton interactions in two different
regimes: For large populations, the nonlinearity gives rise to a
critical slowing down of the intrinsic dynamics as the polaritons
are driven across a first-order dissipative phase transition. The
measurement of photon correlations has been shown to be intimately
linked to the Liouvillian eigenspectrum. This proof-of-principle
experiment enables further research on quantum phase transitions
and lends itself to study quantum many-body phases of polariton
lattices in a straightforward manner. In the single-particle regime,
the onset of quantum correlations between single polaritons has been
observed. Albeit technical difficulties limiting the reproducibility
and an undesired scattering mechanism of polaritons into localized
exciton states, this is a promising result demonstrating the potential
for exciton-polariton quantum dynamics.

An extension of our work is the observation of unconventional


photon blockade recently reported in the microwave domain [223].
In contrast to the blockade mechanism presented in Section 2.7,
this concept exploits quantum interference between two coupled
modes, out of which at least one is slightly nonlinear, to generate

105
6. Conclusion and Outlook

coupling J 2
(a) (b)
4

...

...
n=2 1

(Ep − E)/γ
3

g11 (τ )
U 0

(2)
n=1 2
−1 1
n=0
−2 0
mode 1 mode 2 −2.5 0.0 2.5
τ /(2π/J)

Figure 61: Unconventional blockade. (a) Schematic of unconventional photon


blockade. Two resonator modes, out of which only one is excited and detected,
are coupled to each other with strength J. At least one of them has a slightly
anharmonic spectrum due to a nonlinearity U . Tuning the system parameters
carefully leads to destructive interference of the population in the n = 2 manifold of
mode 1. This interference gives rise to strong oscillatory deviations from coherent
photon statistics shown in (b) for J/γ = 3, U /γ = 0.0428, and F /γ = 0.1, as a
function of delay between the photons and excitation laser detuning.

nonclassical photon statistics. As illustrated in Fig. 61, careful


tuning of the system parameters allows to suppress the population in
the two-polariton manifold of one of the modes, which leads to strong
antibunching signatures for nonlinearities as small as U /γ . 0.05
[224, 225]. Moreover, the coupling causes oscillations between
strongly antibunched and bunched photon statistics with a time
scale given by the mode coupling. The linear polarization splitting
in our system can be mapped exactly onto such a configuration,
and the experimental parameters are close to the ideal ones for
this approach. The ability to obtain strongly antibunched photon
correlations with only moderate nonlinearities as well as the short
polariton lifetime compared to, e.g., quantum dots opens up possible
applications as ultra-bright single-photon sources. Adiyatullin et
al. [226] reported oscillatory mode coupling signatures in photon
correlation measurements, but the low duty cycle of the streak
camera-based measurement [202, 203] prevented access to a regime
in which low populations enable non-classical photon statistics.
Unconventional blockade can be understood as a special case of the
interplay between squeezing and displacement [227] and different

106
excitation and detection schemes have been proposed recently [228]
including an approach based on planar microcavities [229].
Our results suggest that the single-particle exciton-polariton interac-
tions in our system are significant, yet about one order of magnitude
smaller than the intrinsic decay rate and thus just sufficient to ob-
serve the onset of quantum behavior. In order to be usable as a
building block for a quantum simulator, however, the ratio of these
parameters should be of order unity to enable quantum correlations
between single photons at different units. Future efforts should thus
be dedicated to increasing both polariton lifetime and interactions,
and the perspectives discussed in Sections 3.7 and 5.4 present viable
approaches to this end. In addition to enabling strong quantum
correlations, a second direction of future work is the scaling of the
system: Whereas we focused on a single exciton-polariton mode in
this dissertation, coupling many of these to form a lattice would ul-
timately allow for strongly correlated many-body phenomena. The
CO2 -based fabrication technique used here is likely to be limited in
terms of scalability, yet FIB machining allows to create large arrays
of tunnel-coupled microcavities and provides a similar degree of
control as wet etching [103, 109]. Such systems would not only put
solid state driven-dissipative exciton-polariton quantum simulators
within reach, but also allow for more fundamental research. The
combination of strongly correlated many-body phenomena with the
non-unitary dynamics of dissipative systems has recently attracted
a lot of attention due to the novel phenomena that emerge from
this interplay [125, 219, 230–235]. Scaling up the exciton-polariton
single-site system considered in this dissertation would constitute
an exquisite testbed to study such yet unexplored and complex
nonequilibrium physics.

107
Appendix A

Sample Characterization

The sample used in this dissertation (labeled C4114) is MBE-grown


by A. Schade, S. Höfling, and C. Schneider at the Universität
Würzburg. Figure A1 displays the composition of the full sample
structure. It consists of a 17 nm-wide In4 % Ga96 % As QW in the
center of a GaAs λ/n spacer layer on top of an AlAs / GaAs DBR
with 35.5 bilayers. Figure A2 (a) displays photoluminescence (PL)
spectra of the bare sample under helium neon (HeNe) (633 nm)
and diode laser (780 nm) excitation. The extracted PL emission
linewidths (FWHM) are 590 µeV and 568 µeV, respectively, in good
agreement with each other. In Fig. A2 (b) we show the trans-
mitted light spectra of a lower polariton with negligible photon
content (cavity blue-detuned by 14 meV). The linewidth of 220 µeV
extracted from a Lorentzian fit is significantly lower compared to
the bare PL emission, likely due to the additional spectral and
spatial filtering by the cavity. Note that here we only use HeNe
excitation since light at 780 nm is reflected by the stopband. The
transmission under white light excitation shows additional features
at slightly lower energy than the exciton which correspond to cavity
modes with a high transverse mode number belonging to a different
longitudinal mode manifold.

109
A. Sample Characterization

Layer Thickness (nm) Material


Spacer top 112.5 GaAs
Quantum well 17.0 InGaAs
Spacer bottom 112.5 GaAs
DBR 4 679.5
73.2 AlAs
35×
60.5 GaAs
DBR cap 73.2 AlAs
Buffer 240.0 GaAs
Total: 5 234.7
Substrate 500 µm GaAs

Figure A1: Sample structure. Structure of the sample C4114 illustrated on the
left. The respective layers as well as thicknesses and materials are indicated in the
table on the right.

(a) HeNe HeNe (b)


1.0 780 nm WL
norm. intensity

0.5

0.0
1.468 1.470 1.472 1.474 1.468 1.470 1.472 1.474
Energy (eV) Energy (eV)

Figure A2: Exciton emission spectrum. (a) PL excitation of the QW exciton


with HeNe (blue dots) and 780 nm (orange dots) light. The input power in both
cases is less than 1 nW. The green lines denote Lorentzian fits with extracted
linewidths (FWHM) of 590 µeV and 568 µeV, respectively. (b) Transmitted light
of an exciton-like lower polariton under HeNe (blue dots) and white light (orange
dots) excitation. The green lines denote Lorentzian fits with linewidths 220 µeV
and 369 µeV, respectively. The additional features around 1.470 eV for white light
excitation are transverse modes with high transverse mode numbers of a different
longitudinal cavity mode manifold. The orange curves are offset in both panels for
clarity. Note the small energy shift between (a) and (b) due to the slightly different
wavelength calibration of the two spectrometers used.

110
Bibliography

[1] R. P. Feynman, “Simulating physics with computers,” Inter-


national Journal of Theoretical Physics 21, 467–488 (1982).
[2] S. Lloyd, “Universal Quantum Simulators,” Science 273,
1073–1078 (1996).
[3] I. Buluta and F. Nori, “Quantum Simulators,” Science 326,
108–111 (2009).
[4] J. Orenstein and A. J. Millis, “Advances in the Physics of
High-Temperature Superconductivity,” Science 288, 468–474
(2000).
[5] E. Dagotto, “Correlated electrons in high-temperature super-
conductors,” Review of Modern Physics 66, 763–840 (1994).
[6] D. P. DiVincenzo, “The physical implementation of quantum
computation,” Fortschritte der Physik 48, 771–783 (2000).
[7] J. I. Cirac and P. Zoller, “Goals and opportunities in quantum
simulation,” Nature Physics 8, 264 (2012).
[8] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Omran,
et al., “Probing many-body dynamics on a 51-atom quantum
simulator,” Nature 551, 579 (2017).
[9] S. Trotzky, Y.-A. Chen, A. Flesch, I. P. McCulloch, U.
Schollwöck, J. Eisert, and I. Bloch, “Probing the relax-
ation towards equilibrium in an isolated strongly correlated
one–dimensional Bose gas,” Nature Physics 8, 325 (2012).

