You are on page 1of 23

Cavity-Based Single-Photon Sources

Axel Kuhn∗ and Daniel Ljunggren†


University of Oxford, Clarendon Laboratory, Parks Road, OX1 3PU, United Kingdom
(Dated: January 05, 2010)
We introduce the basic concepts and characteristic properties of single-photon emitters based on
resonator effects in optical cavities, and provide a review of the most prominent implementations.
First we discuss the elementary principles of cavity quantum electrodynamics, which determine how
single quantum systems couple to the quantised field modes of optical resonators, and then show how
to exploit these principles in order to generate single photons on demand. Furthermore, we elucidate
how to study the properties of the photons, namely their singleness and their spectral characteristics.
In particular, it is explained how to use a Hanbury-Brown & Twiss interferometer to determine
whether the photons arrive one-by-one, and how to probe their mutual indistinguishability with two-
photon interference experiments of the Hong-Ou-Mandel type. Finally, we present a classification
scheme applicable to most cavity-based photon emitters realised to date. This scheme allows us to
assess the pros and cons of the most significant approaches – in particular with regard to photonic
quantum computing and quantum communication, which are among the most promising applications
to date.

Contents Acknowledgments 20

I. Introduction 1 References 20

II. Atom-photon interaction in resonators 2


I. INTRODUCTION
A. Field quantisation in cavity QED 2
B. Two-level atom 3
C. Three-level atom 4 The ultimate control of individual quantum systems is
D. Cavity-coupling regimes 4 a key element in emerging topics of modern physics, such
as quantum information processing. Here, the achievable
III. Single-photon emission 4 degree of control relies on the ability to arbitrarily ad-
A. Cavity-enhanced spontaneous emission 5 dress, manipulate, and couple individual physical enti-
B. Steady-state evolution in the bad-cavity ties, like single atoms or single photons, which are used
regime 6 as information carriers. For instance, the interfacing of
C. Dynamic evolution in the strong-coupling atoms and photons, the storage and retrieval of single
regime 7 photons, and the mapping of quantum states between
D. Arbitrary photon shaping 8 distant entities, all constitute essential building blocks
of future quantum communication networks and quan-
IV. Single-photon characterisation 8 tum information processors [1]. Ideally, such systems are
A. Statistics: photons one-by-one 9 composed of individual nodes acting as quantum gates
B. Indistinguishability: the photon identity 10 or memories, with optical links in between that allow
Two-photon interference 11 for the entanglement or teleportation of their quantum
Step-by-step analysis 11 states, or for optical quantum information processing us-
Identical and distinguishable photons 12 ing light travelling between the nodes [2]. With individ-
Frequency jitter 13 ual photons acting as messengers in these networks, wide
Time jitter 13 efforts are being undertaken that focus on the production
and characterisation of single photons. Applications rely-
V. Routes to single-photon sources 14 ing on the availability of single photons include quantum
A. Neutral single atoms 15 cryptography, optical quantum computing, light-matter
B. Trapped single ions 17 entanglement, and atom-photon state mapping, which
C. Artificial atoms in solid-state 18 all have successfully been demonstrated in the past few
years.
VI. Future directions 20 In this article, we restrict ourselves to review the con-
cepts, characteristic properties, and major implementa-
tions of state-of-the-art single-photon sources based on
single emitters in cavities. These have all the potential
∗ Corresponding author: axel.kuhn@physics.ox.ac.uk to meet the requirements of optical quantum computing
† Current affiliation: Dept. of Applied Physics, KTH - Royal Insti- and quantum networking schemes, namely
tute of Technology, 106 91 Stockholm, Sweden. Email: daniellj@
kth.se • deterministic single-photon emission with unit effi-
2

ciency, between different cavities has been predicted [11, 12], and
demonstrated [13].
• directed emission into a single mode of the radia- In the next section, we thoroughly discuss the elemen-
tion field, tary principles of cavity-QED that apply whenever a sin-
• indistinguishable photons with immaculate tempo- gle quantum system is coupled to a quantised field mode
ral and spatial coherence, and of a cavity. We then show in section 3 how to exploit
these effects to generate single photons on demand in
• reversible quantum state mapping and entangle- the strong coupling regime and the bad cavity limit, us-
ment between atoms and photons. ing either an adiabatic driving technique or a sudden ex-
citation of the emitter. In section 4, we elucidate how to
We emphasise that no single-photon source has been re- study the properties of the photons: their singleness and
alised so far which meets all of the above criteria. For their temporal and spectral characteristics. In particular,
instance, photon sources based on spontaneous paramet- we focus on the first and second order coherence proper-
ric down conversion are inherently probabilistic, and can- ties of the light. The section explains how to analyse and
not reach the desired degree of determinism, even when measure the photon counting statistics using a Hanbury-
assisted by an heralding signal. For other sources that ex- Brown & Twiss interferometer to determine whether the
ploit the spontaneous emission of single photons into free- photons arrive one-by-one, and how to obtain informa-
space, from emitters like trapped atoms, ions, molecules, tion on their mutual bandwidth and/or temporal jitter
or quantum defects in solid-state systems, the obvious from two-photon interference experiments of the Hong-
lack of directionality can be overcome by collecting pho- Ou-Mandel type. In section 5, we introduce a classifica-
tons from a large solid angle. Results on entangling tion scheme for tracking the performance and properties
atoms or ions with emitted photons have been obtained of cavity-based single-photon emitters. This scheme al-
[3, 4], as well as teleportation of quantum states [5]. How- lows us to assess the advantages and disadvantages of
ever, as the restriction of the solid angle diminishes the the most significant approaches, in particular with re-
efficiency, these sources are rendered probabilistic as well. gard to photonic quantum computing, such as linear op-
Instead, with the single emitter being placed inside an tical quantum computing (LOQC) [2] and quantum com-
optical cavity, the Purcell effect can be utilised, which, munication [14], the two most promising applications to
together with coherent control of the state dynamics, al- date. We also discuss the most prominent experimen-
lows for deterministic generation of single photons. As tal achievements and examine the different approaches
discovered by Purcell [6], the spatial mode density inside for obtaining single photons from cavities using either
a cavity is altered substantially, such that the sponta- atoms, ions or quantum dots as photon emitters.
neous emission rate can be either enhanced (f > 1) or
inhibited (f < 1) by the Purcell factor
II. ATOM-PHOTON INTERACTION IN
3Qλ3 RESONATORS
f= ,
4π 2 V
depending on the cavity’s mode volume, V , and quality In this section we discuss how a single quantum system,
factor, Q. More importantly, the probability of sponta- which shows discrete energy levels like an individual atom
neous emission into the cavity is given by β = f /(f + 1). or ion, couples to a quantised mode of the radiation field.
The field modes in the cavity are quantised, and with a We first introduce the relevant features of cavity-QED
suitable spacing between cavity modes, the emitter only and the Jaynes-Cummings model [15, 16], after which
couples to a single field mode. Provided the mode volume we focus on three-level atoms with two dipole transitions
of the cavity is sufficiently small, the emitter and cavity driven by two radiation fields. One of the fields is from
couple so strongly that the emission probability into the a laser, the other is the cavity field strongly coupled to
cavity, β, is significantly larger than the probability for the atom[115]. Furthermore, we shortly explain how the
spontaneous emission into free space. Hence, in prin- behaviour of a coupled-atom system depends on the most
ciple, a deterministic photon emission into a single field relevant cavity parameters, such as the cavity’s mode vol-
mode can be achieved with an efficiency close to unity for ume and its finesse.
very large f . These effects have first been observed by
Carmichael et al. [7] and De Martini et al. [8]. Moreover,
with the coherence properties uniquely determined by A. Field quantisation in cavity QED
the parameters of the cavity and the driving process, one
should be able to obtain indistinguishable photons from Let us consider the coupling of a discrete quantum
different cavities. Note also that state mapping and en- system to a Fabry-Perot cavity with mirror separation
tanglement between atomic spin and photon polarisation l and reflectivity R. The cavity has a free spectral
has recently been demonstrated in cavity-based single- range ∆ω√ F SR = 2π × c/(2l), and its finesse is defined
photon emitters [9, 10]. Additionally the reversibility of as F = π R/(1 − R). In the vicinity of a resonance,
the photon generation process, and quantum networking the transmission profile is Lorentzian with a linewidth
3

(FWHM) of 2κ = ∆ωF SR /F, which is twice the decay a b c


rate, κ, of the cavity field. Curved mirrors are often used
to restrict the cavity eigenmodes to geometrically stable ΔC |x,n-1〉 |x,1〉 |φ 2+〉
Hermite-Gaussian or Laguerre-Gaussian modes. We as- |g,2〉 |φ 2−〉
ωC Ωn,eff
|φ 1+〉
sume the atom is coupled to just one of these modes, |x,0〉
characterised by its mode function ψC (r) and its reso- ωC ΔC
|g,1〉 |φ 1−〉
nance frequency ωC . In general, the state vector is a su-
ωC
perposition of photon-number states, |ni, and for n pho- |g,0〉
|g,n〉 |g,0〉
tons in the mode the total energy is h̄ωC (n + 21 ). The
equidistant energy spacing imposes an analogue treat-
ment of the cavity as an harmonic oscillator. Conse- FIG. 1: (a) A two-level atom with ground state |gi and ex-
quently, creation and annihilation operators for a pho- cited state |xi coupled to a cavity containing n photons. In
ton, ↠and â, are used to express the Hamiltonian of the the dressed-level scheme of the combined atom-cavity system
with the atom outside (b) or inside (c) the cavity, the state
cavity,
  doublets are either split by ∆C or by the effective Rabi fre-
1 quency, Ωn,eff , respectively.
HC = h̄ωC ↠â + . (1)
2
Note that this Hamiltonian does not take account of any
a b c
losses. In a real cavity, all photon number states de-

atomic bare state energy


cay until thermal equilibrium with the environment is Δ |x,n−1〉 |φ n+〉

dressed state energy


reached. In the optical domain, the latter corresponds to |x,n−1〉
|φ n0〉 Ωsplit
the vacuum state, |0i, with no photons remaining in the Ω
Δ
Ωeff
cavity. |e,n−1〉; |g,n〉
|φ n−〉
|e,n-1〉 |g,n〉

B. Two-level atom
FIG. 2: (a) A three-level atom driven by a classical laser
We now analyse how the cavity field interacts with a field of Rabi frequency Ω, coupled to a cavity containing n
two-level atom with ground state |gi and excited state photons. (b) Dressed-level scheme of the combined system
|xi of energies h̄ωg and h̄ωx , respectively, and transition without coupling, and (c) for an atom interacting
p with laser
dipole moment µxg . The Hamiltonian of the atom reads and cavity. The triplet is split by Ωsplit = 4ng02 + Ω2 + ∆2 .
In the limit of a large detuning ∆, the Raman transition |e, n−
HA = h̄ωg |gihg| + h̄ωx |xihx|. (2) 1i ↔ |g, ni is driven at the effective Rabi frequency Ωeff =
1
2
(Ωsplit − |∆|) ≈ (4ng02 + Ω2 )/|4∆|.
The coupling to the field mode of the cavity is expressed
by the atom-cavity coupling constant,
q
g(r) = g0 ψC (r), with g0 = (µ2xg ωC )/(2h̄0 V ), (3)
where ∆C = ωx − ωg − ωC is the detuning between atom
where V is the mode volume of the cavity. As the atom and cavity. The level splitting p between the two corre-
is barely moving during the interaction, we can safely sponding eigenstates, Ωn,eff = 4ng02 + ∆2C , is the ef-
disregard its external degrees of freedom. Furthermore fective Rabi frequency at which the population oscillates
we assume maximum coupling, i.e. ψC (ratom ) = 1, so between states |g, ni and |x, n − 1i. This means that the
that one obtains g(r) = g0 . In a closed system, any cavity field stimulates the emission of an excited atom
change of the atom’s internal state must be reflected by into the cavity, thus de-exciting the atom and increasing
a corresponding change of the cavity’s photon number, the photon number by one. Subsequently, the atom is
n. Hence the interaction Hamiltonian of the atom-cavity re-excited by absorbing a photon from the cavity field,
system reads and so forth. In particular, an excited atom and a cavity
containing no photon are sufficient to startpthe oscilla-
Hint = −h̄g0 |xihg|â + ↠|gihx| .
 
