Quantum-enhanced Metrology without Entanglement
based on Optical Cavities with Feedback
Lewis A. Clark, Adam Stokes, M. Mubashir Khan, Gangcheng Wang and Almut Beige
The School of Physics and Astronomy, University of Leeds, Leeds LS2 9JT, U.K.
Keywords:
Quantum Metrology, Optical Cavities, Quantum Feedback.
Abstract:
There are a number of different strategies to measure the phase shift between two pathways of light more
efficiently than suggested by the standard quantum limit. One way is to use highly entangled photons. Another
way is to expose photons to a non-linear or interacting Hamiltonian. This paper emphasises that the conditional
dynamics of open quantum systems provides an interesting additional tool for quantum-enhanced metrology.
As a concrete example, we review a recent scheme which exploits the conditional dynamics of a laser-driven
optical cavity with spontaneous photon emission inside a quantum feedback loop. Deducing information from
second-order photon correlation measurements requires neither optical non-linearities nor entangled photons
and should therefore be of immediate practical interest.
1 INTRODUCTION
This paper compares different strategies to decrease
the uncertainty ∆ϕ for measurements of an unknown
phase shift ϕ between two pathways of light when
given a certain amount of resources N. Using N in-
dependent resources, i.e. deducing ϕ from a measure-
ment signal M(ϕ) obtained from the dynamics of a
generator which is linear in N, the scaling of the lower
bound of the uncertainty ∆ϕ of the phase measure-
ment with respect to N is given by the standard quan-
tum limit (Dowling and Seshadreesan, 2015),
∆ϕ
class
N
0.5
. (1)
However, using for example N highly-entangled pho-
tons as a resource, the measurement uncertainty ∆ϕ
can be as low as the Heisenberg limit,
∆ϕ
quant
N
1
. (2)
An alternative way of enhancing metrology beyond
the standard quantum limit is to expose N incoming
single photons to a ϕ-dependent Hamiltonian which
is no longer linear in N (Boixo et al., 2008). In this
case, the uncertainty ∆ϕ of the phase measurement
scales as
∆ϕ
nonlin
N
0.5k
, (3)
where k denotes the order of non-linearity of the
interaction Hamiltonian with respect to N or de-
scribes the interaction between k phase dependent
systems. However, multi-photon entanglement and
highly-efficient optical non-linearities are hard to im-
plement experimentally and have not yet become
readily available for a wide range of applications.
Some suggested schemes avoiding entanglement to
get an enhancement include the use of non-linear op-
tics (Luis, 2007) or using squeezed states (Caves,
1981).
In some situations, the length of the measurement
process, t, and not the number of incoming photons,
N, is the main resource which we want to constrain
(Clark et al., 2016b). If we can write the time t in
such cases as t = N t with t being the length of a
short time interval, then the standard quantum limit
still coincides with Eq. (1). In the following we em-
phasise that the environmental interactions of open
quantum systems naturally result in conditional dy-
namics, like quantum jumps (Blatt and Zoller, 1988),
which can be exploited for quantum computational
tasks (Beige et al., 2000; Lim et al., 2005; Metz et al.,
2006; Clark et al., 2015). It is shown that the gen-
erators of the conditional dynamics can introduce a
non-linear resource-dependence with respect to time
(Clark et al., 2016c). Applying this observation to
quantum metrology provides an additional tool which
allows us to break the standard quantum limit (1) and
helps us to design scheme which can be implemented
relatively easily.
As an example, we review a recent quantum-
enhanced metrology scheme by Clark et al. (Clark
et al., 2016b) which requires only a laser-driven op-
tical cavity inside a quantum-feedback loop. The pro-
posed setup is feasible with current technology (Kuhn
A. Clark L., Stokes A., Mubashir Khan M., Wang G. and Beige A.
Quantum-enhanced Metrology without Entanglement based on Optical Cavities with Feedback.
