Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Quantum Optomechanics

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Annalen der Physik, 16 October 2012

A short walk through quantum optomechanics

arXiv:1210.3619v1 [quant-ph] 12 Oct 2012

P. Meystre

This paper gives an brief review of the basic physics of

tion, control and characterization, including mechanical

quantum optomechanics and provides an overview of

squeezing and pulsed optomechanics. This is followed by

some of its recent developments and current areas of fo-

a discussion of the bottom-up approach that exploits ul-

cus. It first outlines the basic theory of cavity optomechan-

tracold atomic samples instead of nanoscale systems. It

ical cooling and gives a brief status report of the exper-

concludes with an outlook that concentrates largely on the

imental state-of-the-art. It then turns to the deep quan-

functionalization of quantum optomechanical systems and

tum regime of operation of optomechanical oscillators

their promise in metrology applications.

and cover selected aspects of quantum state prepara-

1 Introduction
Broadly speaking, quantum optomechanics provides a
universal tool to achieve the quantum control of mechanical motion [1]. It does that in devices spanning a vast
range of parameters, with mechanical frequencies from a
few Hertz to GHz, and with masses from 1020 g to several
kilos. At a fundamental level, it offers a route to determine and control the quantum state of truly macroscopic
objects and paves the way to experiments that may lead
to a more profound understanding of quantum mechanics; and from the point of view of applications, quantum
optomechanical techniques in both the optical and microwave regimes will provide motion and force detection
near the fundamental limit imposed by quantum mechanics.
While many of the underlying ideas of quantum optomechanics can be traced back to the study of gravitational wave detectors in the 1970s and 1980s [2, 3], the
spectacular developments of the last few years rely largely
on two additional developments: From the top down,
it is the availability of advanced micromechanical and
nanomechanical devices capable of probing extremely
tiny forces, often with spatial resolution at the atomic
scale. And from the bottom-up, we have gained a detailed
understanding of the mechanical effects of light and how
they can be exploited in laser trapping and cooling. These
developments open a path to the realization of macroscopic mechanical systems that operate deep in the quantum regime, with no significant thermal noise remaining.

Copyright line will be provided by the publisher

As a result, they offer both knowledge and control of the


quantum state of a macroscopic object, and increased
sensitivity, precision, and accuracy in the measurement
of feeble forces and fields.
It was Arthur Ashkin [4] who first suggested and
demonstrated that small dielectric balls can be accelerated and trapped using the radiation-pressure forces associated with focused laser beams. In later experiments
these particles, weighting on the order of a microgram,
were levitated against the Earth gravitational field. This advance led to the realization of optical tweezers, whose applications in biological science have become ubiquitous.
In parallel, the use of the strong enhancement provided
by resonant light scattering lead to the laser cooling of
ions and of neutral atoms by D. Wineland, T. W. Hnsch, S.
Chu, W. D. Phillips, C. Cohen-Tannoudji and many others,
resulting in a wealth of extraordinary developments [5]
culminating in 1995 with the invention of atomic BoseEinstein condensates [6, 7], and the subsequent explosion
in the study of quantum-degenerate atomic systems.
Non-resonant light-matter interactions present the
considerable advantage of being largely wavelength independent, providing one with the potential to achieve
optomechanical effects for a broad range of wavelengths
from the microwave to the optical regime. Resonant interactions, on the other hand, can result in a very large
enhancement of the interaction, but at the cost of being limited to narrow ranges of wavelengths. Cavity optomechanics exploits the best of both worlds by achieving
resonant enhancement through an engineered resonant

P. Meystre: A short walk through quantum optomechanics

structure rather than via the internal structure of materials. This could be for example an optical resonator with
a series of narrow resonances, or an electromagnetic resonator such as a superconducting LC circuit. Indeed, numerous designs can achieve optomechanical control via
radiation pressure effects in high-quality resonators. They
range from nanometer-sized devices of as little as 107
atoms to micromechanical structures of 1014 atoms and
to macroscopic centimeter-sized mirrors used in gravitational wave detectors.
That development first appeared at the horizon in the
1960s, but more so in the late 1970s and 1980s. It was
initially largely driven by the developments in optical
gravitational wave antennas spearheaded by V. Braginsky, K. Thorne, C. Caves, and others [2, 3, 8]. These antennas operate by coupling kilogram-size test masses to the
end-mirrors of a large path length optical interferometer.
Changes in the optical path length due to local changes
in the curvature of space-time modulate the frequency of
the cavity resonances and in turn, modulate the optical
transmission through the interferometer. It is in this context that researchers understood fundamental quantum
optical effects on mechanics and mechanical detection
such as the standard quantum limit, and how the basic
light-matter interaction can generate non-classical states
of light.
Braginsky and colleagues demonstrated cavity optomechanical effects with microwaves [9] as early as 1967.
In the optical regime, the first demonstration of these
effects was the radiation-pressure induced optical bistability in the transmission of a Fabry-Prot interferometer,
realized by Dorsel el al. in 1983 [10]. In addition to these
adiabatic effects, cooling or heating of the mechanical motion is also possible, due to the finite time delay between
the mechanical motion and the response of the intracavity
field, see section 2.2. The cooling effect was first observed
in the microwave domain by Blair et al. [11] in a Niobium
high-Q resonant mass gravitational radiation antenna,
and 10 years later in the optical domain in several laboratories around the world: first via feedback cooling of
a mechanical mirror by Cohadon et al. [12], followed by
photothermal cooling by Karrai and coworkers [13], and
shortly thereafter by radiation pressure cooling in several
groups [1419]. Also worth mentioning is that as early as
1998 Ritsch and coworkers proposed a related scheme to
cool atoms inside a cavity [20].
This paper reviews the basic physics of quantum optomechanics and gives a brief overview of some of its
recent developments and current areas of focus. Section
2 outlines the basic theory of cavity optomechanical cooling and sketches a brief status report of the experimental
state-of-the-art in ground state cooling of mechanical os-

Figure 1 (Color online) Generic cavity optomechanical system.


The cavity consists of a highly reflective fixed input mirror and
a small movable end mirror harmonically coupled to a support
that acts as a thermal reservoir.

cillators, a snapshot of a situation likely to be rapidly outdated. Of course ground state cooling is only the first step
in quantum optomechanics. Quantum state preparation,
control and characterization are the next challenges of
the field. Section 3 gives an overview of some of the major
trends in this area, and discusses topics of much current
interest such as the so-called strong-coupling regime, mechanical squeezing, and pulsed optomechanics. Section 4
discusses a complementary bottom-up approach that
exploits ultracold atomic samples instead of nanoscale
systems to study quantum optomechanical effects. Finally,
Section 5 is an outlook that concentrates largely on the
functionalization of quantum optomechanical systems
and their promise in metrology applications.

2 Basic theory
To describe the basic physics underlying the main aspects
of cavity optomechanics it is sufficient to consider an optically driven Fabry-Prot resonator with one end mirror
fixed -and effectively assumed to be infinitely massive,
and the other harmonically bound and allowed to oscillate under the action of radiation pressure from the intracavity light field of frequency L , see Fig. 1. Braginsky
recognized as early as 1967 [9] that as radiation pressure
drives the mirror, it changes the cavity length, and hence
the intracavity light field intensity and phase. This results
in two main effects: the optical spring effect, an optically induced change in the oscillation frequency of the
mirror that can produce a significant stiffening of its effective frequency; and optical damping, or cold damping,
whereby the optical field acts effectively as a viscous fluid
that can damp the mirror motion and cool its center-ofmass motion.

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

One can immediately understand how the optical


spring effect can result in a more quantum behavior of the
oscillator by recalling that in the high temperature limit
the mean number of phonons n m in the center-of-mass
motion of an oscillator of frequency m is given by
n m = k B T /m ,

(1)

where k B is Boltzmans constant and T the temperature.


For a given temperature, increasing m automatically
reduces n m , allowing one to approach the quantum
regime without having to reduce the temperature.
Cold damping, in contrast, does reduce the temperature of the oscillating mirror by opening up a dissipation
channel to a reservoir that is effectively at zero temperature. To see how this work, we first remark that in the
absence of optical field the oscillating mirror is dissipatively coupled to a thermal bath at temperature T . Its average center-of-mass energy, E , results from the balance
between dissipation and heating,
d E
= E + k B T,
dt

(2)

where is the intrinsic mechanical damping rate. When


an optical field is applied, an additional optomechanical
damping channel with damping rate opt comes into play
so that
d E
= E + k B T opt E .
dt

(3)

Importantly, that channel does not come with an additional (classical) thermal bath. Optical frequencies are
much higher than mechanical frequencies, so that the optical field is effectively coupled to a reservoir at zero temperature. In steady state Eq. (3) gives E = k B T /( + opt ),
or
Teff =

T
.
+ opt

(4)

This simple phenomenological classical picture predicts


that the fundamental limit of cooling is T = 0. A more
detailed quantum mechanical analysis does yield a fundamental limit given by quantum noise, see Section IIC, but
in practice, this is usually not a major limitation to cooling the mechanical mode arbitrarily close to the quantum
ground state, n m = 0.
More quantitatively, we consider a single mode of the
optical resonator of nominal frequency c and assume
that radiation pressure results in a displacement x(t ) of
the harmonically bound end-mirror, and consequently in
a change in the optical mode resonance frequency to
0c = c G x(t ),

Copyright line will be provided by the publisher

(5)

where
G = 0c /x.