(4)
tion between |x, 0i and |g, 1i at frequency 4g02 + ∆2C .
For a given excitation number n, only the product states This phenomenon is known as vacuum-Rabi oscillation,
|g, ni and |x, n − 1i are coupled. If the cavity is resonant and for ∆C = 0, the resulting oscillation frequency 2g0
with the atomic transition, the√population oscillates with is called vacuum-Rabi frequency.
the Rabi frequency ΩC = 2g0 n between these states.
To summarise, the atom-cavity interaction splits the
The eigenfrequencies of the total Hamiltonian, H =
photon number states into doublets of non-degenerate
HC + HA + Hint , can be found easily. In the rotating
dressed states, which are named after Jaynes and Cum-
wave approximation, they read
    mings [15, 16]. Only the ground state |g, 0i is not coupled
1 1 to any other state and is therefore not subject to any en-
q
± 2 2
ωn = ωC n + + ∆C ± 4ng0 + ∆C , (5)
2 2 ergy shift or splitting.
4

C. Three-level atom D. Cavity-coupling regimes

Now we consider an atom with a Λ−type three-level In the preceding sections, we have been considering
scheme providing transition frequencies ωxe = ωx − ωe the interaction Hamiltonian and the associated eigenval-
and ωxg = ωx − ωg as depicted in fig. 2. The |ei ↔ |xi ues and dressed eigenstates that one obtains whenever
transition is driven by a classical light field of frequency a two- or three-level quantum system is coupled to a
ωL with Rabi frequency Ω, and a cavity mode with fre- cavity. We have been neglecting the transverse polari-
quency ωC couples to the |gi ↔ |xi transition. If we sation decay rate, γ⊥ , of the quantum system, and also
define the respective detunings as ∆L = ωxe − ωL and the field-decay rate of the cavity, κ, has not been taken
∆C = ωxg − ωC , and assume that the driving laser and into account[116]. It is evident that both relaxation rates
the cavity only couple to their respective transitions, the result in a damping of a possible vacuum-Rabi oscillation
behaviour of the atom-cavity system is described by the between states |x, 0i and |g, 1i, and therefore determine
interaction Hamiltonian how the atom is coupled to the cavity. Two extreme cases
are worth mentioning here, namely the
Hint = h̄[ ∆L |eihe| + ∆C |gihg| − Ω2 (|xihe| + |eihx|)
−g0 (|xihg|a + a† |gihx|)]. • Strong-coupling regime, with g0  {κ, γ⊥ }. In
(6) this case, the damping rates have a negligible effect
Given an arbitrary excitation number n, this Hamilto- on the time evolution of the coupled atom-cavity
nian couples only the three states |e, n − 1i, |x, n − 1i, system, and vacuum-Rabi oscillations can occur.
|g, ni. For this triplet and a Raman-resonant interaction For a three-level atom exposed to driving laser and
with ∆L = ∆C ≡ ∆, the eigenfrequencies of the coupled cavity, the condition sometimes reduces to g0  κ,
system read provided one can eliminate the excited atomic state
  from the description.
0 1
ωn = ωC n + and (7) • Bad-cavity regime, with κ  g02 /κ  γ⊥ . This
2
    results in a strong damping and quasi-stationary
1 1
q
ωn± = ωC n + + ∆ ± 4ng02 + Ω2 + ∆2 . quantum states of the coupled system (see section
2 2 III B).
The Jaynes-Cummings doublets of the two-level atom are Two properties of the cavity can be used to distinguish
now replaced by triplets, between these regimes: First√ the strength of the atom-
|φ0n i = cos Θ|e, n − 1i − sin Θ|g, ni, (8) cavity coupling, g0 ∝ 1/ V ∝ l−3/4 (dependant upon
the dimensions of the√ cavity mode, with the transverse
|φ+
n i = cos Φ sin Θ|e, n − 1i − sin Φ|x, n − 1i mode area πw02 ∝ l for near-planar cavities), and sec-
+ cos Φ cos Θ|g, ni, ond the finesse F of the resonator, which depends on the
− mirror reflectivity. The finesse corresponds to the average
|φn i = sin Φ sin Θ|e, n − 1i + cos Φ|x, n − 1i
+ sin Φ cos Θ|g, ni, number of round trips in the cavity before a photon gets
lost by transmission through one of the cavity mirrors,
where the mixing angles Θ and Φ are given by and can also be expressed as the ratio between free spec-
p tral range ∆ωF SR and cavity linewidth 2κ. Consequently
Ω 4ng02 + Ω2 a trade-off between g0 and κ is needed to keep one high
tan Θ = √ , tan Φ = p .
2g0 n 2
4ng0 + Ω2 + ∆2 − ∆ and the other low, as the latter increases with 1/(F × l).
(9) Fig. 3 shows the two coupling regimes as a function of
We note that the interaction with the light lifts the de- finesse and cavity length. For the strong-coupling regime
generacy of the three eigenstates as soon as the Rabi fre- that relies on a high value of g0 and therefore a short
quencies are non-zero. Furthermore, we emphasise that cavity of small mode volume, only a cavity of sufficiently
neither |φ0n i is subject to an energy shift, nor does the high finesse will keep κ small enough. This usually re-
excited atomic state contribute to it. It is therefore often quires a mirror reflectivity R ≥ 99.999%.
called a ‘dark state’ since it cannot decay by spontaneous
emission.
In the limit of vanishing Ω, the states |φ± III. SINGLE-PHOTON EMISSION
n i correspond
to the Jaynes-Cummings doublet and the third eigen-
state, |φ0n i, coincides with |e, n − 1i. Note that ωn0 is not In this section, we introduce several different ways
affected by Ω or g0 . Therefore transitions between the of producing single photons from a coupled atom-cavity
dark states |φ0n+1 i and |φ0n i are always in resonance with system. This encompasses cavity-enhanced spontaneous
the cavity. This holds, in particular, for the transition emission and Raman transitions stimulated by the vac-
from |φ01 i to |φ00 i ≡ |g, 0i since the n = 0 state does not uum field while driven by classical laser pulses. In par-
split (the corresponding states |e, −1i and |x, −1i do not ticular, we introduce a scheme for adiabatic coupling be-
exist). tween a single atom and an optical cavity, which is based
5

l regime: A three-level system is coupled to the cav-


ity and a driving laser. In the bad-cavity regime,
the cavity decay κ is stronger than g0 and Ω. There-
↑ weak coupling fore the system is overdamped to a quasi-stationary
bad cavity
operation ւ ր state at any time of the interaction. As the station-
↓ strong coupling
ary state of the coupled system depends on Ω(t),
low coupling factor the envelope function of the emitted photon can be
controlled to a large extend. However, with some
high coupling factor transient population placed in the excited state,
one cannot rule out spontaneous emission losses
into non-cavity modes and the efficiency is reduced.
high C
good cavity III. Dynamic evolution in the strong-coupling
regime: A three-level system is strongly coupled
strong coupling
to the cavity and a driving laser, with {Ω, g0 }  κ.
bad cavity operation Therefore the cavity decay may be neglected on the
time scale of the excitation, and the coupled system
undergoes stimulated Raman transitions between
the two atomic ground states and the associated
F zero- and one-photon states of the cavity. For in-
stance, the driving process can be implemented in
FIG. 3: Cavity coupling regimes as determined by cavity fi- form of an adiabatic passage (STIRAP process [17])
nesse, F, and cavity length, l. C is the cooperativity param- or a far-off resonant Raman process to avoid plac-
eter.
ing any transient population into the excited level,
thus reducing losses due to spontaneous emission
on a unitary evolution of the coupled atom-cavity system into free space. In that case, an efficiency for pho-
[17], and is therefore intrinsically reversible. This scheme ton generation close to unity can be reached. Once
either allows one to populate photon-number states on a photon is placed into the cavity, it gets emitted
demand in the cavity, or to emit single optical photons due to the finite cavity-decay rate κ.
into a well-defined mode of the radiation field outside the
We emphasise that these three scenarios represent ex-
cavity.
treme cases. Any real cavity-based single-photon emitter
For a photon emission from the cavity to take place, it
combines several of the above aspects, according to the
is evident that a finite value of κ is mandatory, otherwise
parameters of the coupled system and the specific charac-
any light would remain trapped in the cavity. Moreover,
teristics of the driving field. The following three sections
as κ is the decay rate of the cavity field, the associated
discuss the three different approaches in more detail.
minimum length of an emitted photon is in most cases
given by κ−1 . We also emphasise that γ⊥ plays a cru-
cial role in most experimental settings, since it accounts A. Cavity-enhanced spontaneous emission
for the spontaneous emission into non-cavity modes, and
therefore leads to a reduction of efficiency. The relation
As a starting point, we assume that some excitation
of the atom-cavity coupling constant and the Rabi fre-
process (e.g. a short π pulse with Ω  {g0 , κ, γ⊥ }, or
quency of the driving field to the two decay rates can
some internal relaxation cascade starting from an ener-
be used for marking the difference between three basic
getically higher state) has been implemented to drive the
classes of single-photon emission schemes from cavity-
quantum system suddenly into its excited state |x, 0i.
QED systems:
Hence we are free to treat the system as a simple excited
I. Cavity-enhanced spontaneous emission: A two-level quantum system coupled to an empty optical
sudden excitation process (e.g. a short π pulse with cavity. This particular situation is the text book exam-
Ω  {g0 , κ, γ⊥ }, or some internal relaxation cas- ple of cavity-QED that has been thoroughly analysed in
cade starting from an energetically higher state) the past. In fact, it has been proposed by Purcell [6] and
drives the quantum system into its excited state. demonstrated by Heinzen et al. [18] and Morin et al.
Subsequently, a photon gets spontaneously emit- [19] that the spontaneous emission properties of an atom
ted either into the cavity or into free space. Due to coupled to a cavity are significantly different from those
the Purcell effect, the emission probability into the in free space. For an analysis of the atom’s behaviour, it
cavity is strongly enhanced by the Purcell factor suffices to look at the evolution of n = 1 doublet under
f = 2C, where C is the cooperativity parameter of the influence of the atomic polarisation decay rate γ⊥
the cavity. and the cavity-field decay rate κ. As non-cavity spon-
taneous decay of the atom and photon emission through
II. Steady-state evolution in the bad-cavity one of the cavity mirrors both lead the system into state
6

|g, 0i, i.e. a dressed state outside the n = 1 doublet, we


can deal with these decay processes phenomenologically
by introducing non-hermitian terms into the interaction
Hamiltonian,
0
= −h̄g0 |xihg|â + ↠|gihx| − ih̄γ⊥ |xihx| − ih̄κ↠â.

Hint
(10)
Figure 4(a) shows the time evolution of the atom-cavity
system when κ > g0 . The strong damping of the cavity’s
one-photon state inhibits any vacuum-Rabi oscillation,
since the photon is emitted from the cavity before it can
be reabsorbed by the atom. Therefore the transient pop-
ulation in state |g, 1i is negligible if the atom is initially
in its excited state |xi. In this case, the adiabatic ap-
proximation ċg ≈ 0 can be applied, which gives
d g2
cx = −γ⊥ cx − 0 cx , (11)
dt κ
with the solution
g02
   
cx (t) = exp − γ⊥ + t . (12)
κ
It is straightforward to see that the ratio of the emission
rate into the cavity, g02 /κ, to the spontaneous emission
probability into free space becomes g02 /(κγ⊥ ) ≡ f , i.e.
the Purcell factor. It equals twice the one-atom coop-
erativity parameter, C, originally introduced in the con-
text of optical bistability [20]. Hence the photon-emission FIG. 4: Time evolution of the system and photon emis-
probability from the cavity reads PEmit = 2C/(2C + 1). sion rate Rph = 2κρgg for different coupling cases. In
Note that the atom radiates mainly into the cavity if the upper row, an excited two-level atom is coupled to
g02 /κ  γ⊥ . Together with κ  g0 , this condition con- the cavity and populations ρxx (solid line) and ρgg (dot-
stitutes the bad-cavity regime. ted line) of the product states |x, 0i and |g, 1i are shown.
The other extreme we are considering is the strong- In the lower row, a three-level atom-cavity system pre-
coupling regime, g0  (κ, γ⊥ ). In this case the atom- pared in |e, 0i is exposed to a pump pulse driving |ei − |xi
cavity system is subject to vacuum-Rabi oscillations be- while the cavity couples |xi and |gi. The population ρee
tween |x, 0i and |g, 1i, with both states decaying at the of the initial state is shown as a dashed line. The left
column displays the bad-cavity regime with (g0 , γ⊥ , κ) =
respective rates γ⊥ and κ. Figure 4(b) shows a situa-
2π×(15, 3, 20) MHz, and the right column the strong-coupling
tion where the atom-cavity coupling, g0 , saturates the regime with (g0 , γ⊥ , κ) = 2π × (15, 3, 2) MHz. The pump
|x, 0i ↔ |g, 1i transition. On average, the probabilities pulses read Ω(t) = g0 sin(πt/200 ns) in (c), and Ω(t) =
to find the system in either one of these two states are g0 × t/1 µs in (d). No transient population is found in ρxx
equal, and therefore the average ratio of the emission in the latter case. In all cases,R the overall photon-emission
probability into the cavity to the spontaneous emission probability is given by PEmit = Rph dt.
probability into free space is given by κ/γ⊥ . The vacuum-
Rabi oscillation in this particular case gives rise to a pro-
nounced amplitude modulation of the photons emitted to a cavity. We first follow Law et al. [21, 22] and con-
from the cavity, so that their coherence properties are sider the bad-cavity regime, κ  g02 /κ  γ⊥ , where the
not as favourable as in the other schemes discussed here. loss of excitation into unwanted modes of the radiation
We emphasise that the enhancement of the atom’s field is small. We assume that the atom’s |ei − |xi tran-
spontaneous emission into the cavity cannot be denoted sition is excited by a pump pulse while the atom emits
in terms of simple analytic expressions in intermediate a photon into the cavity by enhanced spontaneous emis-
coupling regimes. Nonetheless, the emission probability sion. In this regime, the cavity-field decay rate κ sets the
into the cavity will be enhanced as long as the cavity fastest time scale, while the spontaneous emission rate
mode is resonant with the atomic transition. into the cavity, g02 /κ, dominates the incoherent decay of
the polarisation from the excited atomic state. Assum-
ing that any decay leads to a loss from the three-level
B. Steady-state evolution in the bad-cavity regime system, the evolution of the wave vector is governed by
the non-Hermitian Hamiltonian
To take the effect of the excitation process into ac-
0
count, we now consider a Λ-type three-level atom coupled Hint = Hint − ih̄κ↠â − ih̄γ⊥ |xihx|, (13)
7