i
Bibliography

[10] A. Mazurenko, C. S. Chiu, G. Ji, M. F. Parsons, M. Kanász-


Nagy, R. Schmidt, F. Grusdt, E. Demler, D. Greif, and M.
Greiner, “A cold-atom Fermi-Hubbard antiferromagnet,”
Nature 545, 462 (2017).
[11] I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simula-
tion,” Review of Modern Physics 86, 153–185 (2014).
[12] C. Weisbuch, M. Nishioka, A. Ishikawa, and Y. Arakawa,
“Observation of the coupled exciton-photon mode splitting
in a semiconductor quantum microcavity,” Physical Review
Letters 69, 3314–3317 (1992).
[13] A. İmamoğlu, R. J. Ram, S. Pau, and Y. Yamamoto,
“Nonequilibrium condensates and lasers without inversion:
Exciton-polariton lasers,” Physical Review A 53, 4250–4253
(1996).
[14] J. Kasprzak, M. Richard, S. Kundermann, A. Baas, P. Jeam-
brun, et al., “Bose-Einstein condensation of exciton polari-
tons,” Nature 443, 409–414 (2006).
[15] R. M. Stevenson, V. N. Astratov, M. S. Skolnick, D. M. Whit-
taker, M. Emam-Ismail, A. I. Tartakovskii, P. G. Savvidis,
J. J. Baumberg, and J. S. Roberts, “Continuous Wave Ob-
servation of Massive Polariton Redistribution by Stimulated
Scattering in Semiconductor Microcavities,” Physical Review
Letters 85, 3680–3683 (2000).
[16] R. Balili, V. Hartwell, D. Snoke, L. Pfeiffer, and K. West,
“Bose-Einstein Condensation of Microcavity Polaritons in a
Trap,” Science 316, 1007–1010 (2007).
[17] J. D. Plumhof, T. Stöferle, L. Mai, U. Scherf, and R. F.
Mahrt, “Room-temperature Bose-Einstein condensation of
cavity exciton-polaritons in a polymer,” Nature Materials
13, 247 (2013).
[18] W. H. Nitsche, N. Y. Kim, G. Roumpos, C. Schneider, M.
Kamp, S. Höfling, A. Forchel, and Y. Yamamoto, “Algebraic
order and the Berezinskii-Kosterlitz-Thouless transition in
an exciton-polariton gas,” Physical Review B 90, 205430
(2014).

ii
Bibliography

[19] K. S. Daskalakis, S. A. Maier, and S. Kéna-Cohen, “Spatial


Coherence and Stability in a Disordered Organic Polariton
Condensate,” Physical Review Letters 115, 035301 (2015).
[20] A. Amo, J. Lefrère, S. Pigeon, C. Adrados, C. Ciuti, I. Caru-
sotto, R. Houdré, E. Giacobino, and A. Bramati, “Superflu-
idity of polaritons in semiconductor microcavities,” Nature
Physics 5, 805 (2009).
[21] A. Amo, D. Sanvitto, F. P. Laussy, D. Ballarini, E. d. Valle,
et al., “Collective fluid dynamics of a polariton condensate
in a semiconductor microcavity,” Nature 457, 291 (2009).
[22] K. G. Lagoudakis, M. Wouters, M. Richard, A. Baas, I.
Carusotto, R. André, L. S. Dang, and B. Deveaud-Plédran,
“Quantized vortices in an exciton-polariton condensate,” Na-
ture Physics 4, 706 (2008).
[23] D. Sanvitto, F. M. Marchetti, M. H. Szymanska, G. Tosi, M.
Baudisch, et al., “Persistent currents and quantized vortices
in a polariton superfluid,” Nature Physics 6, 527 (2010).
[24] C. Schneider, K. Winkler, M. D. Fraser, M. Kamp, Y.
Yamamoto, E. A. Ostrovskaya, and S. Höfling, “Exciton-
polariton trapping and potential landscape engineering,”
Reports on Progress in Physics 80, 016503 (2017).
[25] D. Sanvitto and S. Kéna-Cohen, “The road towards polari-
tonic devices,” Nature Materials 15, 1061 (2016).
[26] A. Amo and J. Bloch, “Exciton-polaritons in lattices: A non-
linear photonic simulator,” Comptes Rendus Physique 17,
934–945 (2016).
[27] P. Y. Yu and M. Cardona, Fundamentals of Semiconductors,
4th ed. (Springer-Verlag, 2010).
[28] M. E. Levinshtein and S. L. Rumyantsev, “GALLIUM AR-
SENIDE (GaAs),” in Handbook series on semiconductor
parameters (World Scientific, 1996), pp. 77–103.
[29] M. P. Mikhailova, “INDIUM ARSENIDE (InAs),” in Hand-
book series on semiconductor parameters (World Scientific,
1996), pp. 147–168.

iii
Bibliography

[30] Y. A. Goldberg and N. M. Shmidt, “GALLIUM INDIUM


ARSENIDE (Gax In1−x As),” in Handbook series on semicon-
ductor parameters (World Scientific, 1996), pp. 62–88.
[31] J. M. Luttinger, “Quantum Theory of Cyclotron Resonance
in Semiconductors: General Theory,” Physical Review 102,
1030–1041 (1956).
[32] W. W. Chow, S. W. Koch, and M. I. Sargent, Semiconductor-
Laser Physics (Springer-Verlag, 1994).
[33] G. Sęk, J. Misiewicz, D. Radziewicz, M. Tłaczała, M. Panek,
and R. Korbutowicz, “Study of the nature of light hole exci-
tonic transitions in InGaAs/GaAs quantum well,” Vacuum
50, 199–201 (1998).
[34] H. Deng, H. Haug, and Y. Yamamoto, “Exciton-polariton
Bose-Einstein condensation,” Review of Modern Physics 82,
1489–1537 (2010).
[35] S. Schmitt-Rink, D. S. Chemla, and D. A. B. Miller, “Linear
and nonlinear optical properties of semiconductor quantum
wells,” Advances in Physics 38, 89–188 (1989).
[36] N. F. Mott, “The Basis of the Electron Theory of Metals, with
Special Reference to the Transition Metals,” Proceedings of
the Physical Society. Section A 62, 416 (1949).
[37] S. Schmitt-Rink, D. S. Chemla, and D. A. B. Miller, “Theory
of transient excitonic optical nonlinearities in semiconductor
quantum-well structures,” Physical Review B 32, 6601–6609
(1985).
[38] G. Rochat, C. Ciuti, V. Savona, C. Piermarocchi, A. Quat-
tropani, and P. Schwendimann, “Excitonic Bloch equations
for a two-dimensional system of interacting excitons,” Physi-
cal Review B 61, 13856–13862 (2000).
[39] M. Combescot and C. Tanguy, “New criteria for bosonic
behavior of excitons,” Europhysics Letters 55, 390 (2001).
[40] M. Combescot, O. Betbeder-Matibet, and F. Dubin, “The
many-body physics of composite bosons,” Physics Reports
463, 215–320 (2008).

iv
Bibliography

[41] L. Kappei, J. Szczytko, F. Morier-Genoud, and B. Deveaud,


“Direct Observation of the Mott Transition in an Optically
Excited Semiconductor Quantum Well,” Physical Review
Letters 94, 147403 (2005).
[42] D. I. Babic and S. W. Corzine, “Analytic expressions for
the reflection delay, penetration depth, and absorptance of
quarter-wave dielectric mirrors,” IEEE Journal of Quantum
Electronics 28, 514–524 (1992).
[43] G. Panzarini, L. C. Andreani, A. Armitage, D. Baxter, M. S.
Skolnick, V. N. Astratov, J. S. Roberts, A. V. Kavokin, M. R.
Vladimirova, and M. A. Kaliteevski, “Exciton-light coupling
in single and coupled semiconductor microcavities: Polariton
dispersion and polarization splitting,” Physical Review B 59,
5082–5089 (1999).
[44] V. Savona, “Linear optical properties of semiconductor micro-
cavities with embedded quantum wells,” in Confined photon
systems, edited by H. Benisty, J.-M. Gérard, R. Houdré, J.
Rarity, and C. Weisbuch (Springer, 1999), pp. 173–242.
[45] N. Ismail, C. C. Kores, D. Geskus, and M. Pollnau, “Fabry-
Pérot resonator: spectral line shapes, generic and related
Airy distributions, linewidths, finesses, and performance at
low or frequency-dependent reflectivity,” Optics Express 24,
16366–16389 (2016).
[46] M. Born and E. Wolf, Principles of optics: electromagnetic
theory of propagation, interference and diffraction of light
(Pergamon Press, 1964).
[47] O. S. Heavens, Optical Properties of Thin Films (Butter-
worth, 1955).
[48] B. Besga, C. Vaneph, J. Reichel, J. Estève, A. Reinhard,
J. Miguel-Sánchez, A. İmamoğlu, and T. Volz, “Polariton
Boxes in a Tunable Fiber Cavity,” Physical Review Applied
3, 014008 (2015).

v
Bibliography

[49] S. Reitzenstein, C. Hofmann, A. Gorbunov, M. Strauß, S. H.