(6)

For a single-mode Fabry-Prot resonator of length L this


becomes simply G = c /L.
Typical mechanical oscillator frequencies are in the
range of m /2 =10Hz to 109 Hz and the mechanical quality factors of the mirrors are in the range of perhaps Q m
103 107 , so that typically the damping rate = m /Q m
of the oscillating mirror is much slower than the damping rate of the intracavity field. One can then gain considerable intuition by first neglecting mirror damping altogether and assuming that its motion is approximately
harmonic,
x(t ) x 0 sin(m t ).

(7)

For a classical monochromatic pump of frequency L and


amplitude in the intracavity field obeys the equation of
motion
p
d (t )
= [i ( +G x(t )) /2] (t ) + in ,
dt

(8)

with the steady-state solution


p
in
=
.
i ( +G x) + /2

(9)

Here we have introduced the detuning


= L c

(10)

and is the intracavity field amplitude, normalized so


that

P
||2 =
( +G x)2 + (/2)2 L

=
(11)
( + c x/L)2 + (/2)2 L
where
P = L |in |2

(12)

is the input laser power driving the cavity mode. This normalization allows for an easy generalization to the case of
quantized fields, in which case will be interpreted as the
square
p root of the mean number of intracavity photons,
with a and a the bosonic annihilation and
= a a,
creation operators of the intracavity field. Note that |in |2
has then the units of photons per second.
For oscillation amplitudes x 0 small enough that G x 0 /m
1, it can be shown that the mirror oscillations simply result in the generation of two sidebands at frequencies

P. Meystre: A short walk through quantum optomechanics

L m , see e.g. Refs. [21, 22]. The time-dependent complex field amplitude (t ) then takes the approximate form
(t ) ' 0 (t ) + 1 (t ) with
p
in
0 (t ) '
,
(13)
i + /2
p

in
G x0
(14)
1 (t ) '
2 i + /2

e i m t
e +i m t

.
i ( + m ) + /2 i ( m ) + /2
The first sideband in Eq. (15) can be interpreted as an antiStokes line, with a resonance at L = c m , and the
second one is a Stokes line. An important feature of these
sidebands is that their amplitudes can be vastly different,
as they are determined by the cavity Lorentzian response
function evaluated at L m and L + m , respectively.

2.1 Static phenomena


Consider first a situation where the cavity damping rate
is much faster than all other characteristic times of the
system. One can then understand the mirror motion as resulting from the combined effects of the harmonic restoring force and the radiation pressure force F rp resulting
from an adiabatic elimination of the intracavity field, see
e.g. Ref. [23],
c
F rp = G||2 = ||2 ,
L

(15)

where ||2 is given by Eq. (11) and the second equality


holds for a simple Fabry-Prot. One can easily show that
the force F rp can be derived from the potential
||2
Vrp =
arctan [2( +G x)/] ,
2

(16)

the mirror of mass m being therefore subject to the total


potential
1
||2
arctan [2( +G x)/] .
V (x) = m2m x 2
2
2

(17)

The potential Vrp slightly displaces the equilibrium position of the mirror to a position x 0 6= 0, as would be intuitively expected, and also changes its spring constant
from its intrinsic value k = m2m to a new value

d 2 Vrp (x)
2
k rp = mm +
.
(18)

d x2
x=x 0

The second term in this expression is the static optical


spring effect. For realistic parameters it can increase the

stiffness of the mechanical system by orders of magnitude.


A third important static effect of radiation pressure is that
in general, there is a range of parameters for which the
potential V (x) can exhibit 3 extrema. Two of them correspond to stable local minima of V (x), and the third one to
an unstable maximum. This results in radiation pressure
induced optical bistability [10], an effect that is physically
similar to the more familiar form of bistability that can
occur in a Kerr nonlinear medium. The difference is that
in one case, it is the optical length of the resonator that
is changed by a Kerr nonlinearity, with its physical length
remaining unchanged, while in the other it is that physical
length that is intensity-dependent.

2.2 Effects of retardation


In general the optical field does not respond instantly to
the motion of the mechanical oscillator, therefore we need
to account for the effects of retardation as well. We proceed by assuming that he system is in equilibrium at some
mirror position x 0 with intracavity field 0 , taken to be
real without loss of generality, and consider the linearized
dynamics of small displacements x(t ) and (t ) from
that state under the effect of an external force F (t ),

+ x
+ 2 x = G0 + ,
x
m
= (i /2) + iG0 x.

(19)

These equations of motion can easily be solved, for instance in Fourier space, to give

iG0
() =
x()
(20)
+ ) + /2
i (
where
= +G x 0 ,

(21)

resulting in a modification of the radiation pressure force

F rp () = G0 () + () .
(22)
Together with Eq. (20) this expression shows that the mirror motion exerts a dynamical back-action on the radiation pressure force, which acquires both a real and an
imaginary component, the physical origin of the imaginary component being the delayed response of the intracavity field. As a result the intracavity power acquires a
component that oscillates out of phase with the mirror
motion, that is, with its velocity. It is through that friction
force that the optical field acts as a viscous field for the
mirror.
The net effect of the real and imaginary components of
F rp can be conveniently cast in terms of the back-action

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

frequency shift opt and a damping rate opt , see e.g.


Ref. [21, 22], with

opt =

G 2 20

+ m

+ m )2 + 2 /4
2mm (

+ m

m )2 + 2 /4
(
(23)

and

opt =

G 2 20

+ m )2 + 2 /4
2mm (

m )2 + 2 /4
(

.
(24)

m the first term in Eq. (24) domFor detunings


inates over the second term, and the dynamical backaction results in an increase in the mechanical damping
of the mechanical oscillator and cooling, see Eq. (4). It
is therefore the asymmetry between the response function of the Fabry-Prot at the frequencies of the two side
modes that is responsible for cooling or anti-damping
if one changes the sign of and uses a blue-detuned instead of a red-detuned driving field. In particular, in the
resolved sideband limit m we find
opt


2 G 2 20
,
mm

(25)

which can in principle be increased arbitrarily (within the


limits of validity of the model) by increasing the incident
optical power.
Together with Eq. (4) this analysis predicts that the
cooling of the center-of-mass motion of the mirror can
be arbitrarily close to Teff = 0, a consequence of the fact
that the optomechanical coupling between the intracavity field and the mirror results in the scattering of the
the driving field into an anti-Stokes line that is strongly
damped due to the high density of states at the cavity res m
onance. Conversely, for the opposite detuning
it is the Stokes line that is strongly damped, resulting in
anti-damping of the mirror motion. This can lead to parametric oscillations and dynamical instabilities, a situation
further discussed in section 3.5.
The quantum description of the next section will show
that cold damping and mirror cooling can also be interpreted in terms of of the annihilation of phonons from the
center-of-mass mode of oscillation when scattering the
driving laser field into the anti-Stokes sideband. Heating
can similarly be understood as resulting from the creation
of phonons associated with the scattering of the driving
field into the lower frequency Stokes side mode.

Copyright line will be provided by the publisher

Figure 2 (Color online) Schematic of sideband cooling: a coherent light field driving the resonator acquires frequency sidebands due to the mirror oscillations. The origin of the high
frequency sideband is the parametric transfer of phonons from
the mirror to the optical field and the lower sideband is due to
the reverse process, see section 2.2. Sideband cooling results
when the upper sideband frequency is resonant with the resonator. The solid black curve depicts the resonator transmission
near its mode of frequency c .