with Hint given in (6). To simplify the analysis, we con- The most promising approach to achieve that goal is to
sider only the vacuum state, |0i, and the one-photon implement an adiabatic passage in the optical domain be-
state, |1i, of the cavity. Hence the state vector can be tween these two states [23, 24]. In fact, adiabatic passage
written as methods have been used for coherent population transfer
in atoms or molecules for many years. For instance, if
|Ψ(t)i = ce (t)|e, 0i + cx (t)|x, 0i + cg (t)|g, 1i, (14) a Raman transition is driven by two distinct pulses of
variable amplitudes, effects like electromagnetically in-
where ce , cx and cg are complex amplitudes. The time duced transparency (EIT) [25, 26], slow light [27, 28],
evolution of the amplitudes is given by the Schrödinger and stimulated Raman scattering by adiabatic passage
d 0
equation, ih̄ dt |Ψi = Hint |Ψi, which yields (STIRAP) [17] are observed. All of these effects have
been demonstrated with classical light fields, and have
1
iċe = 2 Ω(t)cx the property in common that the system’s state vector,
iċx = 1
− iγ⊥ cx |Ψi, always coincides with a single eigenstate, e.g. |φ0n i,
2 Ω(t)ce + g0 cg (15)
of the time-dependent interaction Hamiltonian. In prin-
iċg = g0 cx − iκcg ,
ciple, the time evolution of the system is completely con-
with the initial condition ce (0) = 1, cx (0) = cg (0) = 0 trolled by the variation of this eigenstate. However, a
and Ω(0) = 0. An adiabatic solution of (15) is found as more detailed analysis [23, 29] reveals that the eigen-
long as Ω(t)  g02 /κ, as the decay is so fast that cx and states must change slowly with respect to the eigenfre-
cg are nearly time independent. This allows one to make quency differences. Adiabaticity is assured if the con-
the approximations ċx = 0 and ċg = 0, with the result dition |ωn± − ωn0 |  |hφ± d 0
n | dt |φn i| is met throughout the
interaction, and as long as the system does not decay via
α
 Z t  some other channel. In this context, the non-decaying
2 0 0
ce (t) ≈ exp − Ω (t )dt dark state, |φ0n i, is of enormous significance.
4 0 It follows that a three-level atom-cavity system, once
cx (t) ≈ − 2i αΩ(t)ce (t) (16) prepared in |φ0n i, should stay there, thus allowing one to
cg (t) ≈ − κi g0 cx (t), control the relative population of the contributing prod-
uct states, |e, n − 1i and |g, ni, by simply adjusting the
where α = 2/(2γ⊥ +2g02 /κ). Since photon emissions from pump Rabi frequency, Ω. To show this, let us start with
the cavity only occur if the system is in |g, 1i, the photon- a system initially prepared in state |e, n − 1i. As can
emission rate reads Rph (t) = 2κ|cg (t)|2 , thus yielding a be seen from √ Eq. (8), this state coincides with |φ0n i if the
photon-emission probability of condition 2g0 n  Ω is met in the beginning of the inter-
Z action. Once the system has been successfully prepared
in the dark state, the ratio between the populations of
PEmit = Rph (t)dt (17) the contributing states reads
g2 α g2 α
  Z 
α |he, n − 1|Ψi|2 4ng02
= 0 1 − exp − Ω2 (t)dt −→ 0 . = . (18)
κ 2 κ |hg, n|Ψi|2 Ω2
Note
R that the exponential in (17) vanishes if the area As proposed in [30], we now assume that an atom in
Ω(t)dt of the exciting pump pulse is large enough. In state |ei is placed in a cavity mode populated with n − 1
this limit, the photon-emission probability does not de- photons driving the |g, ni ↔ |x, √n − 1i transition with
pend on the shape and amplitude of the pump pulse. the effective Rabi frequency 2g0 n, such that the initial
With a suitable choice of g0 , α, and κ, high photon- state coincides with |φ0n i. The atom is then exposed to
emission probabilities can be reached [22]. a laser pulse coupling the |ei ↔ |xi transition with √a
slowly rising amplitude that finally leads to Ω  2g0 n.
The system evolves from |e, n − 1i to |g, ni, and the pho-
C. Dynamic evolution in the strong-coupling ton number increases by one. Without cavity decay, the
regime successive application of this method would allow one to
prepare arbitrary photon-number states [31]. To do so,
To study the effect of the exciting laser pulse in the the pump pulse must be turned off suddenly, or the sys-
strong-coupling regime, we again consider a Λ-type three- tem would adiabatically return to its initial state.
level atom coupled to a cavity. We assume that the We emphasise that the situation is the same if one
strong-coupling condition also applies to the Rabi fre- starts with an empty cavity, i.e. with the atom-cavity
quency of the driving field, i.e. {g0 , Ω}  {κ, γ⊥ }. In this system initially in state |e, 0i, application of this method
case, we can safely neglect the effect of the two damping transfers the coupled system into state |g, 1i via the dark
rates on the time scale of the excitation and look for a state |φ01 i, with exactly one photon emitted into the cav-
method to effectively stimulate a Raman transition be- ity mode. In this particular case, the excitation scheme
tween the two ground states, which also places a single constitutes vacuum-stimulated Raman scattering by adi-
photon into the cavity. abatic passage, known as V-STIRAP. In the case of a
8

finite cavity decay time κ−1 much longer than the inter- a b
action time, a photon is emitted from the cavity with a Pump
probability close to unity and with properties uniquely Data
defined by κ, after the system has been excited to |g, 1i.

Events [arb. units]


In contrast to this idealised scenario, Fig. 4(d) shows weak strong
coupling coupling
a more realistic situation where the photon is generated
and emitted from the cavity during the excitation pro-
cess. This is due to the cavity decay time being compara- c d
ble or shorter than the exciting laser pulse. Even in this
case, no secondary excitations or photon emissions take
place, as the system reaches the decoupled state |g, 0i
once the photon escapes. However, the photon-emission
probability is slightly reduced as the non-Hermitian con-
tribution of κ to the interaction Hamiltonian gives rise

Photon rate [arb. units]


to a small admixture of |x, 0i to the eigenstate |φ01 i. The
latter is therefore no pure dark state, as it is weakly e f
ideal pump
affected by spontaneous emission losses. Nonetheless,
Photon pulse
photon-emission probabilities far above 90% are always
attainable if V-STIRAP is applied.

Time [µs]

D. Arbitrary photon shaping FIG. 5: Shaping the photon. (a-d) show photon shapes
realised in several experiments together with the respective
driving laser pulses. To obtain this data, the histogram of
As we have seen from the previous analysis, the dy- the photon-detection time has been recorded using several
namic evolution of the atomic quantum states provides hundred single-photon emissions. The data shown in (a+b)
information of the photon emission rate, and thereby also is taken from [33], with neutral atoms falling through a high-
the photon’s waveform. This raises the question as to finesse cavity acting as photon emitters. The linear increase
what extent we can arbitrarily shape the photons in time in Rabi frequency is the same in both cases, and the differ-
ence in photon shape is caused by variations in the coupling
by controlling the driving field’s envelope. This is an im-
strength to the cavity. The data shown in (c+d) is taken
portant problem for applications such as quantum state from [34], with a single ion trapped between the cavity mir-
mapping, where the availability of photon wave pack- rors. It impressively shows that the photon shape depends
ets that are symmetric in space and time allow for a strongly on the driving laser pulse. (Reprinted by permis-
time-reversal of the emission process [11]. Custom pho- sion from Macmillan Publishers Ltd: Nature (M. Keller et al.
ton shaping is also of interest for generating approximate [34]), copyright 2004.) The last row (e+f ) shows the Rabi
Gaussian pulse shapes which are shown to maximise the frequency one needs to apply to achieve symmetric single or
tolerance against mode-mismatch in interference-based twin-peak photon pulses with an efficiency close to unity. This
quantum information processing schemes [32]. Employ- is a result from an analytic solution of the problem discussed
ing photons of soliton-shape for dispersion-free propa- in [35].
gation in optical fibres could also help boost quantum
communication protocols.
IV. SINGLE-PHOTON CHARACTERISATION
The problem of photon shaping have been addressed
traditionally by solving the Master equation of the atom-
photon system, which yields the time-dependent proba- In this section, we elucidate how the properties of pho-
bility amplitudes, and by consequence also the wave func- tons, namely their singleness, their temporal and their
tion of the photon emitted from the cavity [33, 34]. It has spectral characteristics can be studied. In particular, we
recently been shown [35] that this analysis can also be re- focus on the first- and second-order coherence properties
versed, giving a unambiguous analytic expression for the of the light intensity. The characterisation of a single-
time evolution of the driving field as a function of the de- photon source usually starts with an investigation of the
sired shape of the photon. This model is not only valid for photon statistics, where the observation of antibunching
V-STIRAP in the strong-coupling and bad-cavity regime, indicates that the source tends to emit single photons.
but it generally allows control of the coherence and pop- First, we explain how to measure and analyse the photon
ulation flow in any Raman process. Fig. 5 compares some counting statistics, and how the photon statistics are af-
of the results obtained. fected by several properties of the photon emitter, such as
9

low efficiency, blinking and intermittent operation, back- the time delay τ between photo detections. In terms of
ground noise, and a spurious multi-photon contribution. the intensity of the light field the autocorrelation func-
Second, we focus on the coherence and temporal prop- tion is given by
erties of the emitted photons, and show how to obtain
information on their mutual bandwidth and/or temporal hI1 (t)I2 (t + τ )i
g (2) (τ ) = , (19)
jitter from two-photon interference experiments. hI1 (t)ihI2 (t)i

where In (t) is the light power recorded by photodetector


A. Statistics: photons one-by-one n. Similarly, in terms of probabilities of photon counts for
two single-photon detectors, the autocorrelation function
The most crucial test for a single-photon emitter con- is given by the cross-correlation
sists of a statistical analysis of the emitted photon stream
to determine whether the photons arrive one-by-one. For P12 (τ )
g (2) (τ ) = , (20)
a deterministic source that emits single photons in re- P1 (t) P2 (t + τ )
sponse to some trigger pulse, a direct measurement of
the photon number statistics upon each trigger event where Px : x ∈ {1, 2, 12} is the probability of a single
would completely characterise the singleness. Ideally, or joint detection events, as given by the counting rates
one expects to find probabilities of P (n = 1) = 1 and from the two detectors.
P (n 6= 1) = 0. However, as the efficiency of most sources A few interesting cases are shown in Fig. 6 and dis-
is far below one, it is generally more suitable to verify the cussed in Ref. [37]. Classical light fulfils the Cauchy-
sub-Poissonian statistics of the radiation by measuring Schwarz inequality, leading to g (2) (0) ≥ g (2) (τ 6= 0),
the second-order intensity correlation function, g (2) (τ ), whereas light that violates this inequality must be de-
of the photon stream with a Hanbury-Brown & Twiss scribed by the laws of quantum physics. First experi-
setup [36]. ments demonstrating antibunching, g (2) (0) < g (2) (τ 6=
In simple terms, it is needed that the probability of 0), were performed with a beam of atoms [38]. Limi-
finding another photon is vanishingly small once a first tations imposed by number fluctuations [39] were later
photon has been detected. In principle, this could be ac- eliminated by using a single ion [40], atom [41, 42],
complished by an ideal photon counter that registers the molecule [43], quantum dot [44], or color centre [45, 46].
arrival times of all impinging photons. However, due to For light with g (2) (0) = 0, photon number states with
the limited dead time of real detectors and their inherent n > 1 are totally absent. Hence the value at τ = 0 indi-
incapability of distinguishing between multiple photon cates how the expectation value of the photon number is
events, a pair of detectors is normally used. Hanbury- changed upon a first detection event. Classical (thermal)
Brown & Twiss’ method works by splitting the incom- bunching, g (2) (0) > g (2) (τ 6= 0), has been observed in the
ing photon stream into two using a beamsplitter, after fluorescence of a large number of independently radiating
which a measurement of the cross-correlation using two atoms as early as 1956 [36], and has regained new inter-
detectors (D1 and D2 ) gives a signal proportional to the est in the context of cold-atom physics [47, 48]. Only
autocorrelation of the light. The principle is depicted in recently, one of has been observing a smooth transition
Fig. 6. In the limit of infinitely short detector dead-times, between these two regimes in a cavity-QED experiment
[49], with the photon statistics changing from antibunch-
a b ing to bunching as a function of the atom number in
D1 g (2) (τ ) the cavity. Note that the special case of a flat corre-
lation function with g (2) (τ ) ≡ 1 is found for coherent
single- 2
photons τ light, i.e. laser radiation, as its photon statistics is Pois-
sonian. By consequence, the photon statistics is called
1 sub-Poissonian if g (2) (0) < 1, and super-Poissonian when
trigger g (2) (τ ) g (2) (0) > 1.
D2
0 τ In view of the above, g (2) (0) = 1 implies no change
0
to the mean photon number after detection and corre-
FIG. 6: a) Basic principle of Hanbury-Brown & Twiss’ tech-
sponds to a completely uncorrelated Poissionian statis-
nique to obtain correlation statistics for single-photon charac- tics, g (2) (0) = 2 implies doubling the mean as for ther-
terisation. The autocorrelation function of the photon stream mal statistics, and g (2) (0) = 0 implies that nothing is left
is measured using two detectors behind a beamsplitter with to detect after a first detection, as for the ideal number
one signal delayed by τ . b) The correlation function can show state |1i. In order to quantify the statistics of the pho-
signs of either systematic correlations (solid line), random cor- ton stream emitted from a given source in terms of its
relations (dashed line), or no correlations (dash-dotted line), single-photoness, we are therefore generally interested in
where the latter is an indication of non-classical light being the autocorrelation value at zero-lag, i.e. when τ = 0,
measured.
P12
the full autocorrelation function is found as a function of g (2) (0) = . (21)
P1 P2
10