Kwon, C. Schneider, A. Löffler, S. Höfling, M. Kamp, and
A. Forchel, “AlAs/GaAs micropillar cavities with quality fac-
tors exceeding 150.000,” Applied Physics Letters 90, 251109
(2007).
[50] I. Carusotto and C. Ciuti, “Quantum fluids of light,” Review
of Modern Physics 85, 299–366 (2013).
[51] L. Claudio Andreani, “Optical Transitions, Excitons, and Po-
laritons in Bulk and Low-Dimensional Semiconductor Struc-
tures,” in Confined Electrons and Photons: New Physics
and Applications, edited by E. Burstein and C. Weisbuch
(Springer US, Boston, MA, 1995), pp. 57–112.
[52] L. C. Andreani and A. Pasquarello, “Accurate theory of
excitons in GaAs-Ga1 − x Alx As quantum wells,” Physical
Review B 42, 8928–8938 (1990).
[53] R. Houdré, R. P. Stanley, U. Oesterle, M. Ilegems, and
C. Weisbuch, “Room-temperature cavity polaritons in a
semiconductor microcavity,” Physal Review B 49, 16761–
16764 (1994).
[54] M. S. Skolnick, T. A. Fisher, and D. M. Whittaker, “Strong
coupling phenomena in quantum microcavity structures,”
Semiconductor Science and Technology 13, 645 (1998).
[55] J. J. Hopfield, “Theory of the Contribution of Excitons to the
Complex Dielectric Constant of Crystals,” Physical Review
112, 1555–1567 (1958).
[56] D. M. Whittaker, P. Kinsler, T. A. Fisher, M. S. Skolnick,
A. Armitage, A. M. Afshar, M. D. Sturge, and J. S. Roberts,
“Motional Narrowing in Semiconductor Microcavities,” Phys-
ical Review Letters 77, 4792–4795 (1996).
[57] R. Houdré, R. P. Stanley, and M. Ilegems, “Vacuum-field
Rabi splitting in the presence of inhomogeneous broadening:
Resolution of a homogeneous linewidth in an inhomoge-
neously broadened system,” Physical Review A 53, 2711–
2715 (1996).

vi
Bibliography

[58] M. Tavis and F. W. Cummings, “Exact Solution for an N -


Molecule–Radiation-Field Hamiltonian,” Physical Review
170, 379–384 (1968).
[59] R. H. Dicke, “Coherence in Spontaneous Radiation Pro-
cesses,” Physical Review 93, 99–110 (1954).
[60] E. T. Jaynes and F. W. Cummings, “Comparison of Quantum
and Semiclassical Radiation Theories with Application to the
Beam Maser,” Proceedings of the IEEE 51, 89–109 (1963).
[61] R. W. Boyd, Nonlinear Optics, 3rd ed. (Elsevier Science,
2008).
[62] M. Dinu, F. Quochi, and H. Garcia, “Third-order nonlin-
earities in silicon at telecom wavelengths,” Applied Physics
Letters 82, 2954–2956 (2003).
[63] C. Ciuti, P. Schwendimann, and A. Quattropani, “Theory
of polariton parametric interactions in semiconductor micro-
cavities,” Semiconductor Science and Technology 18, S279
(2003).
[64] C. Ciuti, V. Savona, C. Piermarocchi, A. Quattropani, and
P. Schwendimann, “Role of the exchange of carriers in elastic
exciton-exciton scattering in quantum wells,” Physial Review
B 58, 7926–7933 (1998).
[65] L. Ferrier, E. Wertz, R. Johne, D. D. Solnyshkov, P. Senel-
lart, I. Sagnes, A. Lemaı̂tre, G. Malpuech, and J. Bloch,
“Interactions in Confined Polariton Condensates,” Physical
Review Letters 106, 126401 (2011).
[66] S. R. K. Rodriguez, A. Amo, I. Sagnes, L. Le Gratiet, E.
Galopin, A. Lemaître, and J. Bloch, “Interaction-induced
hopping phase in driven-dissipative coupled photonic micro-
cavities,” Nature Communications 7, 11887 (2016).
[67] S. R. K. Rodriguez, W. Casteels, F. Storme, N. Carlon Zam-
bon, I. Sagnes, et al., “Probing a Dissipative Phase Transition
via Dynamical Optical Hysteresis,” Physical Review Letters
118, 247402 (2017).

vii
Bibliography

[68] Y. Sun, Y. Yoon, M. Steger, G. Liu, L. N. Pfeiffer, K. West,


D. W. Snoke, and K. A. Nelson, “Direct measurement of
polariton-polariton interaction strength,” Nature Physics 13,
870 (2017).
[69] I. Carusotto, T. Volz, and A. İmamoğlu, “Feshbach blockade:
Single-photon nonlinear optics using resonantly enhanced
cavity polariton scattering from biexciton states,” Euro-
physics Letters 90, 37001 (2010).
[70] C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, “Feshbach
resonances in ultracold gases,” Review of Modern Physics
82, 1225–1286 (2010).
[71] N. Takemura, S. Trebaol, M. Wouters, M. T. Portella-Oberli,
and B. Deveaud, “Polaritonic Feshbach resonance,” Nature
Physics 10, 500 (2014).
[72] N. Takemura, S. Trebaol, M. Wouters, M. T. Portella-Oberli,
and B. Deveaud, “Heterodyne spectroscopy of polariton
spinor interactions,” Physical Review B 90, 195307 (2014).
[73] E. Wertz, L. Ferrier, D. D. Solnyshkov, R. Johne, D. Sanvitto,
et al., “Spontaneous formation and optical manipulation
of extended polariton condensates,” Nature Physics 6, 860
(2010).
[74] A. Askitopoulos, H. Ohadi, A. V. Kavokin, Z. Hatzopoulos,
P. G. Savvidis, and P. G. Lagoudakis, “Polariton conden-
sation in an optically induced two-dimensional potential,”
Physical Review B 88, 041308 (2013).
[75] R. J. Glauber, “The Quantum Theory of Optical Coherence,”
Physical Review 130, 2529–2539 (1963).
[76] R. H. Brown and R. Twiss, “A new type of interferometer
for use in radio astronomy,” The London, Edinburgh, and
Dublin Philosophical Magazine and Journal of Science 45,
663–682 (1954).
[77] R. Hanbury Brown and R. Q. Twiss, “Correlation between
Photons in two Coherent Beams of Light,” Nature 177, 27
(1956).

viii
Bibliography

[78] A. Verger, C. Ciuti, and I. Carusotto, “Polariton quantum


blockade in a photonic dot,” Physical Review B 73, 193306
(2006).
[79] M. J. Werner and A. İmamoğlu, “Photon-photon interactions
in cavity electromagnetically induced transparency,” Physical
Review A 61, 011801 (1999).
[80] S. Ferretti and D. Gerace, “Single-photon nonlinear optics
with Kerr-type nanostructured materials,” Physical Review
B 85, 033303 (2012).
[81] Y. Colombe, T. Steinmetz, G. Dubois, F. Linke, D. Hunger,
and J. Reichel, “Strong atom-field coupling for Bose-Einstein
condensates in an optical cavity on a chip,” Nature 450, 272
(2007).
[82] J. Volz, R. Gehr, G. Dubois, J. Estève, and J. Reichel, “Mea-
surement of the internal state of a single atom without energy
exchange,” Nature 475, 210 (2011).
[83] F. Haas, J. Volz, R. Gehr, J. Reichel, and J. Estève, “En-
tangled States of More Than 40 Atoms in an Optical Fiber
Cavity,” Science 344, 180–183 (2014).
[84] M. Steiner, H. M. Meyer, C. Deutsch, J. Reichel, and M. Köhl,
“Single Ion Coupled to an Optical Fiber Cavity,” Physical
Review Letters 110, 043003 (2013).
[85] M. Steiner, H. M. Meyer, J. Reichel, and M. Köhl, “Photon
Emission and Absorption of a Single Ion Coupled to an
Optical-Fiber Cavity,” Physical Review Letters 113, 263003
(2014).
[86] S. Dufferwiel, F. Fras, A. Trichet, P. M. Walker, F. Li, et
al., “Strong exciton-photon coupling in open semiconductor
microcavities,” Applied Physics Letters 104, 192107 (2014).
[87] A. Muller, E. B. Flagg, M. Metcalfe, J. Lawall, and G. S.
Solomon, “Coupling an epitaxial quantum dot to a fiber-
based external-mirror microcavity,” Applied Physics Letters
95, 173101 (2009).

ix
Bibliography

[88] J. Miguel-Sánchez, A. Reinhard, E. Togan, T. Volz, A.