2.3 Quantum limit


The classical prediction that one can in principle reach an
arbitrarily large degree of cooling needs to be revised to
account for the effects of quantum and thermal noise. As
is well known, the open port of the interferometer used
to supply the optical drive of the oscillating mirror also
allows for the coupling of vacuum fluctuations into the
resonator, see e.g. Ref. [24]. This leads to a fundamental limit to the degree of cooling that can be achieved. A
proper quantum description of the system must account
for this effect as well as for the the bosonic nature of the
phonons.
Ignoring in a first step the important effects of fluctuations and dissipation, and in case a single optical mode
of the Fabry-Prot resonator and a single mode of oscillation of the suspended mirror need to be considered, the
optomechanical Hamiltonian is simply
a a +
H = (x)

p 2 1
+ m2m x 2 ,
2m 2

(26)

where a and a are bosonic annihilation and creation operators for the cavity mode of frequency , and p and x are
the momentum and position of the oscillating mirror of
mass m and frequency m . In reality, though, this Hamiltonian is more subtle than may appear at first. This is
because the mode frequency (q) depends on the length
of the resonator, which in turn depends on the intracavity
intensity. Stated differently, the boundary conditions for
the quantization of the light field are changing in time,
and do so in a fashion that depends on the state of that

P. Meystre: A short walk through quantum optomechanics

field and its history. The rigorous quantization of this system is a far-from-trivial problem, but for most cases of
interest in quantum optomechanics the situation is significantly simplified since the transit time c/2L of the light
field through the optical resonator is much faster than the
mechanical frequency m . The intracavity field therefore
learns about changes in its environment in times short
compared to 1/m . Under these conditions one can assume that the cavity frequency follows adiabatically any
change in resonator length,

1
nc

=
= c
c (1 x/L)
(27)
(x)

L + x
1 + x/L
where n is an integer that labels the mode of nominal
frequency c and L is the nominal resonator length (in the
absence of light.) In the classical limit we recover the result
G = c /L valid for a simple Fabry-Prot. The Hamiltonian
(26) reduces then to [2527]
p 2 1
(28)
H = c a a +
+ m2m x 2 G a a x
2m 2
p 2 1
b + b ).
+ m2m x 2 g 0 a a(
2m 2
In the second line we have used the familiar relationship
between the position operator x and the annihilation and
creation operators b and b of the mechanical oscillator,
= c a a +

x = x zpt (b + b )

(29)

with
s
x zpt =

.
2mm

(30)

We also introduced the optomechanical coupling frequency


g 0 = x zpfG = x zpf 0c /x,

(31)

which scales the optomechanical displacement to the


zero-point motion of the mechanical oscillator. The
Hamiltonian (28) is the starting point for most quantum
mechanical discussions of cavity optomechanics.
In order to establish the theoretical limit to cavity optomechanical cooling, it is necessary to expand the description provided by the Hamiltonian (28) to account for
the optical drive of the resonator, cavity damping, and the
mechanical damping of the oscillator. This analysis was
carried out in Refs. [2830]. The main message of these
papers is that at least for constant optomechanical coupling the best cooling can be achieved in the so-called resolved sideband limit, m , with the minimum mean
phonon number
n m =

0
T
opt n m
+ n m

+ opt

(32)

0
Here n m
is the mean steady-state number of phonons in
the absence of mechanical damping, given by the detailed
balance expression

0
+ m )2 + 2 /4
n m
+ 1 (
m
=
exp
,
(33)
0
m )2 + 2 /4
k B Teff
(
n m
T
n m
is the equilibrium phonon occupation determined by
the mechanical bath temperature, and

opt =

g 02 a a

(34)

2mm

+ m )2 + 2 /4
(

m )2 + 2 /4
(

T
For n m
0 one recovers the classical result of Eq. (4).
If the optical damping opt dominates, opt , though,
the mean phonon number is limited in the resolved sideband limit m to

2
0
,
(35)
n m =
4m

which shows that the ground state can be approached, but


not reached, in that case. As expected from the classical
considerations of the preceding section, this is the best
possible case. In practice, the theoretical limit (35) is difficult to reach due to technical noise issues including laser
noise [31, 32], clamping noise [33], etc. but the discussion
of these topics in beyond the scope of this brief review.
Remarkably though, these experimental challenges have
now being overcome in several experiments, see section
2.4. We also note that using pulsed optomechanical interactions may lead to improved cooling limits [34, 35]. We
return briefly to this point in section 3.7.
Importantly, we remark that optomechanical sideband
cooling is formally identical to the cooling of harmonically trapped ions, or more generally of any harmonically
trapped dipole, see Ref. [36] for a nice discussion of this
point. In the case of trapped ions, the resolved sideband
cooling limit was understood as early as 1975 [37, 38], and
the ground state cooling of trapped ions was first demonstrated over 20 years ago [39, 40]. As already mentioned,
the key new element contributed by cavity optomechanics is the use of engineered resonance-enhancing structures.

2.4 Experimental status


Following the pioneering work on gravitational wave antennas, advances in material science and nanofabrication in particular in microelectromechanical systems
(MEMS), nanoelectromechanical systems (NEMS), and

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

b
-29

nd=18

-13

10

n m=0.93

n d =4,500

n m=27

10
Figure
3 (Color online) Artist conception of the microwave
n =0.55
=11,000
n =71
n =22
optomechanical circuit
of Ref.n[42].
Capacitor element of the
LC10circuit
n =280 is formed
n =8.5 by a 15 micrometer diameter membrane
n =0.36
n =28,000
10
lithographically suspended 50 nanometers above a lower elecn =1,100
n =2.9
trode.
Insert: cut through the capacitor
showing the membrane
10
n =0.34
n =89,000
g
/
oscillations.
After
Ref.
[44].
n =4,500
n =0.93
n =0.42
n =180,000
4

S/Po (1/H z)

Sx (m /H z)

-30

-31

-14

8
6

-15

-32

10

O ccupancy

100

10

10.3

10.556
10.558
F requency (M H z)

10

10.4
10.5
10.6
F requency (M H z)

10.7

chanical ground state of center-of-mass motion. In one


case [42] the mechanical resonator was a suspended circular aluminum membrane tightly coupled to a superconducting lithographic microwave cavity. That cavity
was precooled to 20mK, corresponding to an initial occupation of 40 phonons and then further cooled by radiation pressure forces to an average phonon occupation of
n m 0.3. In contrast, Ref. [43] utilized an optomechanical structure with co-located photonic and phononic
band gaps in a suspended on-chip waveguide. The structure was precooled to 20K, corresponding to about 100
thermal quanta, and then cooled via radiation pressure
to n m 0.85. Shortly thereafter, that same group also
observed the motional sidebands generated on a second
probe laser by a mechanical resonator cooled optically
to near its vibrational ground state. They were able to detect the asymmetry in the sideband amplitudes between
up-converted and down-converted photons, a smoking
gun signature of the asymmetry between the quantum
processes of emission and absorption of phonons [45].

nm
nc

3 Beyond the ground state

3.1 Strong coupling regime

0.1
0

10

10

10

10

10

10

D rive P hotons, n d

Figure 4 (Color online) Phonon occupancy (blue) and intracavity photon occupancy (red) as a function of the drive photon
number. In this example sideband cooling reduces the thermal
occupancy of the mechanical mode from n m =40 into the quantum regime, reaching a minimum of n m =0.34 0.05. From
Ref. [42], with permission.

optical microcavities opened up the possibility to extend these ideas in many new directions, leading to the
demonstration of significant cooling in a broad variety
of systems from 2006 on, see Refs. [1518], with the first
demonstration of cooling in the resolved sideband regime
reported in Ref. [36].
More recently these efforts have culminated in the
cooling of the center-of-mass motion of at least three different micromechanical systems with a mean phonon
number within a fraction of a phonon of their ground
state of vibrational motion, n m < 1 [4143]. We postpone a discussion of Ref. [41] until the next section to
concentrate first on the two experiments [42, 43] that
utilized resolved sideband cooling to approach the me-

Copyright line will be provided by the publisher

Cooling mechanical resonators to their ground state of


motion is an essential first step in eliminating the thermal fluctuations that normally mask quantum features.
However, by itself that state is not particularly interesting, so the next challenge is to prepare, manipulate and
characterize quantum states of the mechanical resonator
required for a specific science or engineering goal. An
important first experimental step in that direction was
reported in Ref. [41]. In contrast to Refs. [42] and [43]
this experiment did not rely on radiation pressure cooling to achieve the motional ground state. Because of its
high frequency of about 6 GHZ, a conventional dilution
refrigerator that can reach temperatures of about 25 mK
was sufficient to cool it to n m < 0.07. A key point of the
experiment is that it succeeded in coupling an acoustic
resonator to a two-state system, or qubit, that could detect
the presence of a single mechanical phonon. This is analogous to protocols that have been developed over the years
in cavity quantum electrodynamics, see e.g. Ref. [46], with
the important distinction that photons are now replaced
by phonons.
The coupling between a bosonic field mode and one
or more two-state systems paves the way to a number
of approaches to prepare and to observe genuine quantum features such as the energy quantization of the res-

P. Meystre: A short walk through quantum optomechanics

onator, or to make controlled state manipulations at the


few phonons level. Many of those protocols have already
been developed in quantum optics and can be readily applied to phonon fields, at least in principle. In all cases
dissipation and decoherence must be reduced to a minimum, as they rapidly lead to the destruction of the most
salient quantum features of the state. In the experiment of
Ref. [41] decoherence was just weak enough to observe a
few coherent oscillations of a single quantum exchanged
between the qubit and the mechanical structure. As such
it can be considered as the first demonstration of the capability of coherent control of phonon fields in a micromechanical resonator.
Generally speaking, and in complete analogy with the
situation in quantum optics and in cavity QED, the control of the quantum state of a mechanical oscillator requires that one operates in the so-called strong coupling
regime, where the energy exchange between the mechanical object and the system to which it is coupled an
optical field mode, a qubit, an electron, etc. is not negatively affected by dissipation and decoherence. Section 2.3
showed that at the simplest level the optomechanical interaction takes the form (28),
H = c a a +

p 2 1
b + b ).
+ m2m x 2 g 0 a a(
2m 2

(36)