DOI: 10.5220/0006141702230229
In Proceedings of the 5th International Conference on Photonics, Optics and Laser Technology (PHOTOPTICS 2017), pages 223-229
ISBN: 978-989-758-223-3
Copyright
c
2017 by SCITEPRESS Science and Technology Publications, Lda. All rights reserved
223
et al., 2002; McKeever et al., 2004). Differing from
Ref. (Clark et al., 2016b), this paper does not pay as
much attention to the concrete analysis of the pro-
posed scheme. Instead, it focusses its attention on
what we believe to be the main mechanisms under-
lying the observed enhancement. Our findings com-
plement the work of other authors (Braun and Martin,
2011; Macieszczak et al., 2016; Pearce et al., 2015),
who also observe that quantum-enhancements may
be obtained from quantum correlations and sequential
measurements in open quantum systems.
There are five sections in this paper. Section 2 re-
views the main theoretical models that are commonly
used to describe quantum optical systems with spon-
taneous photon emission. In Section 3, we discuss
how to use the non-linear conditional dynamics of
an open quantum system to measure the phase shift
ϕ between two pathways of light, using the work of
Ref. (Clark et al., 2016b) as an example. Section 4
emphasises that the observed quantum enhancement
is not unexpected by showing that subsequent mea-
surements on a single quantum system provide at least
as much information as single-shot measurements on
entangled states. Finally, we summarise our findings
in Section 5.
2 THE QUANTUM JUMP
DYNAMICS OF OPEN
QUANTUM SYSTEMS
In this section, we review the modelling of open
quantum systems with spontaneous photon emission
(Hegerfeldt, 1993; Dalibard et al., 1992; Carmicheal,
1993; Stokes et al., 2012), thereby emphasising that
the conditional dynamics of single quantum trajecto-
ries, which is associated with quantum jumps, is nat-
urally non-linear.
2.1 The Schr
¨
odinger Equation of
Closed Quantum Systems
As is well known, the dynamics of a closed quan-
tum system in the Schr
¨
odinger picture is given by the
Schr
¨
odinger equation,
|
˙
ψi =
i
~
H |ψi, (4)
where |ψi is a state vector and H is the time-
independent Hamiltonian and energy observable.
Solving the above equation for a given initial state
|ψ(0)i yields the state vector |ψ(t)i,
|ψ(t)i = exp
i
~
Ht
|ψ(0)i. (5)
This shows that the generator for the dynamics of a
closed quantum system is linear in time.
Suppose the Hamiltonian H depends on an un-
known parameter ϕ which we want to measures as
accurately as possible. Then the above closed system
dynamics, i.e. the time evolution of a measurement
signal M(t,ϕ), can be used to deduce ϕ. The longer
the system is observed and the larger the measure-
ment time t, the more information about ϕ becomes
available and the smaller the uncertainty ∆ϕ of this
measurement can become. However ∆ϕ is limited by
the standard quantum limit (1) with respect to the re-
source N which measures time.
2.2 Master Equations of Open
Quantum Systems
When interacting with an environment, the expecta-
tion values of physical observables averaged over an
ensemble of quantum systems can no longer be de-
duced from a single state vector |ψi. The ensem-
ble now needs to be described by a density matrix ρ.
For Markovian systems, ρ necessarily obeys a master
equation of Lindblad form. For example, the mas-
ter equation of an open quantum system with sponta-
neous photon emission and only a single decay chan-
nel can be written as (Stokes et al., 2012)
˙
ρ = L ρ (6)
with the linear superoperator L given by
L ρ =
i
~
[H,ρ] +
1
2
Γ
2LρL
h
L
L,ρ
i
+
.