İmamoğlu, B. Besga, J. Reichel, and J. Estève, “Cavity
quantum electrodynamics with charge-controlled quantum
dots coupled to a fiber Fabry–Perot cavity,” New Journal of
Physics 15, 045002 (2013).
[89] L. Greuter, S. Starosielec, A. V. Kuhlmann, and R. J. War-
burton, “Towards high-cooperativity strong coupling of a
quantum dot in a tunable microcavity,” Physical Review B
92, 045302 (2015).
[90] H. Kaupp, C. Deutsch, H.-C. Chang, J. Reichel, T. W. Hän-
sch, and D. Hunger, “Scaling laws of the cavity enhancement
for nitrogen-vacancy centers in diamond,” Physical Review
A 88, 053812 (2013).
[91] R. Albrecht, A. Bommer, C. Deutsch, J. Reichel, and C.
Becher, “Coupling of a single nitrogen-vacancy center in dia-
mond to a fiber-based microcavity,” Physical Review Letters
110, 243602 (2013).
[92] J. Benedikter, H. Kaupp, T. Hümmer, Y. Liang, A. Bommer,
C. Becher, A. Krueger, J. M. Smith, T. W. Hänsch, and D.
Hunger, “Cavity-Enhanced Single-Photon Source Based on
the Silicon-Vacancy Center in Diamond,” Physical Review
Applied 7, 024031 (2017).
[93] S. Bogdanović, M. S. Z. Liddy, S. B. van Dam, L. C. Co-
enen, T. Fink, M. Lončar, and R. Hanson, “Robust nano-
fabrication of an integrated platform for spin control in a
tunable microcavity,” APL Photonics 2, 126101 (2017).
[94] N. E. Flowers-Jacobs, S. W. Hoch, J. C. Sankey, A.
Kashkanova, A. M. Jayich, C. Deutsch, J. Reichel, and
J. G. E. Harris, “Fiber-cavity-based optomechanical device,”
Applied Physics Letters 101, 221109 (2012).
[95] S. Stapfner, L. Ost, D. Hunger, J. Reichel, I. Favero, and
E. M. Weig, “Cavity-enhanced optical detection of carbon
nanotube Brownian motion,” Applied Physics Letters 102,
151910 (2013).

x
Bibliography

[96] A. B. Shkarin, N. E. Flowers-Jacobs, S. W. Hoch, A. D.


Kashkanova, C. Deutsch, J. Reichel, and J. G. E. Harris,
“Optically Mediated Hybridization between Two Mechanical
Modes,” Physical Review Letters 112, 013602 (2014).
[97] A. D. Kashkanova, A. B. Shkarin, C. D. Brown, N. E. Flowers-
Jacobs, L. Childress, S. W. Hoch, L. Hohmann, K. Ott, J.
Reichel, and J. G. E. Harris, “Superfluid Brillouin optome-
chanics,” Nature Physics 13, 74 (2016).
[98] A. D. Kashkanova, A. B. Shkarin, C. D. Brown, N. E. Flowers-
Jacobs, L. Childress, S. W. Hoch, L. Hohmann, K. Ott, J.
Reichel, and J. G. E. Harris, “Optomechanics in superfluid
helium coupled to a fiber-based cavity,” Journal of Optics
19, 034001 (2017).
[99] D. M. Coles, Y. Yang, Y. Wang, R. T. Grant, R. A. Taylor,
S. K. Saikin, A. Aspuru-Guzik, D. G. Lidzey, J. K.-H. Tang,
and J. M. Smith, “Strong coupling between chlorosomes of
photosynthetic bacteria and a confined optical cavity mode,”
Nature Communications 5, 5561 (2014).
[100] S. Schwarz, S. Dufferwiel, P. M. Walker, F. Withers, A. A. P.
Trichet, et al., “Two-Dimensional Metal–Chalcogenide Films
in Tunable Optical Microcavities,” Nano Letters 14, 7003–
7008 (2014).
[101] S. Dufferwiel, S. Schwarz, F. Withers, A. A. P. Trichet, F. Li,
et al., “Exciton-polaritons in van der Waals heterostructures
embedded in tunable microcavities,” Nature Communica-
tions 6, 8579 (2015).
[102] M. Sidler, P. Back, O. Cotlet, A. Srivastava, T. Fink, M. Kro-
ner, E. Demler, and A. İmamoğlu, “Fermi polaron-polaritons
in charge-tunable atomically thin semiconductors,” Nature
Physics 13, 255–261 (2016).
[103] S. Dufferwiel, F. Li, A. A. P. Trichet, L. Giriunas, P. M.
Walker, I. Farrer, D. A. Ritchie, J. M. Smith, M. S. Skolnick,
and D. N. Krizhanovskii, “Tunable polaritonic molecules in
an open microcavity system,” Applied Physics Letters 107,
201106 (2015).

xi
Bibliography

[104] A. Reinhard, Strong photon-photon interactions in solid state


cavity QED, PhD thesis (ETH Zürich, 2013).
[105] S. Bogdanović, S. B. van Dam, C. Bonato, L. C. Coenen,
A.-M. J. Zwerver, et al., “Design and low-temperature char-
acterization of a tunable microcavity for diamond-based
quantum networks,” Applied Physics Letters 110, 171103
(2017).
[106] E. R. Abraham and E. A. Cornell, “Teflon feedthrough
for coupling optical fibers into ultrahigh vacuum systems,”
Applied Optics 37, 1762–1763 (1998).
[107] T. Steinmetz, Y. Colombe, D. Hunger, T. W. Hänsch,
A. Balocchi, R. J. Warburton, and J. Reichel, “Stable
fiber-based Fabry-Pérot cavity,” Applied Physics Letters 89,
111110 (2006).
[108] J. Orloff, L. Swanson, and M. Utlaut, High Resolution Fo-
cused Ion Beams: FIB and its Applications, The Physics of
Liquid Metal Ion Sources and Ion Optics and Their Applica-
tion to Focused Ion Beam Technology (Springer Science &
Business Media, 2003).
[109] P. R. Dolan, G. M. Hughes, F. Grazioso, B. R. Patton, and
J. M. Smith, “Femtoliter tunable optical cavity arrays,”
Optics Letters 35, 3556–3558 (2010).
[110] D. Hunger, Herstellung und Charakterisierung von
Faserresonatoren hoher Finesse, Diploma thesis (Ludwig-
Maximilians-Universität München, 2008).
[111] C. Deutsch, High Finesse Fibre Fabry-Perot Resonators -
Production, Characterization, and Applications, Diploma
thesis (Ludwig-Maximilians-Universität München, 2008).
[112] D. Hunger, T. Steinmetz, Y. Colombe, C. Deutsch, T. W.
Hänsch, and J. Reichel, “A fiber Fabry–Perot cavity with
high finesse,” New Journal of Physics 12, 065038 (2010).
[113] D. Hunger, C. Deutsch, R. J. Barbour, R. J. Warburton, and
J. Reichel, “Laser micro-fabrication of concave, low-roughness
features in silica,” AIP Advances 2, 012119 (2012).

xii
Bibliography

[114] J. Benedikter, T. Hümmer, M. Mader, B. Schlederer, J.


Reichel, T. W. Hänsch, and D. Hunger, “Transverse-mode
coupling and diffraction loss in tunable Fabry–Pérot micro-
cavities,” New Journal of Physics 17, 053051 (2015).
[115] H. S. Kaupp, Coupling nitrogen-vacancy centers in diamond
to fiber-based Fabry-Pérot microcavities, PhD thesis (Ludwig-
Maximilians-Universität München, Oct. 2017).
[116] L. Greuter, S. Starosielec, D. Najer, A. Ludwig, L. Duem-
pelmann, D. Rohner, and R. J. Warburton, “A small mode
volume tunable microcavity: Development and characteriza-
tion,” Applied Physics Letters 105, 121105 (2014).
[117] J. M. Bennett, “Recent developments in surface roughness
characterization,” Measurement Science and Technology 3,
1119 (1992).
[118] D. M. Sullivan, Electromagnetic simulation using the FDTD
method (John Wiley & Sons, 2013).
[119] G.-M. Schnüringer, Photon lifetimes in fibre-based cavities
with small radii of curvature, Bachelor thesis (ETH Zürich,
2015).
[120] V. Savona, F. Tassone, C. Piermarocchi, A. Quattropani, and
P. Schwendimann, “Theory of polariton photoluminescence
in arbitrary semiconductor microcavity structures,” Physical
Review B 53, 13051–13062 (1996).
[121] A. Trichet, P. R. Dolan, and J. M. Smith, “Strong coupling
between 0D and 2D modes in optical open microcavities,”
Journal of Optics 20, 035402 (2018).
[122] M. Steger, G. Liu, B. Nelsen, C. Gautham, D. W. Snoke,
R. Balili, L. Pfeiffer, and K. West, “Long-range ballistic mo-
tion and coherent flow of long-lifetime polaritons,” Physical
Review B 88, 235314 (2013).
[123] L. Greuter, D. Najer, A. V. Kuhlmann, S. R. Valentin, A.
Ludwig, A. D. Wieck, S. Starosielec, and R. J. Warburton,
“Epitaxial lift-off for solid-state cavity quantum electrody-
namics,” Journal of Applied Physics 118, 075705 (2015).

xiii
Bibliography

[124] T. Fink, A. Schade, S. Höfling, C. Schneider, and A.