= 1, this interaction is
At the single photon level, a a
usually much too weak for its coherent nature to dominate over the incoherent dynamics for realizable levels of
1 the quantum
decay and decoherence. Since for a a
nature of the optical field normally rapidly decreases in
importance, it is therefore challenging to reach situations
where the full quantum nature of the interaction between
the photon and phonon fields is significant. There is a way
around this difficulty, though, the trade-off being that the
intrinsic nonlinear nature of the optomechanical interaction (36) disappears in the process to be replaced by a
linear effective interaction. As we shall see, this is not all
bad, as that effective interaction offers itself a number of
new opportunities.
Our starting point is the observation that strong intracavity optical fields can usually be decomposed as the
sum of a classical, or mean-field part and a small quantum mechanical component. In terms of the mean field

of the optical field mode = a


a + c

(37)

where c is again a photon annihilation operator. The optomechanical coupling term in the Hamiltonian (36) becomes then

Hint = g 0 n(b + b ) g c + c (b + b )
(38)

where we have introduced the optomechanical coupling


strength
p
g = g 0 n,
(39)
n = ||2 , and we have taken to be real for notational convenience. The first term in the Hamiltonian (38) describes
a simple Kerr effect, with a change in resonator length proportional to the classically intracavity intensity. This is the
term that leads to the radiation pressure induced optical
bistability observed e.g. in the experiments of Dorsel et
al [10].
In a frame rotating at the driving field frequency, the
cavity frequency and the mechanical frequency the second term in Eq. (38) can be reexpressed as
h
i
V = g b c e i (+m )t + h.c.
h
i
g b c e i (m )t + h.c.
(40)
This interaction describes the linear coupling between
the quantized component of the optical field and the mechanical oscillator. The coupling g is enhanced from the
single-photon optomechanical coupling frequency g 0 by
p
a factor n, which can be very substantial. Note however
that this enhancement comes at the cost of losing the non b + b ).
linear character of the original interaction g 0 a a(
That nonlinear character is at the origin of a number of
quantum effects that are expected to appear when the
radiation pressure of a single (or of very few) photons
displaces the mechanical oscillator by more than x zpf .
These include two-photon blockade as well as quantitative changes in the output spectrum and cavity response
of the optomechanical system, leading for example to the
possible generation of non-Gaussian steady states of the
oscillator [4749].
The linear coupling of Eq. (40) provides exciting opportunities as well, and these are significantly less challenging
to realize experimentally. On the red-detuned side of the
Fabry-Prot resonance, = m , we have after invoking
the rotating wave approximation

V ' g b c + h.c. ,
(41)
the so-called beam-splitter Hamiltonian of quantum optics. In contrast, in the blue-detuned side of the resonance,
= +m , we have

V ' g b c + h.c. ,
(42)
which describes the parametric amplification of the
phonon mode and the optical field.
This approach has enabled experiments to reach the
regime of strong phonon-photon optomechanical coupling in several micromechanical devices [5052]. A familiar characteristic of strongly coupled systems is the

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

occurrence of normal mode splitting. For the Hamiltonian (40) the normal mode frequencies are

1/2
q

2
1 2
=
+ 2m
2 2m + 4g 2 m
.
(43)
2

The beam-splitter Hamiltonian (41) describes the coherent exchange of cavity photons and mechanical phonons.
One of its remarkable properties is that it offers the potential to precisely transfer the quantum state of the mechanical oscillator to the electromagnetic field, and conversely.

Copyright line will be provided by the publisher

7.4740

nd=5x10

nd=5x10

p /2 (GHz)

0
7.4735

7.4730

nd=5x10

7.4725

Coupling rate, g / (Hz)

nd=5x10

nd=5x10

nd=5x10

g /

nd=5x10

This is seen easily by considering the Heisenberg equations of motion for the annihilation operators b and c in
the absence of decay,
) = b(0)
cos(g t ) + i c(0)
sin(g t ),
b(t
(44)

The optomechanical interaction g can easily be made


time dependent by pulsing the classical driving laser field
intensity, n n(t ). For an interaction time t int and a driving laser pulse intensity such that
t int

dt
0

10

10

10

Figure 5 (Color online) Normalized cavity transmission for


increasing resonator drive intensity n d = ||2 . For moderate
drive intensities the interference between the drive and probe
photons results in a narrow peak in the cavity spectrum, the
onset of electromechanically induced transparency. For higher
intensities the cavity resonance then splits into normal modes.
From Ref. [51], with permission.

g 0

10

10

7.4725
7.4730
7.4735
7.4740
Probe frequency, p /2 (GHz)

sin(g t ).
) = c(0)
cos(g t ) + i b(0)
c(t

3.2 State transfer

-1

nd=5x10

Probe transmission, |T|

The first demonstration of normal mode coupling in an


optomechanical situation was realized by Grblacher
and coworkers [50]. As pointed out by these authors the
optomechanical modes can be interpreted in a dressed
state approach as excitations of mechanical states that
are dressed by the cavity radiation field. Alternatively,
they can also be interpreted as optomechanical polariton modes. Teufel and coworkers [51] carried a series
of experiments in the strong coupling regime of quantum optomechanics. They measured the dressed cavity
states as a function of the pump-probe experiment where
the coupling (39) was controlled by a pump field and
the resonator transmission measured by a weak probe
field. Increasing the strength of g allowed them to monitor the change in cavity transmission as the strong coupling regime was reached, with an intermediate regime
where the interference between the pump and probe field
results in an effect analogous to electromagnetically induced transparenty [53, 54].
It should be emphasized that by itself, the observation
of normal mode splitting, which is both a classical and
quantum feature of coupled systems, does not prove the
existence of coherent exchange of excitations between the
mechanical and optical field modes. An important step
toward the demonstration of quantum coherent coupling
was recently achieved by Kippenberg and coworkers [52]
in a system where the optomechanical coupling is described by the beam splitter interaction (41). This experiment considered a micro-mechanical oscillator cooled
to a mean phonon number of the order of n m 1.7, and
in addition excited the system with a weak classical light
pulses to achieve coherent coupling between the optical
field and the micromechanical oscillator and the level of
less than one quantum on average. These results, while
still preliminary in many ways, open up a promising route
towards the use of mechanical oscillators as quantum
transducers, as well as in microwave-to-optical quantum
links as we now discuss.

n(t )t = /2

10

P. Meystre: A short walk through quantum optomechanics

int ) = c(0)

and c(t
int ) = i b(0),
we then have that b(t
indicative of a perfect state transfer between the optical and
phonon modes assuming of course that dissipation and
decoherence can be ignored during that time interval.

phase of the classical driving field, so that


p
p
g = g 0 n i g 0 n exp(i )

The interest in devices capable of high-fidelity state


transfer between optical and acoustical fields is largely
motivated by its potential for quantum information applications. This is because due to their potentially slow
decoherence rate motional states of mechanical systems
are well suited for information storage. However mechanics does not permit fast information transfer, while optical fields are ideal as information carriers, but are typically subject to fast decoherence that limits their interest for storage [55]. The coherent quantum mapping of
phonon fields to optical modes also promises to be useful in quantum sensing applications, by combining the
remarkable sensitivity of nanoscale cantilevers to feeble
forces and fields with reliable and high-efficiency optical detection schemes. And in addition to standard state
transfer between motional and optical states, phonon
fields could also serve as convenient transducers between
optical fields of different wavelengths, or between optical
and microwave fields.

i
h
V = i g b c g b c

The first theoretical proposal that analyzed a scheme


to transfer quantum states from a propagating light field
to the vibrational state of a movable mirror by exploiting radiation pressure effects is due to Jin Zhang and
coworkers [56]. This work was then expanded in several
directions, especially in the context of quantum optomechanics. For instance Tian and Wang [57] proposed an
optomechanical interface that converts quantum states
between optical field of distinct wavelengths through a
sequence of optomechanical /2 pulses. In another recent proposal, Didier et al. [58] considered exploiting the
beam splitter coupling of a mechanical oscillator and a
microwave resonator to measure and synthesize quantum
phonon states, and also to generate and detect entanglement between phonons and photons. They also proposed
generating the entanglement of two mechanical oscillators and its detection by the cavity field after entanglement swapping. The first experimental demonstration of
state transfer between a microwave field and a mechanical oscillator with amplitude at the single quantum level
was recently achieved by Palomaki et al. [59].