10 a 1.7 b
Some typical examples of g (2) (τ ) are plotted in Fig. 7. N=0.15 N=6.6
9 1.6
The graphs are based on experimental data for different 8
cases of continuous versus discrete (pulsed) sources, and 1.5
7
for free falling versus trapped atoms. It is seen that the 6 1.4
form of g (2) (τ ) essentially resembles that of Fig. 6b for 5 1.3

intensity correlation g(2)(τ)


non-classical states, showing clear signs of antibunching 4
1.2
as discussed above. Note, however, that traces 7a-d seem 3
2 1.1
to indicate super-Poissionian statistics, even after sub-
1 1.0
traction of the background. This is because the measured -4 -2 0 2 4 -4 -2 0 2 4
photon statistics is additionally affected by the statistical 8 c 10 d
distribution of atoms or ions that are available to emit a atoms 8 atoms
6 3.4 ms 10 ms
photon (or by the intermittent operation in the case of
6
quantum dots, so-called blinking), which adds systematic 4
correlations to the data. In contrast to these, trace 7e 4

depicts the photon statistics obtained with a single atom 2 2


trapped inside a cavity. Hence the atom number is not 0 0
subject to statistical variations and the emitted radiation -30 -20 -10 0 10 20 30 -30 -20 -10 0 10 20 30
is sub-Poissonian, as can be seen from the inset. Gener- delay τ [μs] delay τ [μs]
35
ally, a Poisson-distributed atom number adds a unit con-

correlations per atom


30 e 1.5
stant to g (2) (τ < τc ), while a thermal distribution adds single trapped atom
25 1.0
a constant 2, and a fixed atom number state |N ie adds 20
a constant 1 − 1/N . In addition, values are also scaled 0.5
15
by a factor that is inversely proportional to the average 10 0
-40 -20 0 20 40
atom number N̄ (or to the duty-cycle in case of blink- 5 delay τ [μs]
ing), which corresponds to an inclusion of dark events 0
and leads to large values of g (2) (τc < τ < τi ) for N̄ < 1, -400 -300 -200 -100 0 100 200 300 400
where τi is the atom-cavity interaction time (or the blink- delay τ [μs]
ing period). For large delays there are only random corre-
lations remaining, and thus g (2) (τ → ∞) = 1. The ideal FIG. 7: The intensity correlation function g (2) (τ ) measured
case of a single trapped emitter, |1ie , does not affect g (2) using a Hanbury-Brown & Twiss setup for different types of
for any delay, since N = N̄ = 1. With these effects ac- cavity-based photon emitters. (a+b) are recorded for a con-
counted for, e.g. by using a trigger pulse to exclude dark tinuously driven system [49], with a stream of atoms falling
through the cavity. Antibunching is visible for small average
events and/or conditioning the experiment on the actual
atom number, N̄ , and a pronounced thermal bunching peak
presence of a photon emitter in the cavity, the output is found for N̄  1. (c+d) show the photon statistics for
is normally consistent with the requirements of a single- a triggered source, with atoms falling randomly through the
photon source, namely to show both sub-Poisson statis- cavity [33, 50]. The Gaussian envelope is due to the limited
tics and anti-bunching. For further discussions around atom-cavity interaction time, and the comb of peaks reflects
this topic we refer to Ref. [50]. the stream of trigger pulses. With the average atom number
being small, the missing central peak signals antibunching.
Note, however, that the raw data shows no sub-Poissonianess,
as it is affected by the random arrival of atoms in the cav-
B. Indistinguishability: the photon identity
ity. (e) corresponds to the ideal case, with a single emitter
(atom) trapped inside a cavity [51]. Antibunching and sub-
The requirements on a single-photon source for LOQC Poissonianess are evident, and no envelope function is visible
and for the entanglement of two distant atoms go far in the correlation function.
beyond showing sub-Poissonian statistics and antibunch-
ing. The realisation of these proposals relies on indis-
tinguishability between the photons, so that even pho- coincides with the single-mode envelope of the photon
tons from different sources are identical with respect to wavepackets if all photons are identical. A more sophis-
their frequency, duration and shape. Therefore it is desir- ticated analysis is therefore mandatory.
able to investigate the spectral and temporal properties The standard way to determine the coherence time of a
of single photons emitted from a given source. A simple given light source is through measurement of the second-
correlation measurement between the trigger event and order interference using a Mach-Zehnder or Michelson
the detection time of the generated photon will reveal interferometer. This measurement can also be done with
the temporal envelope (shape) of the photonic ensemble single photons, so that each single photon follows both
[33, 34, 52], assuming sufficiently high detector timing paths of the interferometer and interferes with itself.
resolution. However, as this method is insensitive to the This method has been used to measure the coherence
spectral properties of the photons, the measured enve- time of solid-state based single-photon sources [53, 54].
lope generally represents temporal multimodes and only However, these experiments cannot reveal jitter in the
11

emission time, and they are also not the method of choice ity to detect the photons of each pair in different out-
for characterising photons of long coherence time, or put ports of the beamsplitter. As we shall see, the
length, due to the mere impossibility of scanning one joint detection probability is zero for identical photons.
interferometer arm by several hundred meters. With polarisation-entangled photon pairs emitted from
In the following, we shortly summarise how to charac- a down-conversion source, this effect has first been used
terise the very narrow frequency bandwidth of single pho- to test the violation of Bell’s inequality by joint photode-
tons created by cavity-based emitters using two-photon tections [59], and one year later for measuring the delay
interference, avoiding the above restrictions. between two photons with sub-picosecond precision [55].
Two-photon interference phenomena have also been em-
ployed to test the indistinguishability of independently
Two-photon interference generated single photons [53, 57].

The first experiment to investigate the mutual coher- δτ δτ


ence time in photon pairs was performed by Hong et al. 2 3
[55]. They were measuring the second-order interference
between two photons from a parametric down-conversion
source by superimposing the signal and the idler photon
on a symmetric beamsplitter. The coincidence rate of
photodetections at the two output ports of the beam- 1 4
splitter was measured in dependence of a relative arrival-
time delay between the two photons. Due to their bosonic τ
nature, indistinguishable photons coalescence and leave
the beamsplitter as a pair, so that no coincident photo FIG. 9: Two single-photon wave packets impinge with a rel-
detections are recorded in the two output ports, as de- ative delay of δτ on ports ‘1’ and ‘2’ of a beamsplitter. The
scribed in Fig. 8. If the photons differ slightly, e.g. be- pulse duration is longer than the time resolution of the detec-
tors monitoring ports ‘3’ and ‘4’ of the beamsplitter. There-
cause they arrive at different times, coincidences will be
fore, the time difference, τ , between two photon detections
present. Furthermore, the coincidence rate shows a min- must be taken into account in the analysis of the photon
imum as a function of the arrival-time delay. For indis- statistics.
tinguishable photons, the width of this dip equals the
photon coherence time, i.e. duration. Any distinguisha- As a starting point for the following discussion, we
bility between the interfering photons reduces the dip’s review the phenomenon in the number-state picture, i.e.
depth. However, this method is also problematic with we consider two photons that impinge on the entrance
long coherence times. ports labelled ‘1’ and ‘2’ of the beam splitter shown in
Fig. 9. This situation is described by the quantum state
a 1 b 1 c 1 d 1 |11 12 i = â†1 â†2 |00i. The effect of the beamsplitter on the
field is usually described by the unitary transformations,
2 2 2 2
+ + â†1 = √1 (â

+ â†4 ), â†2 = √1 (â

− â†4 ),
2 3 2 3
+ + + +
(22)
â†3 = √1 (â
2 1

+ â†2 ), â†4 = √1 (â
2 1

− â†2 ).

FIG. 8: Two impinging photons lead to four possible pho- If we use these relations to express the initial state, we
ton distributions at the beam-splitter output. In the first two immediately see that this leads to an entangled state with
cases (a) and (b) the photons would be found together. In either both photons in one or the other output,
the remaining two cases (c) and (d) the photons would leave
the beamsplitter through different ports. Since the quantum |11 12 i = â†1 â†2 |00i → 21 (â†3 + â†4 )(â†3 − â†4 )|00i
states of the cases (c) and (d) show different signs, they in- (23)
→ √1 (|23 04 i − |03 24 i).
terfere destructively. 2

In simple physical terms, this can be also understood as


As a remedy, in case of a photon duration that is very destructive interference of the two possible paths leading
long compared to the detector time resolution, the de- to photons in different output ports, as illustrated by
tection time of a photon can also be measured within Fig. 8.
the single-photon envelope by recording the time delay
between photodetections [56–58]. This time-resolved in-
vestigation of two-photon interference effects is a probate Step-by-step analysis
way to determine the coherence time.
We assume that two independent photon streams im- For a quantitative analysis, we now consider free-
pinge on a symmetric beamsplitter such that two pho- running photons in the space-time domain. No bound-
tons are superimposed, and we ask for the probabil- ary conditions restrict the photons to specific frequency
12

modes, and we are free to define operators â†k and âk ce −1 ]


i d en [δt
inc τ)
that create or annihilate photons in arbitrarily chosen
a
spatio-temporal modes (labelled by the index k). In gen-
n co (τ,δ Δ=0
to P 34
eral, these modes can be defined by a one-dimensional ho
o -p ility
b
Tw oba
spatio-temporal mode function
ζk (t) = k (t)e−iφk (t) . (24) pr
0.15
0.1

δt]
Beside an arbitrary phase evolution φk (t), the mode func-
2
0.05

τ[
tion incorporates an amplitude envelope k (t), which we


assume to be normalised so that dt|k (t)|2 = 1. To 0
R
0

ela
analyse the effect of the beamsplitter, we consider the
Dete -2
ctio

nd
n-t
occupied input modes described by ζ1 (t) and ζ2 (t), and
ime 0
diffe -2

oto
we then consider one photon detection after the other.
re nce 2

Ph
That is we ask for the probability to detect a photon in
output port ‘4’ at time t0 + τ conditioned on a photon τ [δt
]
ce −1 ]
detection in output port ‘3’ at time t0 . The probability
i d en [δt
to find a photon at time t0 in port ‘3’ simply reads b inc τ) Δ=3
n co (τ,δ π [δt −1
oto P 34
P3 (t0 ) = hΨin |â†3 â3 |Ψin i|t0
]
- ph lity
(25) o bi
= 1
2 (|1 (t0 )|
2
+ |2 (t0 )|2 ), Tw oba
pr
where |Ψin i = |11 12 i is the initial state. This expression 0.4
includes no interference, since the initial product state 0.3

δt]
is composed of single-photon states that have no relative 0.2 2
0.1

τ[
phase. However, observing a photon at t0 reduces the 0


remaining quantum state to |Ψcond i, which can also be 0
Dete-2

ela
seen as a state conditioned on the detection of a photon ctio

nd
in port ‘3’ at time t0 . It is obtained by applying the n- time 0
diffe -2

oto
annihilation operator â3 to |Ψin i, which yields r ence 2

Ph
τ [δt
|Ψcond i ∝ ζ2 (t0 )|11 02 i + ζ1 (t0 )|01 12 i. (26) ]

This corresponds to a single photon impinging on the


FIG. 10: Joint probability, P34 , for the detection of photon
beamsplitter along two paths, with amplitude and phase
pairs as a function of the relative delay between the two im-
relations given by ζ2 (t0 ) and ζ1 (t0 ). The result therefore pinging photon wave packets, δτ , and the difference between
is classical interference, with both paths in phase at t0 , photon-detection times, τ for the frequency differences ∆ = 0
and subsequent phase evolution according to ζ1 (t) and and ∆ = 3π δt−1 . All times differences are given in units of
ζ2 (t). The probabilities to detect the second photon in the photon-pulse duration δt.
output ports ‘3’ or ‘4’ at time t0 + τ therefore oscillates,
so that the joint photon-detection probability for a first
detection in output port ‘3’ at time t0 and a subsequent ζ1 (t) ≡ ζ2 (t), irrespective of the specific shape of the
photon detection in output port ‘4’ at time t0 + τ reads mode function. The situation is not that simple if the
photons differ from one another. Still no joint detec-
P34 (t0 , τ ) = 41 |ζ1 (t0 + τ )ζ2 (t0 ) − ζ2 (t0 + τ )ζ1 (t0 )|2 . (27) tions occur simultaneously, i.e. P34 (τ = 0) ≡ 0, but the
It is quite remarkable that the step-by-step analysis of photons can be detected in different ports of the beam
two-photon interference reveals its equivalence to classi- splitter as soon as τ 6= 0.
cal interference, with the only difference being the con- To understand how slight differences between photons
ditioning on the detection of a first photon. Note also affect the joint detection probability, we consider Gaus-
that the sequence of photo detections is irrelevant, and sian pulses of unit-time duration δt (half 1/e width) that
for negative detection-time delay, τ , which corresponds impinge on the beamsplitter with a relative delay time
to a reverse order of photon detections, the same result δτ and a slight carrier-frequency difference ∆ = ω2 − ω1 .
is obtained. For these, evaluation of Eq. (27) yields a joint photon-
detection probability of
1
Identical and distinguishable photons P34 (τ, δτ, ∆) = √
2 π
(cosh(2τ δτ ) − cos(τ ∆))
2
(28)
−τ 2
×e−δτ ,
From Eq. (27), we can see that the joint detection
probability is always zero for identical photons, i.e. for which is averaged over t0 . This joint detection proba-
13

∞ zero for τ = 0, and the bandwidth of the photons is


0.5 δω
δω = 3 π / δt only affecting its width. Hence the dip-width seems a
δω = π / δt good measure for the mutual coherence time (or recipro-
Joint detection probability

0.4 cal bandwidth) of the photons.