İmamoğlu, “Signatures of a dissipative phase transition
in photon correlation measurements,” Nature Physics 14,
365–369 (2018).
[125] N. Syassen, D. M. Bauer, M. Lettner, T. Volz, D. Dietze, J. J.
Garcı́a-Ripoll, J. I. Cirac, G. Rempe, and S. Dürr, “Strong
Dissipation Inhibits Losses and Induces Correlations in Cold
Molecular Gases,” Science 320, 1329–1331 (2008).
[126] S. Dürr, J. J. Garcı́a-Ripoll, N. Syassen, D. M. Bauer, M.
Lettner, J. I. Cirac, and G. Rempe, “Lieb-Liniger model of a
dissipation-induced Tonks-Girardeau gas,” Physical Review
A 79, 023614 (2009).
[127] S. Diehl, A. Tomadin, A. Micheli, R. Fazio, and P. Zoller, “Dy-
namical Phase Transitions and Instabilities in Open Atomic
Many-Body Systems,” Physical Review Letters 105, 015702
(2010).
[128] H. J. Carmichael, Statistical Methods in Quantum Optics 1:
Master Equations and Fokker-Planck Equations (Springer-
Verlag, 1999).
[129] H. M. Gibbs, S. L. McCall, and T. N. C. Venkatesan, “Dif-
ferential Gain and Bistability Using a Sodium-Filled Fabry-
Perot Interferometer,” Physical Review Letters 36, 1135–
1138 (1976).
[130] A. Dorsel, J. D. McCullen, P. Meystre, E. Vignes, and H.
Walther, “Optical Bistability and Mirror Confinement In-
duced by Radiation Pressure,” Physical Review Letters 51,
1550–1553 (1983).
[131] L. Collot, V. Lefèvre-Seguin, M. Brune, J. M. Raimond, and
S. Haroche, “Very High- Q Whispering-Gallery Mode Reso-
nances Observed on Fused Silica Microspheres,” Europhysics
Letters 23, 327 (1993).
[132] A. Baas, J. P. Karr, H. Eleuch, and E. Giacobino, “Optical
bistability in semiconductor microcavities,” Physical Review
A 69, 023809 (2004).

xiv
Bibliography

[133] V. R. Almeida and M. Lipson, “Optical bistability on a


silicon chip,” Optics Letters 29, 2387–2389 (2004).
[134] M. Notomi, A. Shinya, S. Mitsugi, G. Kira, E. Kuramochi,
and T. Tanabe, “Optical bistable switching action of Si high-
Q photonic-crystal nanocavities,” Optics Express 13, 2678–
2687 (2005).
[135] P.-F. Braun, B. Urbaszek, T. Amand, X. Marie, O. Krebs, B.
Eble, A. Lemaitre, and P. Voisin, “Bistability of the nuclear
polarization created through optical pumping in In1−x Gax As
quantum dots,” Physical Review B 74, 245306 (2006).
[136] G. A. Wurtz, R. Pollard, and A. V. Zayats, “Optical Bista-
bility in Nonlinear Surface-Plasmon Polaritonic Crystals,”
Physical Review Letters 97, 057402 (2006).
[137] T. C. H. Liew, A. V. Kavokin, and I. A. Shelykh, “Opti-
cal Circuits Based on Polariton Neurons in Semiconductor
Microcavities,” Physical Review Letters 101, 016402 (2008).
[138] A. Amo, T. C. H. Liew, C. Adrados, R. Houdré, E. Giacobino,
A. V. Kavokin, and A. Bramati, “Exciton-polariton spin
switches,” Nature Photonics 4, 361–366 (2010).
[139] V. M. Menon, L. I. Deych, and A. A. Lisyansky, “Towards
polaritonic logic circuits,” Nature Photonics 4, 345 (2010).
[140] F. Marsault, H. S. Nguyen, D. Tanese, A. Lemaître, E. Ga-
lopin, I. Sagnes, A. Amo, and J. Bloch, “Realization of an
all optical exciton-polariton router,” Applied Physics Letters
107, 201115 (2015).
[141] E. P. Gross, “Structure of a quantized vortex in boson sys-
tems,” Il Nuovo Cimento 20, 454–477 (1961).
[142] L. P. Pitaevskii, “Vortex lines in an imperfect Bose gas,”
Soviet Physics JETP 13, 451–454 (1961), [Zh. Eksp. Teor.
Fiz. 40, 646 (1961)].
[143] R. Bonifacio and L. A. Lugiato, “Photon Statistics and Spec-
trum of Transmitted Light in Optical Bistability,” Physical
Review Letters 40, 1023–1027 (1978).

xv
Bibliography

[144] P. D. Drummond and D. F. Walls, “Quantum theory of


optical bistability. I. Nonlinear polarisability model,” Journal
of Physics A: Mathematical and General 13, 725–741 (1980).
[145] K. Vogel and H. Risken, “Quasiprobability distributions in
dispersive optical bistability,” Physical Review A 39, 4675–
4683 (1989).
[146] K. V. Kheruntsyan, “Wigner function for a driven anhar-
monic oscillator,” Journal of Optics B: Quantum and Semi-
classical Optics 1, 225–233 (1999).
[147] T. Schneider, G. Srinivasan, and C. P. Enz, “Phase Tran-
sitions and Soft Modes,” Physical Review A 5, 1528–1536
(1972).
[148] G. Venkataraman, “Soft modes and structural phase transi-
tions,” Bulletin of Materials Science 1, 129–170 (1979).
[149] H. Risken, C. Savage, F. Haake, and D. F. Walls, “Quantum
tunneling in dispersive optical bistability,” Physical Review
A 35, 1729–1739 (1987).
[150] K. Vogel and H. Risken, “Quantum-tunneling rates and
stationary solutions in dispersive optical bistability,” Physical
Review A 38, 2409–2422 (1988).
[151] M. Vojta, “Quantum phase transitions,” Reports on Progress
in Physics 66, 2069 (2003).
[152] D. Bitko, T. F. Rosenbaum, and G. Aeppli, “Quantum Crit-
ical Behavior for a Model Magnet,” Physical Review Letters
77, 940–943 (1996).
[153] P. Coleman, “Theories of non-Fermi liquid behavior in heavy
fermions,” Physica B: Condensed Matter 259-261, 353–358
(1999).
[154] H. von Löhneysen, “Non-Fermi-liquid behaviour in the heavy-
fermion system CeCu6-x Aux ,” Journal of Physics: Condensed
Matter 8, 9689 (1996).
[155] S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar,
“Continuous quantum phase transitions,” Review of Modern
Physics 69, 315–333 (1997).

xvi
Bibliography

[156] S. V. Kravchenko, W. E. Mason, G. E. Bowker, J. E.


Furneaux, V. M. Pudalov, and M. D’Iorio, “Scaling of an
anomalous metal-insulator transition in a two-dimensional
system in silicon at B = 0,” Physical Review B 51,
7038–7045 (1995).
[157] S. Sachdev, “Quantum Criticality: Competing Ground States
in Low Dimensions,” Science 288, 475–480 (2000).
[158] M. P. A. Fisher, P. B. Weichman, G. Grinstein, and D. S.
Fisher, “Boson localization and the superfluid-insulator tran-
sition,” Physical Review B 40, 546–570 (1989).
[159] S. Sachdev, Quantum Phase Transitions, 2nd ed. (Cambridge
University Press, 2011).
[160] E. M. Kessler, G. Giedke, A. İmamoğlu, S. F. Yelin, M. D.
Lukin, and J. I. Cirac, “Dissipative phase transition in a
central spin system,” Physical Review A 86, 012116 (2012).
[161] F. Storme, Dissipative phase transitions in open quantum
lattice systems, PhD thesis (Université Paris Diderot (Paris
7), Sorbonne Paris Cité, Nov. 2017).
[162] R. Roy, R. Short, J. Durnin, and L. Mandel, “First-Passage-
Time Distributions under the Influence of Quantum Fluctu-
ations in a Laser,” Physical Review Letters 45, 1486–1490
(1980).
[163] F. T. Hioe and S. Singh, “Correlations, transients, bistability,
and phase-transition analogy in two-mode lasers,” Physical
Review A 24, 2050–2074 (1981).
[164] P. Lett, W. Christian, S. Singh, and L. Mandel, “Macroscopic
Quantum Fluctuations and First-Order Phase Transition in
a Laser,” Physical Review Letters 47, 1892–1895 (1981).
[165] W. Casteels, R. Fazio, and C. Ciuti, “Critical dynamical prop-
erties of a first-order dissipative phase transition,” Physical
Review A 95, 012128 (2017).
[166] J. M. Fink, A. Dombi, A. Vukics, A. Wallraff, and P.
Domokos, “Observation of the Photon-Blockade Breakdown
Phase Transition,” Physical Review X 7, 011012 (2017).

xvii
Bibliography

[167] C. Carr, R. Ritter, C. G. Wade, C. S. Adams, and K. J.