3.3 Two-mode squeezing


The Hamiltonian (42) is essentially the familiar two-mode
squeezing Hamiltonian of quantum optics. This becomes
readily apparent if one accounts for the (controllable)

10

and
(45)

with the associated evolution operator


S ab (t ) = exp[(g b c g b c )t ],

(46)

the well-known unitary two-mode squeezing operator. Introducing the generalized two-mode quadrature operator
X ab =

1
23/2

(c + c + b + b )

(47)

one finds that the variance of a system initially in a twomode vacuum state is given by [60]
(X )2 =

1 2|g |t
e
cos2 (/2) + e 2|g |t sin2 (/2) .
4

(48)

That same result also holds if the two modes are initially
in coherent states. For the choice = /2 one finds immediately that (X )2 can be well below the standard
quantum limit of 1/4, a signature of two-mode squeezing. Two-mode squeezed states are known to be entangled, indicating that this form of interaction can result in
quantum entanglement between the photon and phonon
modes. As such this configuration represents a useful resource for demonstrating fundamental quantum mechanical effects as well as for exploiting cavity optomechanical
devices in a quantum information context.
We note for completeness that in early work, Fabre
and coworkers [61], and independently Mancini and
Tombesi [62] exploited the analogy between the situation of an optical resonator and a cavity filled by a Kerr
medium to predict single mode squeezing of the reflected
optical field in situations where the motion of the mirror
is dominated by thermal fluctuations and can be treated
classically.

3.4 Squeezing via back-action evading


measurements
As shown by Braginsky et al. [63] and further analyzed by
Clerk et al. [64] it is possible to implement back-action
evading measurements of the membrane position when
driving it with an input field resonant with the cavity frequency c , but modulated at the mirror frequency m .
The mean-field amplitude of the intracavity field is then
p
(t ) = n cos(m t ),
(49)

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

where
p
in
=
.
i m + /2

(50)

Introducing the quadratures

1 i m t

X (t ) = p ce
+ c e i m t ,
2

i i m t

Y (t ) = p ce
c e i m t
2
of the motional mode, with [ X (t ), Y (t )] = i and
p

) = 2x zpt X (t ) cos m t + Y (t ) sin m t ,


x(t

(51)

(52)

and keeping as before only linear terms in the quantum


component of the field, the optomechanical interaction
Hamiltonian reduces then to
p

V = 2g X (1 + cos(2m t )) + Y sin(2m t ) (b + b )
(53)
p
where g = g 0 = g 0 n as before.
In a time-averaged sense the interaction Hamiltonian (53) reduces to
p
V 2g X (b + b )
(54)
and commutes with X , thus giving rise to the possibility of performing a back action evading measurement of
the X quadrature of mirror motion. This was verified experimentally in the classical regime by J. B. Hertzberg et
al. [68], but not yet in the quantum regime so far.
Since the interaction (54) is linear in X it is perhaps
less evident that it can also lead to quadrature squeezing.
This can be achieved by first performing a precise measurement of X , following which its quadrature can clearly
be below the standard quantum limit. Following that measurement the system would normally rapidly relax back to
a classical state, but by applying an appropriate feedback,
the measurement induced squeezing can be turned into
real squeezing. This is discussed in detail in Ref. [64].

3.5 Parametric instability


We have seen that for a driving laser red-detuned from the
cavity frequency c the upper sideband is resonantly enhanced by the cavity, which leads to preferred extraction
of mechanical energy, i.e. cavity cooling. For blue-detuned
light, in contrast, it is the lower sideband that is resonantly enhanced by the cavity, resulting in the preferred
deposition of mechanical energy, i.e. the optical amplification of mechanical motion. Invoking the rotating-wave

Copyright line will be provided by the publisher

approximation for +m one finds that this process


is described at the simplest level by the 2-mode squeezing interaction (42) instead of the beam-splitter Hamiltonian (41) of section 3.2.
In that regime the optomechanical system can display
dynamical instabilities. For appropriate parameters they
result in stable mechanical oscillations somewhat reminiscent of laser action, but for a phononic field [65, 66],
or even in unstable dynamics and chaos [67]. That this
can be the case is already apparent at the classical level
from the fact that opt can become negative for blue detuning, see Eq. (34). If the laser intensity is strong enough
that the total damping rate + opt is itself negative, then
any amplitude oscillation will grow exponentially until it
saturates due to the onset of nonlinear effects.
Following Ludwig et al. [69] we assume that the motion
of the cantilever is approximately sinusoidal,
x(t ) x + A cos(m t ),

(55)

with the average position x given by the radiation pressure


force,
x =

1
m2m

F rad =

G
m2m

|(t )|2

(56)

where (t )|2 is the intracavity light intensity and A is the


amplitude of oscillations of the mirror. Marquardt and
coworkers [71] showed that with Eq. (55), Eq. (8) yields for
the intracavity field
X
(t ) = e i (t ) n e i nm t ,
(57)
n

with
n =

max

J n (G A/m )
.
i nm / + i (G x )/ + 1/2

(58)

Here (t ) = (G A/m ) sin(m t ) and J n are Bessel functions of the first kind. The stability of the system can be
determined simply comparing the mechanical power P rad
due to radiation pressure to the dissipated power P fric due
to friction. When their ratio increases above unity the system starts to undergo self-induced oscillations [69].
Figure 6 is an example of a stability diagram determined from such an analysis. It shows the ratio P rad /P fric
as a function of the detuning and the square of the
(dimensionless) mechanical energy A 2 . Regions with
P rad /P fric > 1 are unstable, and the solid line defines an attractor where there is an exact power balance between amplification and damping. In general the parameter space
(, A) is characterized by the presence of a number of such
attractors. For relatively weak amplitudes A, nonlinear
effects tend to stabilize the oscillations of the cantilever,

11

P. Meystre: A short walk through quantum optomechanics

so that the optomechanical Hamiltonian becomes


2

1 c 2 2

x (b + b ) a a.
H = c a a + M b b +
2 x 2 zpt

(60)

In the rotating wave approximation this reduces to

2
2 c
H = c a a + M b b + x zpf
b
+
1/2
a a
b
x 2

(61)
= c a a + M b b + g 0(2) b b + 1/2 a a
where
2
g 0(2) x zpf

Figure 6 (Color online) Attractor diagram obtained from the


requirement that the optical power fed into mechanical oscillations is balanced by the power lost to friction. Adapted from
From Ref. [69], with permission.

leading to "laser-like" oscillations [65,66], but for larger oscillations amplitudes the system can become chaotic [67].
Quantum mechanically fluctuations are strongly amplified just below threshold, so that the attractor is no longer
sharp [70].

3.6 Quadratic coupling


So far we have considered geometries where the optomechanical coupling is linear in the oscillator displacement.
Other forms of coupling can however be considered, most
interestingly perhaps a coupling quadratic in the displacement. This can be realized in so-called membrane-inthe-middle geometries, as first demonstrated by J. Harris
and his collaborators at Yale [72, 73], see also Refs. [74, 75].
As implied by its name, this geometry involves an oscillating mechanical membrane placed inside a Fabry-Prot
with fixed end-mirrors.
An attractive feature of membrane-in-the-middle configurations is the ability to realize relatively easily either
linear or quadratic optomechanical couplings, depending
on the precise equilibrium position of the membrane. In
case the membrane is located at an extremum of 0c (x), so
that G = 0c /x = 0, see Eq. (6), we have to lowest order
1 2 c
0c (x) c +
2 x 2

12

(59)

2 c
.
x 2

(62)

Quadratic coupling opens up the way to a number of interesting possibilities, including the direct measurement
of energy eigenstates of the mechanical element, rather
than the position detection characteristic of linear coupling. J. Harris and coworkers estimate that it may be possible in the future to use this scheme to observe quantum
jumps of a mechanical system [73]. In another theoretical study, Nunnenkamp and coworkers [76] considered
optomechanical cooling and squeezing via quadratic optomechanical coupling. They showed that for high temperatures and weak coupling, the steady-state phonon
number distribution is nonthermal, and demonstrated
how to achieve mechanical squeezing by driving the cavity with two optical fields.
Another possibility offered by that geometry is to observe the quantum tunneling of an optomechanical system operating deep in the quantum regime through a
classically forbidden potential barrier. One proposed approach [77] relies on adiabatically raising a potential barrier, whose parameters can be widely tuned, at the location of a mechanical element. For the right choice of
parameters the optomechanical potential is a double-well
potential, and it is estimated that quantum tunneling between its wells can occur at rates several orders of magnitude larger than the decoherence rate of the mechanical
membrane. Besides tunneling, that scheme may also allow for the study of the quantum Zeno effect in a mechanical context and provide a comparatively simple scheme
for the preparation and characterization of non-classical
mechanical states of interest for quantum metrology and
sensing.