(7)
Here Γ denotes the spontaneous decay rate and L is
the so-called Lindblad operator. In the case of an op-
tical cavity, L simply coincides with the cavity pho-
ton annihilation operator c and Γ becomes the spon-
taneous cavity decay rate κ. In analogy to Eq. (5) we
now find that
ρ(t) = exp (Lt) ρ(0), (8)
where ρ(0) denotes the initial state of the open quan-
tum system. As in the previous subsection, we see
that the generator of the system dynamics is linear in
time. This is analogously to the situation of closed
quantum systems in the previous subsection. The
standard quantum limit (1) still gives a lower bound
for the uncertainty ∆ϕ for the measurement of an un-
known parameter ϕ with respect to time.
2.3 Unravelling into Quantum
Trajectories
Having a closer look at microscopic derivations of
Eq. (6) (see e.g. Refs. (Hegerfeldt, 1993; Dalibard
PHOTOPTICS 2017 - 5th International Conference on Photonics, Optics and Laser Technology
224
et al., 1992; Carmicheal, 1993; Stokes et al., 2012))
shows that Eq. (6) is the result of averaging over
all possible quantum trajectories that the quantum
system can experience. If the system is continu-
ously monitored and all its photon emission times are
known, then it can still be described by a pure state
vector at all times. To point out the relevant unravel-
ling of Eq. (6), we now write the derivative
˙
ρ as
˙
ρ =
i
~
h
H
cond
ρ ρH
cond
i
+ ΓLρL
(9)
with H
cond
being the non-Hermitian conditional
Hamiltonian given by
H
cond
= H
i
2
~ΓL
L (10)
and with L being the same Lindblad operator as in
Eq. (7). Re-writing our master equation in this com-
pact form allows us to identify two subensembles.
The first term in Eq. (9) is the time derivative of
the unnormalised density matrix of the subensemble
of quantum systems without a photon emission in a
small time interval (t,t + t), while the second term
refers to the subensemble experiencing an emission.
The normalisation of both terms indicates their rela-
tive sizes.
More concretely, under the condition of no photon
emission in (t,t + t), the state vector |ψ(t)i evolves
such that
|ψ(t + t)i = K
0
|ψ(t)i/kK
0
|ψ(t)ik (11)
with the operator K
0
being the conditional no-photon
time evolution operator
K
0
= exp
i
~
H
cond
t
. (12)
The generator of the no-photon time evolution of our
open quantum system is again linear in time. How-
ever, in case of a photon detection in (t,t + t), the
state vector of the quantum system changes into
|ψ(t + t)i = K
1
|ψ(t)i/kK
1
|ψ(t)ik (13)
with the reset operator K
1
given by
K
1
= (Γ∆t)
1/2
L . (14)
In the case of a photon emission, a so-called quan-
tum jump occurs accompanied by sudden jumps of
expectation values (Blatt and Zoller, 1988). The gen-
erators of the conditional dynamics of open quantum
systems, which are associated with spontaneous pho-
ton emission, are in general highly non-linear due to
the constant need for resetting upon emission.
3 TEMPORAL QUANTUM
CORRELATIONS FOR
QUANTUM METROLOGY
Section 1 emphasises that generators of dynamics,
which are non-linear in the relevant resource, can be
used to measure an unknown parameter ϕ with an ac-
curacy ∆ϕ beyond the standard quantum limit. Sec-
tion 2 shows that the generators for the dynamics
of the single quantum trajectories of open quantum
systems are in general non-linear. Combining these
two observations, one can design novel quantum-
enhanced metrology schemes which require neither
entanglement nor a non-linear or interacting Hamil-
tonian. To illustrate this fact we now review a recent
proposal by Clark et al. (Clark et al., 2016b) based
on a single optical cavity inside a quantum feedback
loop for which time is the resource which we want to
constrain.
3.1 A Quantum-enhanced Metrology
Scheme
The quantum metrology scheme in Ref. (Clark et al.,
2016b) consists of two main stages, a preparation and
a measurement stage. These implement the following
tasks:
1. Firstly, the preparation stage prepares the cavity
field in a coherent state |αi with α being of the
form
α = |α|e
iϕ
. (15)
One way of achieving this is to drive the cav-
ity with a laser field that experiences the phase
shift ϕ and to let it relax into its stationary state
(c.f. Fig. 1).