Weatherill, “Nonequilibrium Phase Transition in a Dilute
Rydberg Ensemble,” Physical Review Letters 111, 113901
(2013).
[168] I. Carusotto, Linear and nonlinear optics in Bose fields: light
waves in dielectric structures, matter waves in optical lattices,
PhD thesis (Scuola Normale Superiore – Pisa, 2000).
[169] D. Chen, M. White, C. Borries, and B. DeMarco, “Quantum
Quench of an Atomic Mott Insulator,” Physical Review
Letters 106, 235304 (2011).
[170] J. Dziarmaga, “Dynamics of a quantum phase transition
and relaxation to a steady state,” Advances in Physics 59,
1063–1189 (2010).
[171] F. Letscher, O. Thomas, T. Niederprüm, M. Fleischhauer,
and H. Ott, “Bistability Versus Metastability in Driven Dissi-
pative Rydberg Gases,” Physical Review X 7, 021020 (2017).
[172] D. Renker, “Geiger-mode avalanche photodiodes, history,
properties and problems,” Nuclear Instruments and Methods
in Physics Research Section A: Accelerators, Spectrometers,
Detectors and Associated Equipment 567, 48–56 (2006).
[173] G. H. Goldsztein, F. Broner, and S. H. Strogatz, “Dynami-
cal Hysteresis without Static Hysteresis: Scaling Laws and
Asymptotic Expansions,” SIAM Journal on Applied Mathe-
matics 57, 1163–1187 (1997).
[174] T. W. B. Kibble, “Topology of cosmic domains and strings,”
Journal of Physics A: Mathematical and General 9, 1387
(1976).
[175] W. H. Zurek, “Cosmological experiments in superfluid he-
lium?” Nature 317, 505 (1985).
[176] W. H. Zurek, U. Dorner, and P. Zoller, “Dynamics of a
quantum phase transition,” Physical Review Letters 95,
105701 (2005).
[177] W. Casteels, F. Storme, A. Le Boité, and C. Ciuti, “Power
laws in the dynamic hysteresis of quantum nonlinear photonic
resonators,” Physical Review A 93, 033824 (2016).

xviii
Bibliography

[178] C. Pfleiderer, “Why first order quantum phase transitions


are interesting,” Journal of Physics: Condensed Matter 17,
S987 (2005).
[179] P. G. Savvidis, J. J. Baumberg, R. M. Stevenson, M. S. Skol-
nick, D. M. Whittaker, and J. S. Roberts, “Angle-Resonant
Stimulated Polariton Amplifier,” Physical Review Letters
84, 1547–1550 (2000).
[180] J. J. Baumberg, P. G. Savvidis, R. M. Stevenson, A. I. Tar-
takovskii, M. S. Skolnick, D. M. Whittaker, and J. S. Roberts,
“Parametric oscillation in a vertical microcavity: A polariton
condensate or micro-optical parametric oscillation,” Physical
Review B 62, R16247–R16250 (2000).
[181] A. Zamora, L. M. Sieberer, K. Dunnett, S. Diehl, and M. H.
Szymańska, “Tuning across Universalities with a Driven
Open Condensate,” Physical Review X 7, 041006 (2017).
[182] G. Roumpos, M. Lohse, W. H. Nitsche, J. Keeling, M. H.
Szymańska, et al., “Power-law decay of the spatial correlation
function in exciton-polariton condensates,” Proceedings of
the National Academy of Sciences 109, 6467–6472 (2012).
[183] G. Dagvadorj, J. M. Fellows, S. Matyja śkiewicz, F. M.
Marchetti, I. Carusotto, and M. H. Szyma ńska, “Nonequi-
librium Phase Transition in a Two-Dimensional Driven Open
Quantum System,” Physical Review X 5, 041028 (2015).
[184] N. Šibalić, C. G. Wade, C. S. Adams, K. J. Weatherill, and
T. Pohl, “Driven-dissipative many-body systems with mixed
power-law interactions: Bistabilities and temperature-driven
nonequilibrium phase transitions,” Physical Review A 94,
011401 (2016).
[185] A. Le Boité, G. Orso, and C. Ciuti, “Steady-State Phases
and Tunneling-Induced Instabilities in the Driven Dissipative
Bose-Hubbard Model,” Physical Review Letters 110, 233601
(2013).

xix
Bibliography

[186] R. M. Wilson, K. W. Mahmud, A. Hu, A. V. Gorshkov,


M. Hafezi, and M. Foss-Feig, “Collective phases of strongly
interacting cavity photons,” Physical Review A 94, 033801
(2016).
[187] J. Jin, D. Rossini, M. Leib, M. J. Hartmann, and R. Fazio,
“Steady-state phase diagram of a driven QED-cavity ar-
ray with cross-Kerr nonlinearities,” Physical Review A 90,
023827 (2014).
[188] J. J. Mendoza-Arenas, S. R. Clark, S. Felicetti, G.
Romero, E. Solano, D. G. Angelakis, and D. Jaksch,
“Beyond mean-field bistability in driven-dissipative lattices:
Bunching-antibunching transition and quantum simulation,”
Physical Review A 93, 023821 (2016).
[189] M. Foss-Feig, P. Niroula, J. T. Young, M. Hafezi, A. V.
Gorshkov, R. M. Wilson, and M. F. Maghrebi, “Emergent
equilibrium in many-body optical bistability,” Physical Re-
view A 95, 043826 (2017).
[190] M. Biondi, G. Blatter, H. E. Türeci, and S. Schmidt,
“Nonequilibrium gas-liquid transition in the driven-
dissipative photonic lattice,” Physical Review A 96, 043809
(2017).
[191] F. Vicentini, F. Minganti, R. Rota, G. Orso, and C. Ciuti,
“Critical slowing down in driven-dissipative Bose-Hubbard
lattices,” Physical Review A 97, 013853 (2018).
[192] J. Chun Tat Ngai, Measurement of short-time photon corre-
lations using frequency conversion at picosecond timescale,
Master thesis (ETH Zürich, 2017).
[193] S. Christopoulos, G. B. H. von Högersthal, A. J. D. Grundy,
P. G. Lagoudakis, A. V. Kavokin, et al., “Room-Temperature
Polariton Lasing in Semiconductor Microcavities,” Physical
Review Letters 98, 126405 (2007).
[194] G. Grosso, G. Nardin, F. Morier-Genoud, Y. Léger, and B.
Deveaud-Plédran, “Soliton Instabilities and Vortex Street
Formation in a Polariton Quantum Fluid,” Physical Review
Letters 107, 245301 (2011).

xx
Bibliography

[195] A. Amo, S. Pigeon, D. Sanvitto, V. G. Sala, R. Hivet, et


al., “Polariton superfluids reveal quantum hydrodynamic
solitons,” Science 332, 1167–1170 (2011).
[196] R. Hivet, H. Flayac, D. D. Solnyshkov, D. Tanese, T. Boulier,
et al., “Half-solitons in a polariton quantum fluid behave like
magnetic monopoles,” Nature Physics 8, 724 (2012).
[197] A. Kavokin, G. Malpuech, and M. Glazov, “Optical Spin
Hall Effect,” Physical Review Letters 95, 136601 (2005).
[198] C. Leyder, M. Romanelli, J. P. Karr, E. Giacobino, T. C. H.
Liew, M. M. Glazov, A. V. Kavokin, G. Malpuech, and A.
Bramati, “Observation of the optical spin hall effect,” Nature
Physics 3, 628 (2007).
[199] T. Jacqmin, I. Carusotto, I. Sagnes, M. Abbarchi, D. D.
Solnyshkov, G. Malpuech, E. Galopin, A. Lemaı̂tre, J. Bloch,
and A. Amo, “Direct Observation of Dirac Cones and a
Flatband in a Honeycomb Lattice for Polaritons,” Physical
Review Letters 112, 116402 (2014).
[200] Á. Cuevas, B. Silva, J. C. L. Carreño, M. de Giorgi, C. S.
Muñoz, et al., “Entangling one polariton with a photon:
effect of interactions on a single-polariton quantum state,”
(2017), arXiv:1609.01244v2.
[201] M. Ueda, M. Kuwata, N. Nagasawa, T. Urakami, Y.
Takiguchi, and Y. Tsuchiya, “Picosecond time-resolved pho-
toelectric correlation measurement with a photon-counting
streak camera,” Optics Communications 65, 315–318 (1988).
[202] J. Wiersig, C. Gies, F. Jahnke, M. Aßmann, T. Berstermann,
et al., “Direct observation of correlations between individual
photon emission events of a microcavity laser,” Nature 460,
245 (2009).
[203] M. Aßmann, F. Veit, J.-S. Tempel, T. Berstermann, H. Stolz,
M. van der Poel, J. M. Hvam, and M. Bayer, “Measuring the
dynamics of second-order photon correlation functions inside
a pulse with picosecond time resolution,” Optics Express 18,
20229–20241 (2010).

xxi
Bibliography

[204] J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S.