3.7 Pulsed optomechanics


So far we have largely limited our discussion to situations
where the optomechanical coupling is either constant or
slowly varying in time. One notable exception was the

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

quantum state transfer protocol outlined in Section III.B,


which requires that the interaction g (t ) be turned off at
the precise time when the state transfer has been completed. However there are a number of situations where
pulsed interactions are desirable, as already realized by
Braginsky [8, 78] in his proposal for a back-action evading
position measurement scheme. In a recent paper, Vanner
and coworkers [79] proposed to use a pulsed interaction
of duration short compared to the period of the mechanical oscillator to generate and fully reconstruct quantum
states of mechanical motion: As a result of the interaction the phase of the pulsed driving optical field becomes
correlated with the position of the mechanical oscillator,
while its intensity imparts it a momentum boost. A time
domain homodyne detection scheme can then be used to
measure the phase of the field emerging from the cavity,
thereby providing a measurement of the mechanical position. This scheme can also be used to achieve squeezing
and state purification of the mechanical resonator. It has
also recently been proposed that pulsed optomechanics
could be used, at least in principle, to surpass the limits
of conventional sideband cooling by using an optimized
sequence of driving optical pulses [34, 35].
In an intriguing potential application of pulsed optomechanics, Pikovski and colleagues [80] considered a
scheme to measure the canonical commutator of a massive mechanical oscillator and by doing so to detect possible commutator deformations due to quantum gravity: there are speculations that the existence of a minimum length scale where space-time is assumed to be
quantized, possibly of the order of the Planck length
L P = 1.6 1035 m, could result in such deformations. In
this proposal a sequence of optomechanical interactions
would be used to map the commutator of the mechanical
resonator onto an optical pulse. Remarkably the analysis
of Ref. [80] suggests that as a result Planck-scale physics
might be observable in a relatively mundane quantum
optics experiment.

4 Cold atoms
In a development complementary to the research on
nanoscale mechanical systems, recent quantum optomechanics experiments have also manipulated and controlled at the quantum level the center-of-mass degrees
of freedom of ultracold atomic ensembles [8184]. In the
following we restrict our discussion to the case of a neutral
atomic sample cooled well below its recoil temperature
and trapped inside a single-mode Fabry-Prot resonator.
This could be for example a nearly homogeneous and col-

Copyright line will be provided by the publisher

lisionless Bose-Einstein condensate (BEC) at T 0 or a


sample cooled near the vibrational ground state of one or
a few wells of the optical lattice formed by the optical field.
Side mode excitations of the condensate in the first case,
and the vibrational motion of thermal atoms in the second case, provide formal analogs of one or several moving
mirrors.
To see how this works we consider first a generic model
consisting of a BEC at T = 0 trapped inside a Fabry-Prot
cavity of length L and mode frequency c . The atoms of
mass M are driven by a pump laser of frequency L and
wave number k. When L is far detuned from the atomic
transition frequency a the excited electronic state of the
atoms can be adiabatically eliminated and the atoms interact dispersively with the cavity field. In the dipole and
rotating-wave approximations, the Hamiltonian describing the interaction between the atoms and the optical
field is
H = Hatom + Hfield ,

(63)

where
Hfield = c a a

(64)

and
Z
Hbec =

(x)
d x

p x2
2M

+ U0 cos2 (kx)a a (x).

(65)

Here (x)
is the bosonic Schrdinger field operator for
the atoms, a is the photon annihilation operator as before,
and the atoms interact with the light field via the familiar
off-resonant coupling
U0 = g R2 /(L a ),

(66)

where g R is the single-photon Rabi frequency. As always


H should be complemented by contributions describing
the external driving of the cavity field, dissipation and
collisions.
When the light field can be approximated as a plane
wave the atomic field operator can likewise be expanded
in terms of plane waves as
p X

(x)
= (1/ L) b k e i q x ,
(67)
q

where b q and b q are annihilation and creation operators for atomic bosons with the momentum k, satisfying
the bosonic commutation relations [b q , b q 0 ] = q,q 0 and
[b q , b q 0 ] = 0.
Consider for simplicity the case of scalar bosonic
atoms: In the absence of light field and at T = 0 the ground

13

P. Meystre: A short walk through quantum optomechanics

state of the sample would be a condensate with zero momentum,


|0 = (b 0 )N |0,

order in K z i one finds for L = c [85]


X
H (c + NU0 sin2 0 )a a + z b b i
i

(68)

"

but as a result of virtual transitions the atoms can acquire


a recoil momentum 2`k, where ` = 0, 1, 2, . . . In the
limit of low photon numbers it is sufficient to consider
the lowest diffraction order, ` = 1 and the atomic field
operator can be conveniently expressed in terms of a zeromomentum component and a sine mode,
b 0 0 (x) + b 2 2 (x)

(69)

where
0 (x) is the condensate wave function and 2 (x) =
p
2 cos(2kx). For very weak optical fields the occupation
of the sine mode remains much
p smaller the the zeromomentum mode, so that b 0 ' N and b 2 b 2 N . Substituting then Eq. (69) into the Hamiltonian (65) and in a
frame rotating at the pump laser frequency the Hamiltonian H becomes
h
i
H om,bec = 4rec b 2 b 2 + a a + g 2 (b 2 + b 2 ) ,
(70)

(71)

is the effective atom-field coupling constant, rec =


K 2 /2M the recoil frequency, and = c L +U0 N /2 is
an effective Stark-shifted detuning .
The reduced Hamiltonian (70) describes the coupling
of two oscillators, the cavity mode a and the momentum
b 2 +
side mode b 2 via the optomechanical coupling g 2 a a(
b 2 ). This shows that the condensate momentum side
mode behaves formally like a moving mirror driven by
the radiation pressure of the intracavity field, see Eq. (28)
for comparison.
A similar analogy can be established when considering a sample of ultracold atoms tightly confined to an
harmonic trap of frequency z centered at some location
z 0 along the resonator axis. The position of atom i is then
z i = z 0 + z i , and the vacuum Rabi frequency with which
it interacts with the field is
g R (z i ) = g R sin(0 + 2kz i ),

(72)

where 0 = kz 0 so that Eq. (66) becomes


U0 =

g (z i )2
.
L a

(73)

Summing over all atoms in the sample and expanding


then the far off-resonant atom-field interaction to lowest

14

#
X

kz i

(74)

where N is the number of atoms and the operator b i describes the annihilation of a phonon from the center-ofmass motion of atom i .
The second line of the Hamiltonian (74) describes the
optomechanical coupling of the intracavity optical field
to the collective atomic variable
X
k z i = kN Zcm
(75)
i

which is nothing but a measure of the normal mode of


P
the sample, its center of mass Zcm = N 1 i z i . For small
displacements that mode can be described as a harmonic
oscillator of frequency z and mass N M . In this picture,
the atom-field system is therefore modeled by the optomechanical Hamiltonian

Hom,at = 0c a a + z b b + g N (b + b )a a,

where
p
g 2 = (U0 /2) N /2

+ U0 sin(20 )a a

(76)

where b and b are bosonic annihilation and creation


operators for the center-of-mass
mode of motion of the
p
atomic ensemble, z zpf = /2N mz and
g N = NU0 (K z zpf ) sin2 0

(77)

p
with scales as N . Quantum optomechanics experiments
with non-degenerate ultracold atoms samples have so far
been carried out principally in the group of D. StamperKurn at UC Berkeley, while T. Esslinger and coworkers at
ETH Zrich have concentrated on the use of Bose condensates [86]. In a trailblazing experiment [85] Purdy et al positioned a sample of cold atoms with sub-wavelength accuracy in a Fabry-Prot cavity to demonstrate the tuning
from linear to quadratic optomechanical coupling from
the linear to the quadratic coupling regime. The Berkeley group also observed the measurement back-action
resulting from the quantum fluctuations of the optical
field by measuring the cavity-light-induced heating of the
atomic ensemble [81], the first observation of quantum
back-action on a macroscopic mechanical resonator at
the standard quantum limit. More recent work [87] detected the asymmetric coherent scattering of light by a
collective mode of motion of a trapped ultracold gas with
0.5 phonons of average excitation, a result that complements the work of Safavi-Naeini et al. [45] on the asymmetric absorption of light by a nanomechanical solid-state
resonator, see section 2.4.

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

A)

150

cillator substantially influences the cavity field. In subsequent work, the Bose condensate was irradiated from the
side of the optical resonator, resulting in the demonstration of a second-order quantum phase transition where
the condensed atoms enter a self-organized super-solid
phase, a process mathematically described by the Dicke
model of an ensemble of two-state systems coupled to
a single-mode electromagnetic field. In contrast to the
situation in the usual Dicke model, where the two states
of interest are two atomic electronic levels coupled by a
dipole optical transition, in the present case the relevant
states are two different momentum states coupled to the
cavity field mode [88, 89].

100

Photon spectrum (photons)

50
0
15
10
5
0
1.5
1
0.5
0
-130 -120 -110 -100 -90 90

100

Frequency (kHz)

110

120 130

10
Phonon occupation

B)

6
4
2

1
0.6
0.4
0.4 0.6

Cooperativity

10

20

Figure 7 (Color online) (a) Asymmetric optical scattering from


quantum collective motion, with the measured Stokes sidebands [left panels, (red) circles] and anti- Stokes sidebands
[right panels, (blue) circles] at various mean phonon numbers,
characterized by the so-called cooperatively coefficient C . From
top to bottom C = 9.6; 1.9; 0.4. (b) Measured phonon occupation
vs cooperativity. From Ref. [87], with permission.