2. Afterwards, during the measurement stage, the
cavity is placed inside a quantum feedback loop
(c.f. Fig. 2). Whenever a spontaneously emitted
photon is detected, a laser pulse is applied. This
laser pulse displaces the coherent state inside the
resonator in a certain direction, which should be
independent of ϕ. For simplicity, we assume here
that the feedback pulse is approximately instanta-
neous.
Since the feedback laser does not experience the un-
known phase ϕ, it provides a reference frame. What
the proposed metrology scheme measures is the rel-
ative phase between the laser field applied during the
preparation stage and the quantum feedback laser. Al-
ternatively we could choose the preparation laser as
the reference frame, since this would not change the
Quantum-enhanced Metrology without Entanglement based on Optical Cavities with Feedback
225
Figure 1: Schematic view of the preparation stage. Its pur-
pose is the preparation of the coherent state |αi given in
Eq. (15), which depends on the unknown phase shift ϕ. One
way of achieving this is to drive a leaky optical cavity with
a laser field which experiences ϕ until the system reaches
its stationary state.
Figure 2: Schematic view of the measurement stage. Dur-
ing this stage, the cavity is placed inside a quantum feed-
back loop. The detection of photon emissions now triggers
a laser pulse that displaces the resonator field in a certain
direction. The relative phase between the laser field ap-
plied during the preparation stage and the quantum feed-
back laser can be deduced, for example, from second-order
photon correlation measurements with an accuracy that in-
creases rapidly in time.
dynamics of the system. Although doing so might be
less favourable in practical applications, let us assume
for the rest of the paper that this is the case for the sake
of convenience.
In standard approaches to quantum metrology
(Dowling and Seshadreesan, 2015; Giovannetti et al.,
2006), the relevant resource is the number of photons
experiencing the unknown phase shift ϕ. This is due
in part to the nature of interferometric experiments.
Usually N photons are passed through an interferom-
eter that contains the unknown phase before being
measured at the end. In the following we extract ϕ
from the photon statistics of the optical resonator dur-
ing the measurement stage. Hence in our scheme the
main resource is not the number of photons N pass-
ing through the setup but the number of observations
posed to deduce the photon statistics. This number is
directly proportional to the duration of the measure-
ment stage t which we can write as t = N t.
3.2 The Non-linear Conditional
Dynamics of an Optical Cavity with
Feedback
As we have seen in Section 2, the expectation values
of an open quantum system with spontaneous pho-
ton emission averaged over all possible quantum tra-
jectories behave as if they were generated by linear
operators. However, to enhance quantum metrology
beyond the standard quantum limit without using en-
tanglement, we require our setup to behave as if its
dynamics were generated non-linearly. One way of
achieving this is to deduce the unknown parameter ϕ
from a measurement signal M(t,ϕ) corresponding to
a pre-selected subset of quantum trajectories, which
involves quantum jumps. In the following we there-
fore have a closer look at the dynamics of subsets
of quantum trajectories of the experimental setup in
Fig. 2.
For an optical cavity with spontaneous photon
emission the Lindblad operator is L = c, where c is
the usual bosonic photon annihilation operator. This
leads to a peculiar effect. If prepared in a coherent
state |αi, as it is in general the case for a laser-driven
optical cavity (Clark et al., 2016b), the spontaneous
emission of a photon does not change the field inside
the resonator. In other words, there are no quantum
jumps in this case. The reason for this is that the co-
herent states are the eigenstates of the photon annihi-
lation operator c. To use the setup in Fig. 1 neverthe-
less for quantum-enhanced metrology, we either need
to prepare the cavity field in a non-coherent state or
we need to replace L by another Lindblad operator.