Pershan, “Interactions between Light Waves in a Nonlinear
Dielectric,” Physical Review 127, 1918–1939 (1962).
[205] H. Suchowski, B. D. Bruner, A. Arie, and Y. Silberberg,
“Broadband Nonlinear Frequency Conversion,” Optics and
Photonics News 21, 36–41 (2010).
[206] O. Kuzucu, F. N. C. Wong, S. Kurimura, and S. Tovstonog,
“Time-resolved single-photon detection by femtosecond up-
conversion,” Optics Letters 33, 2257–2259 (2008).
[207] M. T. Rakher, L. Ma, O. Slattery, X. Tang, and K. Srinivasan,
“Quantum transduction of telecommunications-band single
photons from a quantum dot by frequency upconversion,”
Nature Photonics 4, 786 (2010).
[208] M. T. Rakher, L. Ma, M. Davanço, O. Slattery, X. Tang, and
K. Srinivasan, “Simultaneous Wavelength Translation and
Amplitude Modulation of Single Photons from a Quantum
Dot,” Physical Review Letters 107, 083602 (2011).
[209] L. Ma, M. T. Rakher, M. J. Stevens, O. Slattery, K. Srini-
vasan, and X. Tang, “Temporal correlation of photons fol-
lowing frequency up-conversion,” Optics Express 19, 10501–
10510 (2011).
[210] G. Muñoz-Matutano, A. Wood, M. Johnson, X. V. Asen-
sio, B. Baragiola, et al., “Quantum-correlated photons from
semiconductor cavity polaritons,” (2017), arXiv:1712.05551.
[211] S. Klembt, P. Stepanov, T. Klein, A. Minguzzi, and M.
Richard, “Thermal Decoherence of a Nonequilibrium Polari-
ton Fluid,” Physical Review Letters 120, 035301 (2018).
[212] A. Reinhard, T. Volz, M. Winger, A. Badolato, K. J. Hen-
nessy, E. L. Hu, and A. Imamoglu, “Strongly correlated
photons on a chip,” Nature Photonics 6, 93 (2011).
[213] S. Smolka, W. Wuester, F. Haupt, S. Faelt, W. Wegscheider,
and A. İmamoğlu, “Cavity quantum electrodynamics with
many-body states of a two-dimensional electron gas,” Science
346, 332–335 (2014).

xxii
Bibliography

[214] S. Ravets, P. Knüppel, S. Faelt, O. Cotlet, M. Kroner, W.


Wegscheider, and A. İmamoğlu, “Polaron Polaritons in the
Integer and Fractional Quantum Hall Regimes,” Physical
Review Letters 120, 057401 (2018).
[215] S. Ravets, P. Knüppel, S. Faelt, M. Kroner, W. Wegscheider,
and A. İmamoğlu, “Enhancing photon-photon interactions
using fractional quantum hall states,” in preparation, 2018.
[216] I. Rosenberg, D. Liran, Y. Mazuz-Harpaz, K. West,
L. Pfeiffer, and R. Rapaport, “Strongly interacting
dipolar-polaritons,” (2017), arXiv:1802.01123v1.
[217] P. Cristofolini, G. Christmann, S. I. Tsintzos, G. Deligeor-
gis, G. Konstantinidis, Z. Hatzopoulos, P. G. Savvidis, and
J. J. Baumberg, “Coupling Quantum Tunneling with Cavity
Photons,” Science 336, 704–707 (2012).
[218] H.-T. Lim, E. Togan, S. Fält, W. Wegscheider, and A.
İmamoğlu, in preparation, 2018.
[219] I. Carusotto, D. Gerace, H. E. Türeci, S. De Liberato, C.
Ciuti, and A. İmamoğlu, “Fermionized Photons in an Array
of Driven Dissipative Nonlinear Cavities,” Physical Review
Letters 103, 033601 (2009).
[220] M. J. Hartmann, “Polariton Crystallization in Driven Arrays
of Lossy Nonlinear Resonators,” Physical Review Letters
104, 113601 (2010).
[221] D. G. Angelakis, M. F. Santos, and S. Bose, “Photon-
blockade-induced Mott transitions and XY spin models
in coupled cavity arrays,” Physical Review A 76, 031805
(2007).
[222] M. J. Hartmann, F. G. S. L. Brandao, and M. B. Plenio,
“Strongly interacting polaritons in coupled arrays of cavities,”
Nature Physics 2, 849–855 (2006).
[223] C. Vaneph, A. Morvan, G. Aiello, M. Féchant, M. Aprili,
J. Gabelli, and J. Estève, “Observation of the anomalous
photon blockade effect in the microwave domain,” (2018),
arXiv:1801.04227.

xxiii
Bibliography

[224] T. C. H. Liew and V. Savona, “Single Photons from Coupled


Quantum Modes,” Physical Review Letters 104, 183601
(2010).
[225] M. Bamba, A. İmamoğlu, I. Carusotto, and C. Ciuti, “Origin
of strong photon antibunching in weakly nonlinear photonic
molecules,” Physical Review A 83, 021802 (2011).
[226] A. F. Adiyatullin, M. D. Anderson, H. Flayac, M. T. Portella-
Oberli, F. Jabeen, C. Ouellet-Plamondon, G. C. Sallen, and
B. Deveaud, “Periodic squeezing in a polariton Josephson
junction,” Nature Communications 8, 1329 (2017).
[227] M.-A. Lemonde, N. Didier, and A. A. Clerk, “Antibunching
and unconventional photon blockade with Gaussian squeezed
states,” Physical Review A 90, 063824 (2014).
[228] H. Flayac and V. Savona, “Unconventional photon blockade,”
Physical Review A 96, 053810 (2017).
[229] M. van Regemortel, S. Ravets, A. İmamoğlu, I. Carusotto,
and M. Wouters, “Engineering Gaussian states of light from
a planar microcavity,” (2017), arXiv:1712.08012.
[230] R. O. Umucalılar and I. Carusotto, “Artificial gauge field for
photons in coupled cavity arrays,” Physical Review A 84,
043804 (2011).
[231] D. Gerace, H. E. Türeci, A. İmamoğlu, V. Giovannetti, and
R. Fazio, “The quantum-optical Josephson interferometer,”
Nature Physics 5, 281 (2009).
[232] R. O. Umucalılar and I. Carusotto, “Fractional Quantum
Hall States of Photons in an Array of Dissipative Coupled
Cavities,” Physical Review Letters 108, 206809 (2012).
[233] S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. Büchler,
and P. Zoller, “Quantum states and phases in driven open
quantum systems with cold atoms,” Nature Physics 4, 878–
883 (2008).
[234] M. Kiffner and M. J. Hartmann, “Dissipation-induced Tonks-
Girardeau gas of polaritons,” Physical Review A 81, 021806
(2010).

xxiv
Bibliography

[235] J. Cho, D. G. Angelakis, and S. Bose, “Fractional Quantum


Hall State in Coupled Cavities,” Physical Review Letters
101, 246809 (2008).
[236] T. Fink and H. Bluhm, “Distinguishing Quantum and
Classical Baths via Correlation Measurements,” (2014),
arXiv:1402.0235.
[237] T. Fink and H. Bluhm, “Noise Spectroscopy Using Cor-
relations of Single-Shot Qubit Readout,” Physical Review
Letters 110, 010403 (2013).
[238] B. Jungbluth, S. Nyga, E. Pawlowski, T. Fink, and J. Wuep-
pen, “Efficient frequency conversion of pulsed microchip and
fiber laser radiation in PPSLT,” Proceedings of SPIE 7912,
79120K (2011).