Turning now to quantum degenerate gases, Brenneke


et al. [82] studied the dynamics of a Bose condensate of
87
Rb atoms trapped inside a high-finesse Fabry-Prot and
driven by a feeble optical field. This experiment demonstrated the optomechanical coupling of a collective density excitation of the condensate, showing that it behaves
precisely as a mechanical oscillator coupled to the cavity
field, in quantitative agreement with a cavity optomechanical model of Eq. (70). These authors also succeeded in
approaching the strong coupling regime of cavity optomechanics, where a single excitation of the mechanical os-

Copyright line will be provided by the publisher

5 Outlook Functionalization and hybrid


systems
The rapid progress witnessed by quantum optomechanics makes it increasingly realistic to consider the use of
mechanical systems operating in the quantum regime
to make precise and accurate measurements of feeble
forces and fields [90]. In many cases, these measurements
amount to the detection of exceedingly small displacements, and in that context the remarkable potential for
functionalization of opto and electromechanical devices
is particularly attractive. Their motional degree(s) of freedom can be coupled to a broad range of other physical
systems, including photons via radiation pressure from
a reflecting surface, spin(s) via coupling to a magnetic
material, electric charges via the interaction with a conducting surface, etc. In that way, the mechanical element
can serve as a universal transducer or intermediary that
enables the coupling between otherwise incompatible
systems. This potential for functionalization also suggests
that quantum optomechanical systems have the potential
to play an important role in classical and quantum information processing, where transduction between different
information carrying physical systems is crucial.
Much potential for the functionalization of optomechanical devices is offered by interfacing them with a single quantum object. This could be an atom or a molecule,
but also an artificial atom such as a nitrogen vacancy
center (NV center) in diamond [91], a superconducting
qubit [41, 92, 93] or a Bose-Einstein condensate [94]. Several theoretical proposals [91, 94100] and more recently
experimental realizations [101, 102] involving atomic systems have been reported. For example, a recent experiment [103] realized a hybrid optomechanical system by
coupling ultracold atoms trapped in an optical lattice to a
micromechanical membrane, their coupling being medi-

15

P. Meystre: A short walk through quantum optomechanics

ated by the light field. Both the effect of the membrane motion on the atoms and the back-action of the atomic motion on the membrane were observed. Singh and coworkers [104] considered a variation on that scheme where a
Bose condensate is trapped inside a Fabry-Prot with a
moveable end mirror driven by a feeble optical field. They
showed that under conditions where the optical field can
be adiabatically eliminated one can achieve high fidelity
quantum state transfer between a momentum side mode
of the condensate, see Eq. (69), and the oscillating endmirror.
Artificial atoms such as NV centers are of much interest for hybrid optomechanical systems [91] due to the
attractive combination of their optical and electronic spin
properties. Their ground state is a spin triplet [105] that
can be optically initialized, manipulated and read-out by
a combination of optical and microwave fields, and they
are characterized by remarkably long room-temperature
coherence times for solid-state systems. As such, they
offer much promise for applications e.g. in quantum information processing and ultrasensitive magnetometry,
where the spin is used as an atomic-sized magnetic sensor [106108]. In this context, a spin-oscillator system
of particular interest consists of a magnetized cantilever
coupled to the electronic spin of the NV center. A recent
experiment by Arcizet and colleagues demonstrated the
coupling of a nanomechanical oscillator to such a defect
in a diamond nanocrystal attached to its extremity [109].
In two further recent demonstrations of the potential
of hybrid optomechanical systems, a mechanical oscillator was used to achieve the coherent quantum control
of the spin of a single NV center [111], and the coherent
evolution of the spin of an NV center was coupled to the
motion of a magnetized mechanical resonator to sense
its motion with a precision below 6 picometers [110]. The
authors of that experiment comment that it may soon
become possible to detect the mechanical zero-point fluctuations of the oscillator.
More speculatively perhaps, micromechanical oscillators in the quantum regime offer a route toward new
tests of quantum theory at unprecedented sizes and mass
scales. For instance, spatial quantum superpositions of
massive objects could be used to probe various theories of decoherence and shed new light on the transition from quantum to classical behavior: In contrast to
the generally accepted view that it is technical issues
such as environmental decoherence that rapidly destroy
such superpositions in massive objects and establish
the transition from the quantum to the classical world,
some authors [112116] have proposed collapse models that are associated with more fundamental mechanisms and the appearance of new physical principles.

16

Bouwmeester [117] has pioneered the idea that quantum


optomechanics experiments may shed light on this issue
and on possible unconventional decoherence processes,
and in recent work Romero-Isart has analyzed the requirements to test some of these models and discussed the
feasibility of a quantum optomechanical implementation
using levitating dielectric nanospheres [118, 119].
Acknowledgements. This work is supported by the US National Science Foundation, by the DARPA ORCHID and
QuASAR programs through grants from AFOSR and ARO,
and by the US Army Research Office. We acknowledge enlightening discussions with numerous colleagues, in particular
M. Aspelmeyer, L. Buchmann, A. Clerk, H. Jing, M. Lukin, R.
Kanamoto, K. Schwab, H. Seok, S. Singh, S. Steinke, M. Vengalattore, E. M. Wright and K. Zhang.

References
[1] For recent reviews see T. J. Kippenberg and K. J. Vahala, Science 321, 1172 (2008); M. Aspelmeyer et al.,
J. Opt. Soc. Am. B 27, A189 (2010); F. Marquardt and
S. M. Girvin, Physics 2, 40 (2009); M. Aspelmeyer,
P. Meystre and K. Schwab, Physics Today 65, 29
(2012);D. M. Stamper-Kurn, arXiv:1204.4351, to appear in Cavity optomechanics, edited by M. Aspelmeyer, T. Kippenberg, and F. Marquardt, Springer
Verlag.
[2] V. B. Braginsky, Y. L. Vorontsov, and K. S. Thorne,
Science 209, 547 (1980); C. M. Caves et al., Rev. Mod.
Phys. 52, 341 (1980).
[3] P. Meystre and M. O. Scully, eds, Quantum Optics,
Experimental Gravitation, and Measurement Theory,
Plenum Press, New York and London (1983).
[4] A. Ashkin, Phys. Rev. Lett. 24, 156 (1970).
[5] For a review see e.g. H. J. Metcalf and P. van der
Straten, Laser Cooling and Trapping,, Springer Verlag, Berlin (1999).
[6] M.H. Anderson et al., Science 269, 198 (1995)
[7] K. B. Davis et al., Phys. Rev. Lett. 75, 3969 (1995).
[8] V. B. Braginsky and F. Ya. Khalili, Quantum Measurement, Cambridge University Press, Cambridge
(1992).
[9] V. B. Braginsky and A Manukin, Sov. Phys. JETP 25,
653 (1967).
[10] A. Dorsel et al., Phys. Rev. Lett. 51, 1550 (1983).
[11] D. G. Blair et al., Phys. Rev. Lett. 74, 1908 (1995).
[12] P. F. Cohadon, A. Heidmann and M. Pinard, Phys. Rev.
Lett. 83, 3174 (1999).
[13] C. Hhberger-Metzger and K. Karrai, Nature 432,
1002 (2004).
[14] S. Gigan et al., Nature 444, 67 (2006).
[15] O. Arcizet et al., Nature 444, 71 (2006).
[16] D. Kleckner et al., Nature 444, 75 (2006).
[17] A. Schliesser et al., Phys. Rev. Lett. 97, 243905 (2006).