2 T1
It should be noted here that the above is only appli-
2 T2
0.3 cable for photons long compared to the detector time-
resolution. If solid state systems are considered, the light
0.2
pulses are usually far too short to measure any detection-
time difference. In this latter case, the coincidence prob-
ability is only a function of the relative photon delay, δτ ,
0.1 and is given by

0.0 2 cos2 ϕ
 
1 2 2

-3 -2 -1 0 1 2 3 hP34 (δτ )iτ = 1− √ e−δτ /δt . (30)


2 4 + δt2 δω 2
Detection-time difference τ [ δt ]
Any frequency jitter, δω, now leads to a decreased depth
of the Gaussian-shaped dip, while the width of this dip is
FIG. 11: Average joint detection probability, P̄34 , as a func- not affected and always identical to the photon duration.
tion of the detection-time difference, τ , for simultaneously
impinging photons, δτ = 0, of identical polarisation. The
photons are subject to a frequency jitter of width δω. Time jitter

Even perfectly identical photons could be subject to


bility is displayed in Fig. 10 as a function of the photon-
a jitter in their mutual arrival time at the beam split-
delay, δτ , and the detection-time delay, τ , for two dif-
ter. This gives rise to a relative delay between the two
ferent values of the frequency difference, ∆. The case
mode functions, and hence to a non-zero joint detection
∆ = 0 is depicted in Fig. 10a. As expected, no coinci-
probability for τ > 0. Assuming a Gaussian arrival-time
dences are found if the photons arrive simultaneously,
distribution of width ∆τ , the average P34 shows again a
i.e. for δτ = 0, which is well explained by the model
central dip reaching zero, which now has the width
of two interfering indistinguishable photons. Moreover,
even if the two photons are delayed with respect to each p
δt2 + δt4 /∆τ 2 (31)
other, we see that simultaneous photon detections (with
τ = 0) never occur. Figure 10b shows a situation where As before, time-resolved two-photon interference can be
the two impinging photon wave packets have a frequency used to determine the variation in the emission time of
difference of ∆ = 3π, in units of δt−1 . Simultaneous de- the photons. However, since the shapes of the joint detec-
tections (with τ = 0) still do not occur, although the two tion probabilities for frequency and emission-time jitter
photons are now distinguishable. However, for simulta- are identical, it is in general not possible to distinguish
neously arriving photon wave packets (with δτ = 0), the between the two broadening mechanisms. Nonetheless,
coincidence probability now oscillates as a function of τ one can determine the maximum values for both, as well
with the frequency difference ∆. This beat note always as all pairs of frequency and emission-time jitter combi-
starts at τ = 0, due to the in-phase starting condition nations matching the data. This is discussed in [58].
imposed by the detection of the first photon. In this section, we have shown that the Hanbury-
Brown & Twiss and Hong-Ou-Mandel techniques con-
stitute two excellent tools to characterise single photons.
Frequency jitter In the latter technique, two photons are superimposed
on a beamsplitter and the joint detection probability is
With the in-phase starting condition not depending on in the two output ports of the beam splitter measured as
the mutual frequency difference ∆ between the photons, a function of the arrival plus detection-time difference of
the coincidence probability at τ = 0 remains zero even the photons. For identical photons, the joint detection
in case of an inhomogeneous broadening of frequencies. probability is expected to be zero. Variations of the spa-
To illustrate this, we assume a Gaussian frequency dis- tiotemporal modes of the photons lead to joint photode-
tribution of bandwidth δω (half width at 1/e maximum), tections except for zero detection-time difference. There-
which yields an average photon detection probability of fore the joint detection probability shows a pronounced
2
dip. From the width of this dip, one can estimate the
e−τ  2 2
 maximum emission-time jitter and the minimum coher-
P̄34 (τ, δτ = 0, δω) = √ 1 − e−τ δω /4 . (29) ence time of the photons. Figure 12 illustrates how to
2 π
use the Hong-Ou-Mandel technique to characterise a sin-
As shown in Fig. 11, a 2/δω-wide dip is found in the gle photon emitter. The first of a pair of successively
two-photon detection probability. The dip goes down to emitted photons is send along an optical delay line to
14

less at random times due to the spontaneous nature of


the generation process. To some extent this can be re-
sisted by post-selection, but it is still not a very efficient
method. In addition, the frequency bandwidth originat-
ing from the phase-matching conditions of parametric
down-conversion is in general very broad on the scale
of atomic transitions, and thus unsuitable for many ap-
plications like, e.g., atom-based qubit memories.
During the last decade also other photon generation
schemes have been demonstrated, such as a single-photon
turnstile device based on the Coulomb blockade mecha-
nism in a quantum dot [65], the fluorescence of a single
molecule [66, 67], or a single colour centre (Nitrogen va-
cancy) in diamond [45, 68], or the photon emission of
a single quantum dot into free space [44, 69, 70]. All
FIG. 12: Typical setup for performing a two-photon inter- these new schemes emit photons upon an external trigger
ference experiment to characterise a single photon emitter
event. However, the photons are spontaneously emitted
[57, 60]. The first of a pair of successively emitted photons is
sent along an optical delay line, e.g. a fibre of several 100 m into many modes of the radiation field, i.e. into all direc-
length, to arrive simultaneously with the second photon at a tions, and they usually show a very broad energy distri-
beam splitter. Coincidences in the output ports of this beam- bution. For the same reason, the emission process cannot
splitter are recorded. Rotating the relative polarisation of the be described by a Hamiltonian evolution. Hence the pro-
interfering photons is a way to obtain a reference signal, as cess is hardly reversible, and does not allow for coherent
light of perpendicular polarisation yields no interference. quantum state mapping from source to photon and back.
This does not hamper one from using these photons for
quantum cryptography and communication, but it rep-
arrive simultaneously with the second photon at a beam resents a major obstacle to all applications in quantum
splitter. Hence the joint detection probability is primar- computing or quantum networking. As discussed in the
ily affected by the mutual difference between successive previous sections, most of these limitations can be over-
photons. Long-term drifts of spectral properties will have come by cavity-enhanced emission techniques into well-
a negligible effect on the signal. defined modes of the radiation field. The remainder of
this section will focus on these.
In section III, we have outlined three different ways of
V. ROUTES TO SINGLE-PHOTON SOURCES producing photons from a cavity-based system. Fig. 13
illustrates these three schemes, along with the basic fea-
In this section we review different types of single- tures of the underlying physical methods and the prop-
photon sources that have been implemented to date erties of the photons obtained. All cavity-based photon
where the Purcell effect plays a major role in enhanc- sources realised so far nicely fit into one of these three
ing the emission from single-atoms, single-ions, or artifi- categories, which one therefore can use to classify a large
cial atoms (quantum dots) placed in optical cavities. In variety of existing single-photon sources. Apart from the
order to compare each of the implementations in terms applied scheme, further relevant parameters are the cav-
of their ability to deliver well-prepared single-photons, ity’s mode volume V , its finesse F, coupling and damp-
and their ease of operability, we introduce a classifica- ing rates {g0 , κ, γ⊥ }, cooperativity C = g02 /(2κγ⊥ ), and
tion scheme based on the three classes of cavity-QED all the benchmark parameters characterising the photons
emission schemes discussed in section 3. Along with a produced, namely their coherence time τc , the trigger-
discussion of the pros and cons of the different sources pulse rate RT and the efficiency η, which is the proba-
we also provide a table of benchmarks for some selected bility for a photon emission upon a single trigger pulse
realisations. We furthermore give a short view on the sent to the system. With some of the sources showing a
many types of cavities which advance in technology has blinking or intermittent behaviour of operation, both RT
made available in recent years. and η must refer to the active mode of operation to allow
Hitherto, undoubtedly the most popular technique to for any comparison of different implementations. For the
generate single photons is parametric down-conversion same reason, it is rather inconvenient to simply ask for
in non-linear crystals, which can mainly be explained by g (2) (0). Instead, the anti-bunching ratio g (2) (0)/g (2) (1),
its simple operation and low cost. Even though it can defined as a ratio of pulse areas in the autocorrelation
quite easily produce high-rate streams of photon-pairs, function, i.e. the area of the pulse around τ = 0 to
and thus also single-photons by heralding one photon on the area of its nearest neighbours, is more appropriate
the detection of the other – resulting in low values of the [50]. Table I is listing all these properties and parame-
g (2) autocorrelation, it suffers from the major drawback ters for several different implementations of single-photon
of not producing photons deterministically, but more or sources, along with the type of cavity being used. We will
15

Type Class Cavity V F g0 /2π γ⊥ /2π κ/2π C g (2) (0) τc RT η


[µm3 ] or Q† [MHz] [MHz] [MHz] g (2) (1) [ns] [MHz]

atom[52] III FPI 104 4 × 105 16 2.6 4.2 10 0.05 102 0.1 0.6
atom[60] III FPI 10 6
6 × 10 4
3.0 3.0 1.3 1.2 0.07 470 0.5 0.4
qdot[61] I MD ∼ 16 - - 80 - - 0.29 3.4 82 -
qdot[53] I MP 5 − 80 1000† 104 80 3 × 105 2.5 < 0.04 0.2 77 0.0016
qdot[62, 63] I PC 1 3000† 5000 80 105 2.5 0.54 0.06 80 -
ion[64] II/III FPI 10 6
7 × 10 5
1.6 11 0.05 2.3 < 0.02 < 105 0.002 0.88
ion[34] II/III FPI 106 104 1.4 0.8 1.2 1.0 < 0.02 102 0.1 0.08

TABLE I: Typical performances for different types and classes of cavity-based systems: single atoms, single ions, and solid-
state (quantum dots). Cavity parameters; V : mode volume; F: finesse (or Q-factor); g0 : atom-cavity coupling; γ⊥ : polarisation
decay rate; κ: cavity-field decay rate; C: cooperativity (Purcell factor f = 2C). Photon benchmark parameters (only if the
cavity has been used for single-photon generation); g (2) (0)/g (2) (1): ratio of pulse areas in the intensity auto-correlation (see
text); τc : coherence time (underlined values verified by two-photon interference); RT : single-photon trigger rate; η: combined
efficiency for single-photon generation and escape; (RT and η given for the operational mode in case of blinking sources). The
cavity types stated in the table are FPI: Fabry-Perot; MD: micro disk; MP: micro pillars; PC: photonic crystal. Note that some
of the listed parameters were not given in the original publications and therefore are inferred from other information, like, e.g.,
g0 for the three quantum-dot systems. Also the efficiency for the actual photon generation and escape from the cavity was not
specified in all cases.