A straightforward way of changing L is to intro-
duce quantum feedback. This is why we propose to
place the cavity during the measurement stage into a
quantum feedback loop, as illustrated in Fig. 2. If
the feedback operation depends on the unknown pa-
rameter ϕ, then its introduction moreover results in an
effective ϕ-dependence of the jump operation. More-
over, changing the state of the cavity upon the detec-
tion of photon essentially generates temporal corre-
lations in the conditional dynamics of the quantum
system. These correlations mean that the informa-
tion corresponding to an individual quantum trajec-
tory with respect to two or more different time steps
is no longer additive, thereby allowing for scaling be-
yond the standard quantum limit.
Taking quantum feedback into account in the the-
oretical model, which we introduced in Section 2, is
straightforward. The no-photon time evolution re-
mains the same. However, the operator K
1
in Eq. (14)
now needs to be replaced by R(ϕ)K
1
where R(ϕ) is
a unitary operator (Clark et al., 2016b; Clark et al.,
PHOTOPTICS 2017 - 5th International Conference on Photonics, Optics and Laser Technology
226
2015; Wiseman and Milburn, 2010). In other words,
all we have to do is to replace the Lindblad operator
L by another operator L(ϕ),
L L(ϕ) = R(ϕ)L . (16)
For example, suppose every feedback operation R(ϕ)
displaces the coherent state inside the resonator by a
certain amount β(ϕ) such that
R(ϕ)|αi = |α + β(ϕ)i. (17)
Then the detection of a photon results indeed in an
effective quantum jump. An effective non-linearity
has been created which can be explored for quantum
metrology.
Breaking the standard quantum limit in Eq. (1)
when time is the resource which we want to constrain
requires that the dynamics of the cavity is very sensi-
tive to changes of the unknown parameter ϕ. In order
to be able to distinguish two parameters ϕ
1
and ϕ
2
,
the corresponding measurement signals M(t,ϕ
1
) and
M(t,ϕ
2
) need to evolve such that their distance grows
non-linearly in time. That this can be the case for the
experimental setup which we consider here is illus-
trated in Fig. 3. Suppose β = |α(0)|e
iϕ
and ϕ = π.
Then the detection of a photon at t = 0 prepares the
cavity in its vacuum state, i.e. in the coherent state |αi
with α = 0. Fig. 3 is a logarithmic plot of |α(t)| aver-
aged over all quantum trajectories with a photon de-
tection during the first time step of the measurement
stage (0,t) for different values ϕ which are all close
to π. All curves separate very quickly from the curve
corresponding to ϕ = π. Since the plot is logarithmic,
we see clearly that this happens in a highly non-linear
fashion with respect to time.
3.3 Second Order Photon Correlation
Functions
In order to obtain information about ϕ more effi-
ciently than suggested by the standard quantum limit,
we need to find a measurement signal M(t,ϕ), which
cannot be written as an ensemble average but depends
strongly on the appearance of quantum jumps. Taking
the discussion in the previous subsection into account,
we now have a closer look at the second-order photon
correlation function G
(2)
(t,t
0
), which is given by the
joint probability
G
(2)
(t,t
0
) I(t|t
0
)I(t
0
), (18)
where I(t|t
0
) denotes the probability for the detection
of a photon at a time t conditional on the detection of
a photon at t
0
. Second-order correlation functions are
usually normalised by the product of the photon emis-
sion rate at t
0
and at t. Doing so and dividing Eq. (18)
-2
0
2
0 0.2 0.4 0.6 0.8 1
Log (|α(t)|)
t
units of κ
1
ϕ = π +
5π
50
ϕ = π +
4π
50
ϕ = π +
3π
50
ϕ = π +
2π
50
ϕ = π +
π
50
Figure 3: Logarithmic plot of |α(t)| averaged over all the
possible quantum trajectories of the subensemsbles with a
photon detection and a subsequently applies feedback op-
eration R(ϕ) during the first time step of the measurement
stage (0,t) for five different values ϕ which are all close to
π. The figure is the result of a numerical simulation, which
averaged over 10
5
randomly generated quantum trajecto-
ries. Here we assume α(0) = 2 and β = |α(0)|e
iϕ
. More-
over, each time step t is much smaller than the cavity pho-
ton life time 1/κ.