Copyright line will be provided by the publisher

Ann. Phys. (Berlin) 0, No. 0 (2012)

[18] T. Corbitt et al., Phys. Rev. Lett 98, 150802 (1997);


Phys. Rev. Lett 99, 1608-1 (2007).
[19] A. Vinante et al., Phys. Rev. Lett. 101, 033601 (2008).
[20] G. Hechenblaikner et al, Phys. Rev. A 58, 3030 (1998).
[21] A. Schliesser, Cavity Optomechanics and Optical
Frequency Comb Generation with Silica WhisperingGallery-Mode Microresonators, PhD Thesis, LudwigMaximilians University Munich, Germany (2009).
[22] T. J. Kippenberg and K. J. Vahala, Optics Express 15,
17173 (2007).
[23] F. Marquardt, Quantum Optomechanics," Lecture
Notes, Les Houches School on "Quantum Machines,
to be published.
[24] C. M. Caves, Phys. Rev. Lett. 45, 75 (1980).
[25] C. K. Law, Phys. Rev. A 51, 2537 (1995).
[26] S. Mancini, V. I. Manko, and P. Tombesi, Phys. Rev. A
55, 3042 (1997).
[27] S. Bose, K. Jacobs and P. L. Knight, Phys. Rev. A 56,
4175 (1997).
[28] F. Marquardt et al., Phys. Rev. Lett. 99, 093902 (2007).
[29] I. Wilson-Rae et al., Phys. Rev. Lett. 99, 093901 (2007).
[30] See also F. Marquardt, A. Clerk and S. Girvin, J. Mod.
Optics 55, 3329 (2008) for a nice short review.
[31] P. Rabl et al., Phys. Rev. A 80, 063819 (2009).
[32] G. Phelps and P. Meystre, Phys. Rev. A 83, 063838
(2011).
[33] I. Wilson-Rae, Phys. Rev. B 77, 245418 (2008).
[34] X. Wang et al., Phys. Rev. Lett. 107, 177204 (2011).
[35] S. Machnes et al., ArXiv:1104.5448v2 (2012).
[36] A. Schliesser et al, Nature Phys. 4, 415 (2008).
[37] D. Wineland and H. Dehmelt, Bull. Am. Phys. Soc. 20,
637 (1975).
[38] W. Neuhauser et al, Phys. Rev. Lett. 41, 233 (1975).
[39] F. Diedrich et al., Phys. Rev. Lett. 62, 403 (1989).
[40] C. Monroe et al., Phys. Rev. Lett. 75, 4011 (1995).
[41] A. D. OConnell et al., Nature 464, 697 (2010).
[42] J. D. Teufel et al., Nature 475, 359 (2011).
[43] J. Chan et al., Nature 478, 89 (2011); A. H. SafaviNaeini et al., arXiv:1108.4680v1 (2011).
[44] P. Meystre, Science 333, 832 (2011).
[45] A. H. Safavi-Naeini et al., Phys. Rev. Lett. 108, 033602
(2012).
[46] S. Haroche and J. M. Raimond, Exploring the Quantum: Atoms, Cavities, and Photons, Oxford University Pres (2006).
[47] P. Rabl, Phys. Rev. Lett. 107, 063601 (2011).
[48] U. Akram et al., New J. Phys. 12, 083030 (2010).
[49] A. Nunnenkamp, K. Brkje, and S. M. Girvin, Phys.
Rev. Lett. 107, 063602 (2011).
[50] S. Grblacher et al., Nature 460, 724-727 (2009).
[51] J. Teufel et al., Nature 471, 204 (2011).
[52] E. Verhagen et al., Nature 482, 63 (2012).
[53] K. J. Boller, I. Imamoglu and S. E. Harris, Phys. Rev.
Lett. 66, 2593 (1991).
[54] S. Weis et al., Science 300, 1520 (2010).
[55] A. S. Parkins and H. J. Kimble, J. Opt. B: Quantum
Semiclass. Opt. 1, 496 (1999).

Copyright line will be provided by the publisher

[56] Jing Zhang, K. Pen and S. L. Braunstein, Phys. Rev. A


68, 013808 (2003).
[57] L. Tian and Hailin Wang, Phys. Rev. A 82, 053806
(2010).
[58] N. Didier et al., arXiv:1201.6293v1 (2012).
[59] T. A. Palomaki et al. ArXiv:1206.5562v1 (2012).
[60] R. Loudon and P. L. Knight, J. Modern Optics 34, 709
(1987).
[61] C. Fabre et al., Phys. Rev. A 49, 1337 (1994).
[62] S. Mancini and P. Tombesi, Phys. Rev. A 49, 4055
(1994).
[63] V. B. Braginsky, Y. I. Voronstsov and K. P. Thorne, Science 209, 547 (1980).
[64] A. A. Clerk, F. Marquardt and K. Jacobs, New J. Phys.
10, 1 (2008).
[65] K. Vahala et al., Nature Physics 5, 682 (2009).
[66] I. S. Grudinin et al., Phys. Rev. Lett. 104, 083901
(2010).
[67] T. Kippenberg and K. Vahala, Optics Express 15,
17172 (2007).
[68] J. B. Hertzberg et al., Nature Physics 6, 213 (2010).
[69] M. Ludwig, B. Kubala and F. Marquardt, New J. Phys.
10, 095013 (2008).
[70] Jiang Qian et al., ArXiv:1112.6200 (2011).
[71] F. Marquardt, J. G. E. Harris and S. M. Girvin, Phys.
Rev. Lett. 96, 103901 (2006).
[72] J. D. Thompson et al. Nature 452, 72 (2008).
[73] A. M. Jayich et al., New J. Phys. 10, 095008 (2008).
[74] P. Meystre et al., J. Opt. Soc. Am. B 2 1830 (1985).
[75] M. Bhattacharya, H. Uys and P Meystre, Phys. Rev. A
77 033819 (2008).
[76] A. Nunnenkamp et al., Phys. Rev. A 82, 021806(R)
(2010).
[77] L. Buchmann et al. Phys. Rev. Lett. 108, 210403
(2012).
[78] V. Braginsky, Y. I. Vorontsov and F. Ya. Khalili, JETP
Lett. 27, 276 (1978).
[79] M. R. Vanner et al., Proc. Nat. Acad. Sci. 108, 16182
(2011).
[80] I. Pikovski et al., Nature Phys. 8, 393 (2012).
[81] K. W. Murch et al., Nature Phys. 4, 561 (2008).
[82] F. Brennecke et al., Science 322, 235 (2008).
[83] R. Kanamoto and P. Meystre, Phys. Rev. Lett. 104,
063601 (2010).
[84] M. Schleier-Smith et al., Phys. Rev. Lett. 107, 143005
(2011).
[85] T. P. Purdy et al., Phys. Rev. Lett. 105, 133602 (2010).
[86] For a recent review of cavity optomechanics with
cold atoms see D. M. Stamper-Kurn, arXiv:1204.4351,
to appear in Cavity optomechanics, edited by M. Aspelmeyer, T. Kippenberg, and F. Marquardt, Springer
Verlag.
[87] N. Brahms et al., Phys. Rev. Lett. 108, 133601 (2012).
[88] K. Baumann et al., Nature 464, 1301 (2010).
[89] K. Baumann et al., Phys. Rev. Lett. 107, 140402 (2011).
[90] T. J. Kippenberg and K. J. Vahala, Science 321, 1172
(2008).
[91] P. Rabl et al., Phys. Rev. B 79, 041302(R) (2009).

17

P. Meystre: A short walk through quantum optomechanics

[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
[108]
[109]
[110]
[111]
[112]

[113]
[114]
[115]
[116]
[117]
[118]
[119]

18

A. D. Armour et al., Phys. Rev. Lett. 88, 148301 (2002).


M. D. LaHaye et al., Nature 459, 960 (2009).
P. Treutlein et al., Phys. Rev. Lett.99, 140403 (2007).
C. Genes, D. Vitali and P. Tombesi, Phys. Rev. A 77,
050307 (2008).
S. Singh et al., Phys. Rev. Lett. 101 263603 (2008).
A. A. Geraci and J. Kitching, Phys. Rev. A 80, 032317
(2009).
K. Hammerer et al., Phys. Rev. Lett. 103, 063005
(2009).
C. Genes, H. Ritsch and D. Vitali, Phys Rev A 80
061803(R) (2009).
K. Hammerer et al., Phys. Rev. A 82, 021803 (2010).
D. Hunger et al., Phys. Rev. Lett. 104, 143002 (2010).
P. Treutlein et al., Phys. Rev. Lett. 107, 223001 (2011).
S. Camerer et al. , Phys. Rev. Lett. 107, 223001
(2011).
S. Singh et al, arXiv:1202.6100.
M. W. Doherty et al., Phys. Rev. B 85, 205203 (2012).
J.M. Taylor et al., Nat. Phys. 4 810 (2008).
J. R. Maze et al., Nature 455, (2008).
G. Balasubramanian et al., Nature 455, 648 (2008).
O. Arcizet et al., Nature Physics 7, 879 (2011).
M. Kolkowitz et al., Science 335, 1603 (2012).
S. Hong et al., Nano Lett. 12, 3920 (2012).
G. C. Ghirardi, A. Rimini, and T. Weber, Phys. Rev.
D34, 470 (1986); G. C. Ghirardi, P. Pearle, and A. Rimini, Phys. Rev. A 42, 78 (1990).
A. Frenkel, Found. Phys. 20, 159 (1990).
R. Penrose, Gen. Relativ. Gravit. 28, 581 (1996).
L. Diosi, Phys. Lett. A 105, 199 (1984); ibid ,J. Phys. A:
Math. Theor. 40, 2989 (2007).
J. Ellis, N. E. Mavromatos, and D. V. Nanopoulos,
Phys. Lett. B 293, 37 (1992).
W. Marshall et al., Phys. Rev. Lett. 91, 130401 (2003).
O. Romero-Isart et al., New J. Phys. 12, 033015
(2010).
O. Romero-Isart, Phys. Rev. A 84, 052121 (2012).

Copyright line will be provided by the publisher

You might also like