now discuss these approaches in more detail. random times, no sub-Poissonian behaviour is found. A
possible way to overcome this minor drawback is by con-
ditioning the experiment on the actual presence of an
A. Neutral single atoms atom in the cavity [49]. A good example for doing so is
the characterisation of the photons by two-photon inter-
ference discussed in section IV B. For these experiments,
A straightforward implementation of a cavity-based pairs of photons are needed that meet simultaneously at
single-photon source consists of a single atom placed be- a beamsplitter. As discussed previously, see Fig. 12, this
tween two separate cavity mirrors, with a stream of laser is achieved by delaying the first photon of a pair of suc-
pulses travelling perpendicular to the cavity axis to trig- cessively emitted photons. With the occurrence of these
ger photon emissions. The most simplistic approach to photon pairs being the precondition to observe any corre-
achieve this is by sending a dilute atomic beam through lation and the probability for successive photon emissions
the cavity, with an average number of atoms in the mode being vanishingly small without atoms, the presence of
far below one. However, for a thermal beam, the obvious an atom is actually assured whenever data is recorded.
drawback would be an interaction time between atom and
cavity far too short to achieve any control on the exact For instance, two-photon interference has been stud-
photon emission time. Hence cold (and therefore slow) ied in a time-resolved manner in these experiments. For
atoms are required to overcome this limitation. One of us photons of 1 µs duration, a dip-width (and therefore co-
actually followed this route [33], using a magneto-optical herence time) of 470 ns is found, which is about one order
trap to cool a cloud of atoms to a few tens of µK at a dis- of magnitude longer than the decay rate of either cavity
tance of 20 cm above the cavity. Atoms released from the or atom. This proves that the photon’s coherence is to
trap eventually fall through the cavity, and interact with a large extend controlled by the Raman process, without
its mode for 17.5 µs. Within this limited interaction time, being limited by the decay channels within the system.
up to seven single-photon emissions have been triggered. Figure 15 shows the coincidences measured behind the
Fig. 14 illustrates this setup, together with some possi- beam splitter of the Hong-Ou Mandel experiment as a
bilities of using either transitions between hyperfine or function of the detection time delay. In this particular
Zeeman substates in rubidium to establish the adiabatic case, 87 Rb-atoms are excited in a way that the cavity
passage technique discussed in section III C to generate emits photons of alternating polarisation [60]. Coinci-
single photons. dences are of course found if photons of perpendicular
Bursts of single photons are emitted from the cavity polarisation impinge on the beam splitter, as they do not
whenever a single atom passes its mode, and strong an- interfere. This signal constitutes a nice reference that can
tibunching is found in the photon statistics, as shown easily be compared to the coincidence rate found if the
in Fig. 7c+d. Nonetheless, with the atoms arriving at photons are nearly indistinguishable, i.e. with parallel
16

F=
x

III. dynamic
Pump Atom-

co
x
y
laser cavity
vi t

he
Magneto-
ΩP(t) coupling
l ca

STIRAP
optical trap

ren
2g
smal

ce
co it y e
op
era g bil
lla
tiv
ity n tro
co
Cavity Photon
emission
I. spontaneous II. steady-state
x Time
x Photo diodes
losses

e Laser
pulses Time
g g

FIG. 14: Scheme of the first single-photon source based on


an atom-cavity system [33]. A cloud of atoms released from
a magneto-optical trap falls through a cavity 20 cm below.
FIG. 13: A simple classification scheme for cavity-based Laser pulses controlling the emission travel perpendicular to
single-photon sources. Quantum dots operate today mainly the cavity axis. The light is analysed using a Hanbury-Brown
as class I, with very small cavity volumes and hence very high & Twiss setup with a pair of photodiodes of 50% quantum
coupling factors, but suffer losses from incoherent atomic and efficiency. The relevant energy levels and transitions in 85 Rb
cavity decay. With ions, a substantially higher degree of co- are shown as well. The atomic states labeled |ei, |xi and |gi
herence can be reached in class II to III. In class II, the idea are involved in the Raman process, and the states |0i and |1i
is to get the photon out as fast as possible since one cannot denote the photon number in the cavity.
avoid spontaneous emission losses. With the even smaller cav-
ity volumes available to atoms, full dynamic control is reached
in class III with high cooperativity parameters. Since spon-
taneous losses are then avoided, we can pull the photon out ton emitters in Fabry-Perot type cavities is that a rel-
slower with greater probability. Ideally, future advances al- atively short cavity (some 100 µm) of high finesse (be-
low to combine single atoms with photonic crystal or fibre-tip tween 105 and 106 ) can be used. As stated in table I, one
cavities, such that class III operation at very small cavity thus obtains strong atom-cavity coupling, and the pho-
volumes become possible. ton generation can be driven either in the steady-state
regime (class II) or dynamically by vacuum-stimulated
Raman adiabatic passage (V-STIRAP, class III). This
allows one to control the coherence properties and the
polarisation. As can be seen from Fig. 15, almost no shape of the photons to a large extent, as discussed in
more coincidence counts are found in this case, which section 3.4. Photon generation efficiencies as high as
proves the mutual coherence of successively emitted pho- 40% have been demonstrated with these systems. Fur-
tons. thermore, based on the excellent coherence properties,
Only lately, refined versions of this type of photon first applications such as atom-photon entanglement and
emitter have been realised, with a single atom held in atom-photon state mapping [9, 10] have recently been
the cavity using a dipole-force trap. McKeever et al. [52] demonstrated. At first glance, a significant problem in all
managed to hold a single Cs atom in the cavity with these experiments seems the thermal motion of the atom
a dipole-trapping beam running along the cavity axis, in the surrounding dipole trap. The region explored by
while Hijlkema et al. [51] are using a combination of the atom brings it out of the Lamb-Dicke regime, and
dipole trapping beams running perpendicular and along the atomic motion in the trap leads to variations of the
the cavity to catch and hold a single Rb atom in the cav- atomic levels due to the dynamic Stark shift. However,
ity mode. As illustrated in Fig. 16, the trapped atom is these effect are fortunately not detrimental to transitions
in both cases exposed to a sequence of laser pulses alter- adiabatically driven between atomic ground states, as has
nating between triggering the photon emission, cooling been largely discussed in [71].
and repumping the atom to its initial state to repeat the Apart from the above Fabry-Perot type cavities, many
sequence. As the atom is trapped, the photon statistics other micro-structured cavities have been explored dur-
is no more affected by fluctuations in the atom number ing the last years. These usually provide a much smaller
and therefore is sub-Poissonian (see Fig. 7e). Moreover, mode volume and hence boost the atom-cavity coupling
with trapping times for single atoms up to a minute, a strength by about an order of magnitude. However, this
quasi-continuous bit-stream of photons is obtained. goes hand-in-hand with increased cavity losses and thus
The major advantage of using neutral atoms as pho- a much larger cavity linewidth, which might be in conflict
17

Two-Photon-Interference
with polarised photons

parallel
30 perp. σ+ σ− σ+ σ−
number of coincidences [a.u.]

cav

cav
ity
ity

las

las
er
er
25

cav
cav

las
las

er

er
ity

ity
20 +1 +1
0 0
-1 -1
15

10

0
-0.4 0 0.4 0.8
detection-time delay t [µs]

FIG. 15: Time-resolved two-photon interference signal for


polarised photons emitted from an atom-cavity system [60],
recorded with the setup depicted in Fig. 12. The dotted line
shows the correlations found when distinguishable photons of
perpendicular polarisation meet on the beam splitter, while
the solid line depicts the correlations found if the photon po-
larisation is parallel. The inset shows the level scheme in
87
Rb used to generate photons of alternating polarisation by
driving Raman transitions in between the magnetic sublevels
mF = ±1 of the F = 1 hyperfine state. As illustrated, a mag-
netic field along the cavity axis is used to lift the degeneracy
between magnetic sublevels.
FIG. 16: Atom-cavity systems with a single atom at rest in
the cavity mode. Top: One of us has been using a dipole
trap running perpendicular to the cavity axis is holding a
with the desired addressing of individual atomic transi- single rubidium atom trapped in the cavity [51]. The cav-
tions. Among the most relevant new developments are ity is asymmetric, and photons emitted through its output
fibre-tip cavities, which use dielectric Bragg stacks at the coupler are directed to a pair of photon counters to record
the second-order correlation function of the photon stream.
tip of an optical fibre as cavity mirrors [72, 73]. Due to
Bottom: J. McKeever et al. [52] are using a dipole trap
the small diameter of the fibre, either two fibre tips can running along the cavity axis is used to hold a single caesium
be brought very close together, or a single fibre tip can atom in the cavity. The cavity is symmetric, and only half the
be complemented by a micro-structured mirror on a chip photons are directed towards a pair of detectors for analysing
to form a high-finesse optical cavity. A slightly different the photon statistics (From J. McKeever et al., Science 303,
approach are ring-cavities realised in solid state, guiding 1992–1994 (2004). Reprinted with permission from AAAS).
the light in a whispering gallery mode. An atom can be In both cases, the trapped atom is exposed to a sequence of
easily coupled to the evanescent field of the cavity mode, laser pulses that trigger the photon emission, cool the atom,
provided it can be brought close to the surface of the and re-establish the initial condition by optical pumping.
substrate. Nice examples are microtoroidal cavities re-
alised at the California Institute of Technology [74, 75],
and bottle-neck cavities in optical fibres [76]. These cav-
ities have no well-defined mirrors and therefore no out- and has proven to be a useful method to acquire infor-
put coupler, so one usually arranges for emission into mation on the atom statistics. Last but not least, large
well-defined spatio-temporal modes via evanescent-field efforts have been made to study cavity-mediated forces
coupling to the core of an optical fibre. on either single atoms or atomic ensembles [82–87], which
eventually lead to the development of cavity-mediated
We would like to remind the reader at this point that
cooling techniques.
a large variety of other exciting cavity-QED experiments
has been performed that were not aiming at a single-
photon emission in the optical regime. Most important
amongst these are the coupling of Rydberg atoms [77] B. Trapped single ions
or superconductive SQUIDs [78] to microwave cavities,
which is also a well-established way of placing single pho- Although neutral-atom systems have their advantages
tons into a cavity using either π pulses [79] or dark reso- for the generation of single photons, such experiments are
nances [80]. Also the coupling of ultracold quantum gases sometimes hampered by fluctuating atomcavity coupling
to optical cavities has been studied extensively [72, 81], and multi-atom effects. Also trapping times are still lim-
18

ited in the intra-cavity dipole-trapping of single atoms. favourable mode structure of the cavity and far-off res-
A possible way around these problems is to use a strongly onant Raman transitions between magnetic sublevels of
localised single ion in an optical cavity, as has first been the trapped ion.
demonstrated by M. Keller et al. [34]. In their experi-
ment, an ion is optimally coupled to a well-defined field
mode, resulting in the reproducible generation of single- C. Artificial atoms in solid-state
photon pulses with precisely defined timing. The stream
of emitted photons is uninterrupted over the storage time Excellent reviews of photon sources based on artifi-
of the ion, which, in principle, could last for several days. cial atoms in solid-state systems have been published
by M. Oxborrow and A. G. Sinclair [89], B. Lounis and
M. Orrit [90] and A. J. Shields [91]. Therefore we only
discuss the basic features and key technologies here.
Quantum dots are often considered as artificial atoms
par excellence, as they usually possess several discrete en-
ergy levels for electron-hole pairs, with optical transitions
between these levels comparable to electronic transitions
in atoms. In first experiments, self-assembled individual
quantum dots have been exposed to either exciting laser
pulses or to a continuous beam, and the fluorescence was
collected and analysed using a microscope of high nu-
merical aperture. While antibunching has been found in
the light emitted from a single quantum dot [44, 69, 70],
the isotropic emission of photons resulted in a very low
photon-collection efficiency.

FIG. 17: Arrangement of ion-trap electrodes and cavity in


the experiment by M. Keller et al. [34]. The ion is shuttled
to the cavity region after loading. Upon excitation of the ion
from the side of the cavity, a single photon gets emitted into
the cavity mode. Reprinted by permission from Macmillan
Publishers Ltd: Nature (G. Guthöhrlein et al. [88]), copyright
2001.