by I(t
0
)I(t), we define the renormalised second order
photon correlation function, g
(2)
(t,t
0
), by
g
(2)
(t,t
0
)
I(t|t
0
)
I(t)
. (19)
Now g
(2)
(t,t
0
) depends no longer on the efficiency of
the detector shown in Fig. 2 and can be measured ac-
curately and relatively easily, even when using imper-
fect single-photon detectors.
Ref. (Clark et al., 2016b) uses measurements
of the second order photon correlation function
g
(2)
(t, 0), where t = N t denotes the length of the
measurement stage, to deduce information about an
unknown phase ϕ between two pathways of light.
This means, we propose to measure the joint proba-
bility of detecting a photon at the start of the mea-
surement stage and another photon at the end, after a
time t. For simplicity, we ignore photon emissions be-
tween these two points. Nevertheless, we found that
the uncertainty ∆ϕ scales as N
0.71
,
∆ϕ N
0.71
, (20)
when ϕ = π and β = |α|, which surpasses the stan-
dard quantum limit (Clark et al., 2016b). Although
no entanglement is used, the origin of the quantum
enhancement in Eq. (20) is still of a quantum nature.
The second-order correlation function has no classi-
cal analogue and its measurement requires the detec-
tion of individual photons, although unit detection ef-
ficiency is not required.
What might seem most surprising about Eq. (20)
is that the measurement uncertainty ∆ϕ decreases
Quantum-enhanced Metrology without Entanglement based on Optical Cavities with Feedback
227
rapidly, as t increases. The longer one waits, the
more information is unvealed about ϕ. This again is
due to an interesting property of optical cavities in-
side instantaneous quantum feedback loops. A more
detailed analysis of the dynamics of the experimental
setup in Fig. 2 shows that the field inside the resonator
never reaches a stationary state (Clark et al., 2016b;
Clark et al., 2016a) unless when being placed exactly
into its vacuum state. Whenever a photon is emit-
ted, a quantum feedback pulse occurs and increases
the number of photons inside the resonator. This in-
creases the probability for another photon emission
and so on. If the quantum feedback loop is truly
instantaneous, the mean number of photons inside
the cavity may easily diverge. Hence the longer one
waits, the more easily it becomes to distinguish these
two scenarios and to determine whether the system is
initially in its vacuum state or not. Here this question
is equivalent to asking whether ϕ = π or not.
Indeed the quantum enhanced-metrology scheme
in Ref. (Clark et al., 2016b) exploits the fact that
the cavity possesses two different types of dynam-
ics which separate in time. Similar effects have been
studied for example in Ref. (Macieszczak et al., 2016)
for quantum metrology applications. Due to the ef-
fective infinite dimensional Hilbert space of the field
inside an optical cavity, the divergence between both
types of dynamics may become arbitrarily large in
principle, thereby allowing for the scaling in Eq. (20)
to be preserved for an indefinite time, which is not the
case for the scenarios studied in Ref. (Macieszczak
et al., 2016).
4 TEMPORAL QUANTUM
CORRELATIONS AND
ENTANGLEMENT
Suppose subsequent generalised measurements are
performed on a two-dimensional quantum system
prepared in |ψi. Moreover we assume that these mea-
surements can be described by two Kraus operators
K
0
and K
1
of the form
K
i
= |
˜
ξ
i
ihξ
i
|, (21)
where |ξ
0
i and |ξ
1
i are two orthogonal states. How-
ever, no such constraint applies to |
˜
ξ
0
i and |
˜
ξ
1
i.