xxv
List of Figures

11 Historical publication data . . . . . . . . . . . . . . 2

21 InAs/GaAs band structure . . . . . . . . . . . . . . 8


22 Excitons in Inx Ga1-x As quantum wells . . . . . . . . 10
23 Electric field and spectrum of a DBR . . . . . . . . . 14
24 Electric field and spectrum of a microcavity . . . . . 16
25 Exciton-polaritons . . . . . . . . . . . . . . . . . . . 18
26 Micropillar cavity . . . . . . . . . . . . . . . . . . . . 25
27 Photon statistics of different states . . . . . . . . . . 26
28 Polariton blockade . . . . . . . . . . . . . . . . . . . 29
29 g (2) (0) detuning and power-dependence . . . . . . . 30

31 Fiber cavity setup . . . . . . . . . . . . . . . . . . . 33


32 Fiber fabrication setup . . . . . . . . . . . . . . . . . 35
33 Dimple profile on fiber facet . . . . . . . . . . . . . . 37
34 Dimple profile measurement . . . . . . . . . . . . . . 37
35 Dimple profile analysis . . . . . . . . . . . . . . . . . 38
36 Dimple AFM measurement . . . . . . . . . . . . . . 40
37 Electric field and spectrum of a fiber cavity . . . . . 41
38 FDTD simulations . . . . . . . . . . . . . . . . . . . 43
39 Numerical aperture . . . . . . . . . . . . . . . . . . . 45
310 Beam waist measurement . . . . . . . . . . . . . . . 46
311 Cavity spectroscopy . . . . . . . . . . . . . . . . . . 47
312 Transverse modes . . . . . . . . . . . . . . . . . . . . 49

xxvii
List of Figures

313 Cavity noise spectrum . . . . . . . . . . . . . . . . . 50


314 Cavity transmission spectra . . . . . . . . . . . . . . 51

41 Principle of bistability . . . . . . . . . . . . . . . . . 59
42 Nonlinear power dependence . . . . . . . . . . . . . 61
43 Liouvillian eigenspectrum . . . . . . . . . . . . . . . 63
44 Continuous phase transition . . . . . . . . . . . . . . 65
45 Photon bunching and Liouvillian eigenspectrum . . . 68
46 Polariton intensity distribution . . . . . . . . . . . . 70
47 Time traces and photon correlations . . . . . . . . . 71
48 Liouvillian eigenspectrum with dephasing . . . . . . 73
49 Polariton correlation function . . . . . . . . . . . . . 75
410 Photon bunching and timescale . . . . . . . . . . . . 76
411 Bunching decay around the critical point . . . . . . 77
412 Bunching scaling . . . . . . . . . . . . . . . . . . . . 78
413 Critical point and exponent scaling . . . . . . . . . . 79
414 Comparison first- and second-order coherence . . . . 80

51 Sum frequency generation . . . . . . . . . . . . . . . 87


52 Schematic of upconversion-based g (2) setup . . . . . 88
53 Photon correlations of an EOM-modulated CW laser 91
54 Resolution of upconversion setup . . . . . . . . . . . 92
55 Biexponential polariton decay . . . . . . . . . . . . . 94
56 Same-time correlations . . . . . . . . . . . . . . . . . 96
57 g (2) (t, t) detuning dependence . . . . . . . . . . . . . 97
58 Sub-Poissonian photon statistics . . . . . . . . . . . 99

61 Unconventional blockade . . . . . . . . . . . . . . . . 106

A1 Sample structure . . . . . . . . . . . . . . . . . . . . 110


A2 Exciton emission spectrum . . . . . . . . . . . . . . 110

xxviii
List of Tables

31 Overview of dimple parameters . . . . . . . . . . . . 39

51 Photon detection parameters . . . . . . . . . . . . . 85

xxix
Acknowledgments

Ataç İmamoğlu, I am deeply grateful for the opportunity to follow


my passion for physics with you. I truly enjoyed my time in the
Quantum Photonics Group and am thankful you for your support,
guidance, discussions, and permanent willingness to help when I
felt stuck. Thank you very much Jacqueline Bloch for reading
my manuscript on such short notice as well as your comments and
the insightful discussions we had at my defense. I am indebted to
Dario Gerace for welcoming me in Pavia in the beginning of my
PhD and teaching me his FDTD skills. Martin Kroner, thank
you very much for sharing your enormous expertise in lab work
and your feedback. You always had an open ear for me when I felt
frustrated and helped me find back my motivation. In addition,
thanks for your thorough proofreading of my thesis. Thanks to
Katharina Rodharth and Manuela Weber-Semler for keeping
the group running smoothly. For the samples I have been working
on my gratitude goes to Javier Miguel-Sánchez, Stefan Fält,
as well as the Würzburg team Anne Schade, Sven Höfling, and
Christian Schneider. During the last months of my PhD I had
the pleasure of working with Aymeric Delteil – thank you for
the fruitful collaboration and Franco-German friendship, the cakes
and our linguistic adventures, as well as your comments on my
manuscript. Among my fellow PhD students, a special thanks goes
to Yves Delley. I learned a great deal from you, thanks for being
a great office mate, running partner, friend, and for luring me into

xxxi
Acknowledgments

Python.
My gratitude goes to all past and present QPG members for
the nice atmosphere in the lab as well as during special seminars,
coffee breaks, numerous kicker games, the trip to Kandersteg, snow
shoe hiking, SOLA, and many more. I’d like to thank in particular
Andreas Reinhard for handing over his nice setup to me, as
well as the G8 crew Emre Togan and Jan Klärs, and Meinrad
Sidler with whom I shared all of my PhD, many enjoyable con-
ferences, and one infamous Arosa meeting in particular. Patrick
Knüppel, thanks for taking over Yves’ role as a great office mate.
Gian-Marco Schnüringer and Olivier Faist, thank you for do-
ing your theses with me. I hope you enjoyed the time as much
as I did working with you. Sylvain Ravets, thanks for creating
a fun atmosphere and good humor. Thank you very much Wolf
Wüster, Florian Haupt, and Adrian Maier for welcoming some-
one from the ’far north’ of Germany among you and the many beers,
Schnitzels, and Cordon Bleus we shared. Michelle Suppiger and
Andreas Wyss as well as Sarah Smolka and Stephan Smolka
(and Tom!), thank you for your friendship which made my time in
Zurich so enjoyable.
Meiner Familie gilt mein Dank für ihre Hilfe bei allem was ich tue,
selbst wenn nicht ganz klar ist was das überhaupt ist. Ich bin sehr
froh, dass ich immer auf Euch zählen kann. Vielen Dank Franziska
für Dein Verständnis, Deine Unterstützung und dafür, dass Du mich
ankerst und glücklich machst!

xxxii
Curriculum Vitae

Personal details
Name Thomas Fink
Date of Birth December 4, 1989
Place of Birth Mönchengladbach, Germany
Citizenship German
E-mail finkt@phys.ethz.ch

Education
2013 – 2018 Doctoral student, Institute of Quantum Elec-
tronics, ETH Zürich
2010 – 2012 Master of Science in Physics, RWTH Aachen
2008 – 2010 Bachelor of Science in Physics, RWTH Aachen

List of publications
(1) T. Fink, A. Schade, S. Höfling, C. Schneider, and A. İmamoğlu,
“Signatures of a dissipative phase transition in photon correla-
tion measurements,” Nature Physics 14, 365–369 (2018)
(2) S. Bogdanović, M. S. Z. Liddy, S. B. van Dam, L. C. Coenen,
T. Fink, M. Lončar, and R. Hanson, “Robust nano-fabrication

xxxiii
Curriculum Vitae

of an integrated platform for spin control in a tunable micro-


cavity,” APL Photonics 2, 126101 (2017)
(3) S. Bogdanović, S. B. van Dam, C. Bonato, L. C. Coenen,
A.-M. J. Zwerver, B. Hensen, M. S. Z. Liddy, T. Fink, A. Reis-
erer, M. Lončar, and R. Hanson, “Design and low-temperature
characterization of a tunable microcavity for diamond-based
quantum networks,” Applied Physics Letters 110, 171103
(2017)
(4) M. Sidler, P. Back, O. Cotlet, A. Srivastava, T. Fink, M. Kro-
ner, E. Demler, and A. İmamoğlu, “Fermi polaron-polaritons
in charge-tunable atomically thin semiconductors,” Nature
Physics 13, 255–261 (2016)
(5) T. Fink and H. Bluhm, “Distinguishing Quantum and Classical
Baths via Correlation Measurements,” (2014), arXiv:1402.
0235
(6) T. Fink and H. Bluhm, “Noise Spectroscopy Using Correlations
of Single-Shot Qubit Readout,” Physical Review Letters 110,
010403 (2013)
(7) B. Jungbluth, S. Nyga, E. Pawlowski, T. Fink, and J. Wuep-
pen, “Efficient frequency conversion of pulsed microchip and
fiber laser radiation in PPSLT,” Proceedings of SPIE 7912,
79120K (2011)

xxxiv

Das könnte Ihnen auch gefallen