The major difficulty in combining an ion trap with


a high-finesse optical cavity comes from the dielectric
cavity mirrors, which influence the trapping potential if
they get too close to the ion. This effect might be detri- FIG. 18: The microdisk toroidal cavity used by P. Michler et
mental in case the mirrors get electrically charged dur- al. [61] consists of a 5 µm-diameter disk and a 0.5 µm post.
ing loading of the ion trap, e.g. by the electron beam The cavity has a whispering gallery mode running around the
used to ionise the atoms. Fig. 17 shows how this prob- 200 nm thick GaAs disk, which is containing InAs quantum
lem has been solved in [34] by shuttling the trapped ion dots. From P. Michler et al., Science 290, 2282–2285 (2000).
from a spatially separate loading region into the cavity. Reprinted with permission from AAAS.
Nonetheless, the cavity in these experiments is typically
more than 10–20 mm long to avoid distortion of the trap. To improve the probability for photon collection, sev-
Thus the coupling to the cavity is weak, and albeit op- eral research teams successfully coupled individual quan-
timised pump pulses were used, the single-photon effi- tum dots to micro-structured cavities, consisting either of
ciency in [34] did not exceed (8.0 ± 1.3)%. This is in micro disks guiding a whispering gallery mode [61] (see
good accordance with theoretical calculations, which also Fig. 18) or micro-pillars with Bragg-stacks of dielectric
show that the efficiency can be substantially increased in surfaces [53, 92–97]. Fig. 19 shows a scanning electron
future experiments by reducing the cavity length. It is micrograph of such a pillar, along with a calculation of
important to point out that the low efficiency does not in- the electric field amplitude in the cavity mode. For all
terfere with the singleness of the photons. Hence the g (2) these systems, the cavity resonance frequency needs to
correlation function of the emitted photon stream corre- be equal to a transition frequency of the quantum dot.
sponds to the one depicted in Fig. 7e, with g (2) (0) → 0. Both can be altered by changing the temperature of the
Only recently, H. G. Barros et al. [64] have demonstrated whole setup, but the tuning range is often insufficient.
an improved version of an ion-cavity setup. Albeit their Nonetheless, the random scatter of resonance frequen-
cavity is of similar length length, they were able to reach cies available within a large ensemble of micro cavities
a single-photon efficiency of (88 ± 17)%, using a more and quantum dots normally allows one to find a match-
19

ing pair. Upon continuous or pulsed excitation, such shows the defect in the photonic crystal structure acting
a resonant pair of quantum-dot and cavity is emitting as high-Q cavity, and the modification of the spectrum in
sub-Poissonian light, with a photon statistics similar to response to the cavity-etching process. As in experiments
the one shown in Fig. 7(e). Furthermore, two-photon in- with micro-pillar cavities, photon-antibunching has been
terference can be studied by delaying the first of two observed in the light scattered off the cavity.
successively emitted photons, so that both meet on a
beam splitter [53]. At zero time-delay between the im- 942 944 946 948
a 0 7 nm
pinging photons, the coincidence rate between the two c
beam-splitter exit ports shows a pronounced dip, which
proves that independently generated photons are mutu-
ally coherent. Target
exciton

b d

Cavity

942 944 946 948


Wavelength (nm)
FIG. 19: (a) Micro-pillar cavities used by D. C. Unitt et al.
[96] with a 2 µm diameter. (b,c) show the calculated field
amplitude of the cavity mode in the pillar. By courtesy of FIG. 20: Photonic crystal cavity coupled to a single buried
A. Shields (Toshiba UK), Copyright (2005) IOP Publishing quantum dot used by K. Hennessy et al. [63]. (a) AFM to-
Limited [96]. pography of the photonic crystal cavity aligned to a hill of
material arising from a quantum dot. (b) Field intensity of
the cavity mode, showing that the quantum dot overlaps the
In a variant of the aforementioned micro-pillar exper- field maximum. (c+d) Photoluminescence spectra of a sin-
iments, M. Benyoucef et al. [98] were able to produce gle quantum dot before and after cavity fabrication, showing
entangled photon pairs from a quantum dot coupled to emission from the cavity at 942.5 nm. Reprinted by permis-
a micro-pillar cavity. The two photons are emitted in sion from Macmillan Publishers Ltd: Nature (K. Hennessy et
a relaxation cascade and therefore generated almost si- al. [63]), copyright 2007.
multaneously in a single step. We emphasise that this is
substantially different from other approaches which pro- For many solid-state systems, the reduced light-
duce entangled photons one-after-the-other [9, 10]. Very collection efficiency seems to be a substantial limitation.
recently, more elaborate schemes for pulsed on-demand Some cavity designs have either no dedicated output cou-
sources have been proposed [99], which might even lead pler, or are plagued by a high jump in refractive index
to the emission of large photon clusters from single quan- between bulk and outside world. Due to that, the effi-
tum dots. With the photon emissions being caused by a ciencies listed for quantum dots in table I are either very
fast relaxation cascade, such a source would constitute a small or not stated at all. However, this does not consti-
‘photon machine gun’. tute a principal limit, as promising proposals have been
Possible alternatives to micro-disk or micro pillar cavi- made for accessing the photons using integrated optics
ties are photonic-bandgap cavities, which consist of artifi- [102]. One might also think that using artificial atoms
cial defects in an otherwise regular photonic crystal. Such embedded in solid state cavities will outperform most
cavities have the potential to allow for an easy coupling of other approaches, as the presence of a single emitter in
the light into optical fibres embedded on a chip surface, the cavity mode is assured once and forever – which un-
and have therefore been studied by a large number of doubtedly is the major advantage of the solid-state ap-
research teams [62, 63, 100, 101]. The defects in the pho- proach. However, no two quantum dots are equal, and
tonic crystal give rise to local trapping of light, and thus mutual coherence between photons from different sources
form an effective cavity on the nano-scale. The mode vol- is hard to achieve. Moreover, the finesse of the solid-state
ume of these cavities is typically of the order of λ3 , result- cavities is roughly one order of magnitude smaller than
ing in atom-cavity coupling strengths up to 100 GHz. To the typical finesse obtained with dielectric mirrors, and
bring a quantum dot and a photonic-crystal cavity into the solid-state cavities also have a very tiny volume. This
resonance, the team around A. Imamoglu has developed leads to a very fast photon loss from the cavity, com-
a digital etching technique [62], which works by chang- bined with a very strong atom-cavity coupling. In turn,
ing the radii of the crystal-forming holes, layer for layer the timescale for the process is rather short, and no adi-
of atoms, until the resonance condition is met. Fig. 20 abatic control of the emission process is possible. The
20

process typically starts with a driving laser pulse exciting radiation field, with efficiencies that can be very close to
the quantum dot to an energetically higher level, followed unity. Therefore these photons are ideal for all-optical
by a relaxation cascade down to the upper level used for quantum computation schemes, as proposed by Knill,
photon emission. The last step in this cascade then is Laflamme, and Milburn [2]. Moreover, the atom- and
the cavity-enhanced emission of the desired photon. The ion-based sources are expected to play a significant role
emission into the cavity is therefore dominantly sponta- in the implementation of quantum networking [11] and
neous (class I source), and the driving laser pulses have quantum communication schemes [109], as they have al-
barely any influence on the shape or coherence properties ready shown to be capable of entangling and mapping
of the photons. To some extend, this reduces the useful- quantum states between atoms and photons [9, 10]. Pro-
ness and in particular the scalability of present solid-state cesses like entanglement swapping between distant atoms
photon emitters. or ions, that have so far been studied without the aid of
Last but not least, we would like to emphasise that cavities [3, 4, 110–112] are expected to profit enormously
all the single-photon emitters we discussed so far re- from the introduction of cavity-based techniques, as their
quire an optical excitation of the emitter, be it an atom, success-probability scales with the square of the efficiency
ion, or quantum dot. Very recently, A. J. Shields and of the photon generation process. The high efficiency of
D. A. Ritchie have begun exploring a very promising al- cavity-based photon sources also opens new routes to-
ternative by implementing an electrically driven LED- wards a highly scalable quantum network, which is es-
like system that contains a single quantum dot acting sential for providing cluster states in one-way quantum
as emitter [70, 103]. They were successful in coupling computing [113] and for the quantum simulation of com-
the quantum dot to a VCSEL-like cavity [104–106], and plex solid-state systems [114].
have recently observed a pronounced Hong-Ou-Mandel
dip in the interference of two successively emitted pho-
tons [107, 108]. This new development could open up the
route towards versatile on-chip integrated single-photon Acknowledgments
emitters.

We gratefully acknowledge support by the En-


VI. FUTURE DIRECTIONS gineering and Physical Sciences Research Council
(EP/E023568/1), the Research Unit 635 of the Ger-
We have been reviewing many different ways of pro- man Research Foundation, and the EU through the re-
ducing single photons from simple quantum systems cou- search and training network EMALI (MRTN-CT-2006-
pled to optical cavities. Most of the photon-production 035369) and the integrated project SCALA. In addition,
methods lead to on-demand emission of narrowband and DL wishes to express his gratitude to the Swedish Re-
indistinguishable photons into a well defined mode of the search Council (VR) for financial support.

[1] D.P. DiVincenzo, Real and realistic quantum computers, terface, Science 317 (2007), p. 488.
Nature 393 (1998), pp. 113–114. [10] B. Weber et al., Photon-Photon Entanglement with a
[2] E. Knill, R. Laflamme, and G.J. Milburn, A scheme for Single Trapped Atom, Phys. Rev. Lett. (2009), 030501.
efficient quantum computing with linear optics, Nature [11] J.I. Cirac et al., Quantum State Transfer and Entangle-
409 (2001), pp. 46–52. ment Distribution Among Distant Nodes in a Quantum
[3] B.B. Blinov et al., Observation of entanglement between Network, Phys. Rev. Lett. 78 (1997), pp. 3221–3224.
a single trapped atom and a single photon, Nature 428 [12] D.P. DiVincenzo, The Physical Implementation of
(2004), pp. 153–157. Quantum Computation, Fortschr. Phys. 48 (2000), p.
[4] J. Volz et al., Observation of Entanglement of a Sin- 771.
gle Photon with a Trapped Atom, Phys. Rev. Lett. 96 [13] A.D. Boozer et al., Reversible State Transfer between
(2006), 030404. Light and a Single Trapped Atom, Phys. Rev. Lett. 98
[5] S. Olmschenk et al., Quantum Teleportation Between (2007), 193601.
Distant Matter Qubits, Science 323 (2009), pp. 486–489. [14] N. Gisin and R. Thew, Quantum Communication, Na-
[6] E.M. Purcell, Spontaneous Emission Probabilities at Ra- ture Photonics 1 (2007), pp. 165–171.
dio Frequencies, Phys. Rev. 69 (1946), p. 681. [15] E.T. Jaynes and F.W. Cummings, Comparison of Quan-
[7] H.J. Carmichael, Photon Antibunching and Squeezing tum and Semiclassical Radiation Theories with Appli-
for a Single Atom in a Resonant Cavity, Phys. Rev. cation to the Beam Maser, Proc. IEEE 51 (1963), pp.
Lett. 55 (1985), pp. 2790–2793. 89–109.
[8] F. De Martini et al., Anomalous Spontaneous Emission [16] B.W. Shore and P.L. Knight, The Jaynes-Cummings
Time in a Microscopic Optical Cavity, Phys. Rev. Lett. Model, J. Mod. Opt. 40 (1993), p. 1195.
59 (1987), pp. 2955–2958. [17] N.V. Vitanov et al., Coherent manipulation of atoms
[9] T. Wilk et al., Single-Atom Single-Photon Quantum In- and molecules by sequential laser pulses, Adv. At. Mol.
21

Opt. Phys. 46 (2001), pp. 55–190. 39 (1977), pp. 691–695.