In case of two measurements, the quantum system
changes such that
|ψi →
K
0
|ψi
K
0
K
0
|ψi
K
1
K
0
|ψi
K
1
|ψi
K
0
K
1
|ψi
K
1
K
1
|ψi
(22)
up to normalisation factors. Moreover, suppose we
perform a single-shot measurement of K
0
and K
1
on
two quantum systems prepared in an effective state
|ψ
eff
i,
|ψ
eff
i =
p
00
|ξ
0
ξ
0
i+
p
01
|ξ
0
ξ
1
i
+
p
10
|ξ
1
ξ
0
i+
p
11
|ξ
1
ξ
1
i (23)
with the coefficients p
i j
equal to
p
i j
= kK
j
K
i
|ψik
2
. (24)
One can easily see that both measurements yield the
outcome i j with exactly the same probability. The
state |ψi and |ψ
eff
i have the same information con-
tent. However, |ψ
eff
i is in general an entangled state.
For example, if K
0
= |ξ
1
ihξ
0
| and K
1
= |ξ
0
ihξ
1
|, we
find that
|ψ
eff
i =
p
01
|ξ
0
ξ
1
i+
p
10
|ξ
1
ξ
0
i. (25)
Analogously, one can show that N successive mea-
surements on a single system are in general equivalent
to a single-shot measurement of N entangled quantum
systems.
The quantum-enhanced metrology scheme that we
propose in Ref. (Clark et al., 2016b) extracts informa-
tion about the unknown phase ϕ between two path-
ways of light by performing N successive measure-
ments on a single quantum system. This means, our
scheme is equivalent to performing single-shot mea-
surements on a combination of N entangled quantum
systems. It is therefore not surprising that our scheme
can be used to break the standard quantum limit, as
the system possesses correlations. These are corre-
lations between the system and its environment. It
is the measurements upon the environment, i.e. the
measurement of photon emission, that accesses these
correlations.
5 CONCLUSIONS
This paper emphasises that the environmental inter-
actions of open quantum systems with spontaneous
photon emission naturally result in non-linear con-
ditional dynamics which can be exploited for quan-
tum metrology and other applications. More con-
cretely, we propose to deduce an unknown parame-
ter ϕ by measuring an expectation value M(t, ϕ) av-
eraged over a subset of preselected quantum trajec-
tories instead of measuring ensemble averages. If
the signal M(t, ϕ) evolves with a non-linear genera-
tor, the accuracy of the measurement ∆ϕ can exceed
the scaling proposed by the standard quantum limit
(Clark et al., 2016c). As an example, we reviewed
a recent quantum-enhanced metrology scheme which
PHOTOPTICS 2017 - 5th International Conference on Photonics, Optics and Laser Technology
228
measures the phase shift between two pathways of
light using the open system dynamics of the elec-
tromagnetic field of an optical cavity inside a quan-
tum feedback loop (Clark et al., 2016b). This scheme
should be of immediate practical interest, since it re-
quires neither efficient optical non-linearities nor en-
tangled photons.
ACKNOWLEDGEMENTS
AS and AB acknowledge financial support from
the UK EPSRC-funded Oxford Quantum Technology
Hub for Networked Quantum Information Technolo-
gies NQIT. MMK acknowledges a postdoctoral re-
search fellowship funding from the Higher Education
Commission of the Government of Pakistan. GW ac-
knowledges financial support from the NSF of China
(Grant No. 11405026) and Government of China
through a CSC (Grant No. 201506625070 ).
REFERENCES
Beige, A., Braun, D., Tregenna, B., and Knight, P. L.
(2000). Quantum computing using dissipation to re-
main in a decoherence-free subspace. Phys. Rev. Lett.,
85:1762.
Blatt, R. and Zoller, P. (1988). Quantum jumps in atomic
systems. Eur. J. Phys., 9:250.