[18] D.J. Heinzen et al., Enhanced and inhibited spontaneous [39] H.J. Carmichael et al., Intensity correlations in reso-
emission by atoms in a confocal resonator, Phys. Rev. nance fluorescence with atomic number fluctuations, J.
Lett. 58 (1987), pp. 1320–1323. Phys. A 11 (1978), pp. L121–L126.
[19] S.E. Morin, C.C. Yu, and T.W. Mossberg, Strong atom- [40] F. Diedrich and H. Walther, Nonclassical radiation of a
cavity coupling over large volumes and the observation single stored ion, Phys. Rev. Lett. 58 (1987), pp. 203–
of subnatural intracavity atomic linewidths, Phys. Rev. 206.
Lett. 73 (1994), pp. 1489–1492. [41] V. Gomer et al., Single-atom dynamics revealed by pho-
[20] L.A. Lugiato, Theory of optical bistability, in Progress ton correlations, Phys. Rev. A 58 (1998), pp. R1657–
in Optics, E. Wolf ed., Vol. XXI Elsevier Science Pub- R1660.
lishers, B. V., 1984, pp. 71–216. [42] J. McKeever et al., Experimental realization of a one-
[21] C.K. Law and J.H. Eberly, Arbitrary Control of a Quan- atom laser in the regime of strong coupling, Nature 425
tum Electromagnetic Field, Phys. Rev. Lett. 76 (1996), (2003), pp. 268–270.
p. 1055. [43] F. De Martini, G.D. Giuseppe, and M. Marrocco,
[22] C.K. Law and H.J. Kimble, Deterministic Generation Single-Mode Generation of Quantum Photon States by
of a Bit-Stream of Single-Photon Pulses, J. Mod. Opt. Excited Single Molecules in a Microcavity Trap, Phys.
44 (1997), pp. 2067–2074. Rev. Lett. 76 (1996), pp. 900–903.
[23] A. Kuhn and G. Rempe, Optical Cavity QED: Fun- [44] P. Michler et al., Quantum Correlation Among Photons
damentals and Application as a Single-Photon Light from a Single Quantum Dot at Room Temperature, Na-
Source, in Experimental Quantum Computation and In- ture 406 (2000), pp. 968–970.
formation, F. De Martini and C. Monroe, eds., Vol. 148, [45] C. Kurtsiefer et al., Stable Solid-State Source of Single
IOS-Press, Amsterdam, 2002, pp. 37–66. Photons, Phys. Rev. Lett. 85 (2000), pp. 290–293.
[24] A. Kuhn, M. Hennrich, and G. Rempe, Strongly- [46] A. Beveratos et al., Nonclassical radiation from dia-
Coupled Atom-Cavity Systems, in Quantum Information mond nanocrystals, Phys. Rev. A 64 (2001), 061802.
Processing, in T. Beth and G. Leuchs, eds., Wiley-VCH, [47] C. Jurczak et al., Atomic Transport in an Optical Lat-
Berlin, 2003, pp. 182–195. tice: An Investigation through Polarization-Selective In-
[25] S.E. Harris, Electromagnetically Induced Transparency tensity Correlations, Phys. Rev. Lett. 77 (1996), pp.
with Matched Pulses, Phys. Rev. Lett. 70 (1993), pp. 1727–1730.
552–555. [48] S. Bali et al., Measurements of intensity correlations of
[26] S.E. Harris, Electromagnetically Induced Transparency, scattered light from laser-cooled atoms, Phys. Rev. A 53
Phys. Today 50 (1997), p. 36. (1996), pp. 3469–3472.
[27] L.V. Hau et al., Light speed reduction to 17 metres per [49] M. Hennrich, A. Kuhn, and G. Rempe, Transition from
second in an ultracold atomic gas, Nature 397 (1999), antibunching to bunching in cavity QED, Phys. Rev.
pp. 594–598. Lett. 94 (2005), 053604.
[28] D.F. Phillips et al., Storage of Light in Atomic Vapor, [50] M. Hennrich et al., Photon statistics of a non-stationary
Phys. Rev. Lett. 86 (2001), pp. 783–786. periodically driven single-photon source, New J. Phys. 6
[29] A. Messiah, 1958, Vol. 2, chapter 17. in Quantum Me- (2004), p. 86.
chanics J. Wiley & Sons, NY. [51] M. Hijlkema et al., A single-photon server with just one
[30] A. Kuhn et al., Controlled Generation of Single Photons atom, Nature Physics 3 (2007), pp. 253–255.
from a Strongly Coupled Atom-Cavity System, Appl. [52] J. McKeever et al., Deterministic Generation of Single
Phys. B 69 (1999), pp. 373–377. Photons from One Atom Trapped in a Cavity, Science
[31] A.S. Parkins et al., Synthesis of Arbitrary Quantum 303 (2004), pp. 1992–1994.
States Via Adiabatic Transfer of Zeeman Coherence, [53] C. Santori et al., Indistinguishable photons from a
Phys. Rev. Lett. 71 (1993), p. 3095. single-photon device, Nature 419 (2002), pp. 594–597.
[32] P.P. Rohde, T.C. Ralph, and M.A. Nielsen, Optimal [54] F. Jelezko et al., Coherence length of photons from a
photons for quantum-information processing, Phys. Rev. single quantum system, Phys. Rev. A 67 (2003), 041802.
A 72 (2005), 052332. [55] C.K. Hong, Z.Y. Ou, and L. Mandel, Measurement of
[33] A. Kuhn, M. Hennrich, and G. Rempe, Deterministic Subpicosecond Time Intervals between Two Photons by
single-photon source for distributed quantum network- Interference, Phys. Rev. Lett. 59 (1987), pp. 2044–2046.
ing, Phys. Rev. Lett. 89 (2002), 067901. [56] T. Legero et al., Time-Resolved Two-Photon Quantum
[34] M. Keller et al., Continuous generation of single photons Interference, Appl. Phys. B 77 (2003), pp. 797–802.
with controlled waveform in an ion-trap cavity system, [57] T. Legero et al., Quantum beat of two single photons,
Nature 431 (2004), pp. 1075–1078. Phys. Rev. Lett. 93 (2004), 070503.
[35] G.S. Vasilev, D. Ljunggren, and A. Kuhn, Sin- [58] T. Legero et al., Characterization of single photons us-
gle Photons Made-to-Measure, submitted (2009) ing two-photon interference, Adv. At. Mol. Opt. Phys.
arXiv:0907.0761v1 [quant-ph]. 53 (2006), p. 253.
[36] R.H. Brown and R.Q. Twiss, A test of a new type of [59] C.O. Alley and Y.H. Shih, A new type of EPR experi-
stellar interferometer on Sirius, Nature 178 (1956), pp. ment, in Proceedings of the Second International Sym-
1046–1448. posium on Foundations of Quantum Mechanics in the
[37] L. Mandel and E. Wolf Optical coherence and quantum Light of New Technology, M.Namiki et al. ed., Tokyo,
optics, Cambridge University Press, Cambridge, UK, 1986, pp. 47–52.
1995. [60] T. Wilk et al., Polarization-Controlled Single Photons,
[38] H.J. Kimble, M. Dagenais, and L. Mandel, Photon An- Phys. Rev. Lett. 98 (2007), 063601.
tibunching in Resonance Fluorescence, Phys. Rev. Lett. [61] P. Michler et al., A Quantum Dot Single Photon Turn-
22

stile Device, Science 290 (2000), pp. 2282–2285. [85] P. Maunz et al., Cavity cooling of a single atom, Nature
[62] A. Badolato et al., Deterministic Coupling of Single 428 (2004), pp. 50–52.
Quantum Dots to Single Nanocavity Modes, Science 308 [86] J.K. Thompson et al., A High-Brightness Source of Nar-
(2005), pp. 1158–1161. rowband, Identical-Photon Pairs, Science 313 (2006),
[63] K. Hennessy et al., Quantum nature of a strongly pp. 74–77.
coupled single quantum dot-cavity system, Nature 445 [87] K.M. Fortier et al., Deterministic Loading of Individ-
(2007), pp. 896–899 ual Atoms to a High-Finesse Optical Cavity, Phys. Rev.
[64] H.G. Barros et al., Deterministic single-photon source Lett. 98 (2007), 233601.
from a single ion, New J. Phys. 11 (2009), 103004. [88] G.R. Guthörlein et al., A single ion as a nanoscopic
[65] J. Kim et al., A Single Photon Turnstile Device, Nature probe of an optical field, Nature 414 (2001), pp. 49–51.
397 (1999), pp. 500–503. [89] M. Oxborrow and A.G. Sinclair, Single-Photon Sources,
[66] C. Brunel et al., Triggered Source of Single Pho- Contemporary Physics 46 (2005), pp. 173–206.
tons Based on Controlled Single Molecule Fluorescence, [90] B. Lounis and M. Orrit, Single-Photon Sources, Rep.
Phys. Rev. Lett. 83 (1999), pp. 2722–2725. Prog. Phys. 68 (2005), pp. 1129–1179.
[67] B. Lounis and W.E. Moerner, Single Photons on De- [91] A.J. Shields, Semiconductor quantum light sources, Na-
mand from a Single Molecule at Room Temperature, Na- ture Photonics 1 (2007), pp. 215–223.
ture 407 (2000), pp. 491–493. [92] O. Benson et al., Regulated and entangled photons from
[68] R. Brouri et al., Photon Antibunching in the Fluores- a single quantum dot, Phys. Rev. Lett. 84 (2000), pp.
cence of Individual Color Centers in Diamond, Opt. 2513–2516.
Lett. 25 (2000), pp. 1294–1296. [93] E. Moreau et al., Single-mode solid-state single photon
[69] C. Santori et al., Triggered Single Photons from a Quan- source based on isolated quantum dots in pillar micro-
tum Dot, Phys. Rev. Lett. 86 (2001), pp. 1502–1505. cavities, Appl. Phys. Lett. 79 (2001), pp. 2865–2867.
[70] Z. Yuan et al., Electrically Driven Single-Photon Source, [94] M. Pelton et al., An Efficient Source of Single Pho-
Science 295 (2002), pp. 102–105. tons: A Single Quantum Dot in a Micropost Microcav-
[71] L.M. Duan, A. Kuzmich, and H.J. Kimble, Cavity QED ity, Phys. Rev. Lett. 89 (2002), 233602.
and quantum information processing with “hot” trapped [95] A.J. Bennett et al., High performance single photon
atoms, Phys. Rev. A 67 (2003), 032305. sources from photolithographically defined pillar micro-
[72] Y. Colombe et al., Strong atom-field coupling for Bose- cavities, Optics Express 13 (2005), pp. 50–55.
Einstein condensates in an optical cavity on a chip, Na- [96] D.C. Unitt et al., Quantum dots as single-photon sources
ture 450 (2007), pp. 272–276. for quantum information processing, Journal of Optics
[73] M. Trupke et al., Atom detection and photon production B: Quantum and Semiclassical Optics 7 (2005), pp.
in a scalable, open, optical microcavity, Phys. Rev. Lett. S129–S134.
99 (2007), 063601. [97] D. Press et al., Photon Antibunching from a Single
[74] B. Dayan et al., A Photon Turnstile Dynamically Regu- Quantum-Dot-Microcavity System in the Strong Cou-
lated by One Atom, Science 319 (2008), pp. 1062–1065. pling Regime, Phys. Rev. Lett. 98 (2007), 117402.
[75] T. Aoki et al., Efficient routing of single photons by one [98] M. Benyoucef et al., Enhanced correlated photon pair
atom and a microtoroidal cavity, Phys. Rev. Lett. 102 emission from a pillar microcavity, New J. Phys. 6
(2009), 083601. (2004), p. 91.
[76] M. Pöllinger et al., Ultrahigh-Q tunable whispering- [99] N.H. Lindner and T. Rudolph, Proposal for Pulsed
gallery-mode microresonator, Phys. Rev. Lett. 103 On-Demand Sources of Photonic Cluster State Strings,
(2009), 053901. Physical Review Letters 103(11) (2009), 113602.
[77] A. Rauschenbeutel et al., Step by step engineered many [100] A. Kress et al., Manipulation of the spontaneous emis-
particle entanglement, Science 288 (2000), p. 2024. sion dynamics of quantum dots in two-dimensional pho-
[78] A. Wallraff et al., Strong coupling of a single photon to tonic crystals, Phys. Rev. B 71 (2005), 241304.
a superconducting qubit using circuit quantum electro- [101] S. Laurent et al., Indistinguishable single photons from a
dynamics, Nature 431 (2004), p. 162. single-quantum dot in a two-dimensional photonic crys-
[79] X. Maı̂tre et al., Quantum Memory with a Single Photon tal cavity, Appl. Phys. Lett. 87 (2005), 163107.
in a Cavity, Phys. Rev. Lett. 79 (1997), pp. 769–772. [102] W. Barnes et al., Solid-state single photon sources: light
[80] S. Brattke, B.T.H. Varcoe, and H. Walther, Generation collection strategies, The European Physical Journal D
of Photon Number States on Demand via Cavity Quan- - Atomic, Molecular, Optical and Plasma Physics 18
tum Electrodynamics, Phys. Rev. Lett. 86 (2001), pp. (2002), pp. 197–210.
3534–3537. [103] M.B. Ward et al., Electrically driven telecommunication
[81] F. Brennecke et al., Cavity QED with a Bose-Einstein wavelength single-photon source, Appl. Phys. Lett. 90
condensate, Nature 450 (2007), pp. 268–271. (2007), 063512.
[82] V. Vuletić and S. Chu, Laser cooling of atoms, ions, [104] A.J. Bennett et al., Microcavity single-photon-emitting
or molecules by coherent scattering, Phys. Rev. Lett. 84 diode, Appl. Phys. Lett. 86 (2005), 181102.
(2000), pp. 3787–3790. [105] D.J.P. Ellis et al., Oxide-apertured microcavity single-
[83] V. Vuletić, H.W. Chan, and A.T. Black, Three- photon emitting diode, Appl. Phys. Lett. 90 (2007),
dimensional cavity Doppler cooling and cavity sideband 233514.
cooling by coherent scattering, Phys. Rev. A 64 (2001), [106] D.J.P. Ellis et al., Cavity-enhanced radiative emission
033405. rate in a single-photon-emitting diode operating at 0.5
[84] J. McKeever et al., State-insensitive cooling and trap- GHz, New J. Phys. 10 (2008), 043035.
ping of single atoms in an optical cavity, Phys. Rev. [107] R.B. Patel et al., Postselective Two-Photon Interfer-
Lett. 90 (2003), 133602. ence from a Continuous Nonclassical Stream of Photons
23

Emitted by a Quantum Dot, Phys. Rev. Lett. 100 (2008), (2007), pp. 538–541.
207405. [113] R. Raussendorf and H.J. Briegel, A One-Way Quantum
[108] A.J. Bennett et al., Indistinguishable photons from a Computer, Phys. Rev. Lett. 86 (2001), pp. 5188–5191.
diode, Appl. Phys. Lett. 92 (2008), 193503. [114] H.P. Büchler et al., Atomic Quantum Simulator for Lat-
[109] H.J. Briegel et al., Quantum Repeaters: The Role of Im- tice Gauge Theories and Ring Exchange Models, Phys.
perfect Local Operations in Quantum Communication, Rev. Lett. 95 (2005), 040402.
Phys. Rev. Lett. 81 (1998), pp. 5932–5935. [115] We use the term atom throughout the whole paper
[110] B. Sun, M.S. Chapman, and L. You, Atom-photon en- as a synonym for any quantum system showing dis-
tanglement generation and distribution, Phys. Rev. A crete energy levels with radiative transitions between
69 (2004), 042316. them. Amongst these are neutral atoms, ions, quantum
[111] J. Beugnon et al., Quantum interference between two dots, SQUIDs, Rydberg atoms and many other artificial
single photons emitted by independently trapped atoms, quantum systems with atom-like properties.
Nature 440 (2006), pp. 779–782. [116] Note that the population decay rate of the atom is 2γ⊥ ,
[112] P. Maunz et al., Quantum interference of photon pairs and the photon loss rate from the cavity is 2κ.
from two remote trapped atomic ions, Nature Physics 3

You might also like