Boixo, S., Datta, A., Flammia, S. T., Shaji, A., Bagan, E.,
and Caves, C. M. (2008). Quantum-limited metrology
with product states. Phys. Rev. A, 77:012317.
Braun, D. and Martin, J. (2011). Heisenberg-limited sensi-
tivity with decoherence-enhanced measurements. Nat.
Commun., 2:223.
Carmicheal, H. (1993). An Open systems Approach
to Quantum Optics, Lecture Notes in Physics, vol-
ume 18. Springer-Verlag, Berlin.
Caves, C. M. (1981). Quantum-mechanical noise in an in-
terferometer. Phys. Rev. D, 23:1693.
Clark, L. A., Huang, W., Barlow, T. M., and Beige, A.
(2015). Hidden Quantum Markov Models and Open
Quantum Systems with Instantaneous Feedback, vol-
ume 14, page 143. Springer International Publishing
Switzerland.
Clark, L. A., Maybee, B., Torzewska, F., and Beige, A.
(2016a). Non-ergodiciity and non-classical correla-
tions through quantum feedback. arXiv: 1611.03716.
(submitted).
Clark, L. A., Stokes, A., and Beige, A. (2016b). Quantum-
enhanced metrology through the single mode coher-
ent states of optical cavities with quantum feedback.
Phys. Rev. A, 94:023840.
Clark, L. A., Stokes, A., Khan, M. M., and Beige, A.
(2016c). Quantum jump metrology. to be submitted.
Dalibard, J., Castin, Y., and Mølmer, K. (1992). Wave-
function approach to dissipative processes in quantum
optics. Phys. Rev. Lett., 68:580.
Dowling, J. P. and Seshadreesan, K. P. (2015). Quan-
tum Optical Technologies for Metrology, Sensing, and
Imaging. J. Lightwave Technol., 33:2359.
Giovannetti, V., Lloyd, S., and Maccone, L. (2006). Quan-
tum metrology. Phys. Rev. Lett., 96:010401.
Hegerfeldt, G. C. (1993). How to reset an atom after a pho-
ton detection. Applications to photon counting pro-
cesses. Phys. Rev. A, 47:449.
Kuhn, A., Hennrich, M., and Rempe, G. (2002). Deter-
ministic single-photon source for distributed quantum
networking. Phys. Rev. Lett., 89:067901.
Lim, Y. L., Beige, A., and Kwek, L. C. (2005). Repeat-
until-success linear optics distributed quantum com-
puting. Phys. Rev. Lett., 95:030505.
Luis, A. (2007). Quantum limits, nonseparable transforma-
tions, and nonlinear optics. Phys. Rev. A, 76:035801.
Macieszczak, K., Guta, M., Lesanovsky, I., and Garrahan,
J. P. (2016). Dynamical phase transitions as a re-
source for quantum enhanced metrology. Phys. Rev.
A, 93:022103.
McKeever, J., Boca, A., Boozer, A. D., Miller, R., Buck,
J. R., Kuzmich, A., and Kimble, H. J. (2004). De-
terministic Generation of Single Photons from One
Atom Trapped in a Cavity. Science, 303:1992.
Metz, J., Trupke, M., and Beige, A. (2006). Robust entan-
glement through macroscopic quantum jumps. Phys.
Rev. Lett., 97:040503.
Pearce, M. E., Mehringer, T., von Zanthier, J., and Kok, P.
(2015). Precision Estimation of Source Dimensions
from Higher-Order Intensity Correlations. Phys. Rev.
A, 92:043831.
Stokes, A., Kurcz, A., Spiller, T. P., and Beige, A. (2012).
Extending the validity range of quantum optical mas-
ter equations. Phys. Rev. A, 85:053805.
Wiseman, H. M. and Milburn, G. J. (2010). Quantum Mea-
surement and Control. Cambridge University Press.
Quantum-enhanced Metrology without Entanglement based on Optical Cavities with Feedback
229