Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Coherent feedback cooling of a nanomechanical membrane with atomic spins

Gian-Luca Schmid,∗ Chun Tat Ngai,∗ Maryse Ernzer, Manel


Bosch Aguilera, Thomas M. Karg,† and Philipp Treutlein‡
Department of Physics and Swiss Nanoscience Institute, University of Basel, 4056 Basel, Switzerland

Coherent feedback stabilises a system towards a target state without the need of a measurement,
thus avoiding the quantum backaction inherent to measurements. Here, we employ optical coherent
feedback to remotely cool a nanomechanical membrane using atomic spins as a controller. Direct
manipulation of the atoms allows us to tune from strong-coupling to an overdamped regime. Making
use of the full coherent control offered by our system, we perform spin-membrane state swaps
combined with stroboscopic spin pumping to cool the membrane in a room-temperature environment
to T = 216 mK (n̄m = 2.3 × 103 phonons) in 200 µs. We furthermore observe and study the effects of
delayed feedback on the cooling performance. Starting from a cryogenically pre-cooled membrane,
arXiv:2111.09802v1 [quant-ph] 18 Nov 2021

this method would enable cooling of the mechanical oscillator close to its quantum mechanical
ground state and the preparation of nonclassical states.

I. INTRODUCTION temperature bath. Coherent feedback is achieved by cou-


pling the two systems, thus reducing the noise in the
Hybrid quantum systems in which a mechanical os- mechanical system by transferring it to the spin, where
cillator is coupled to a spin are a promising platform it is dissipated. Additional coherent control of the spin
for fundamental quantum science as well as for quantum enhances the cooling performance.
sensing [1–3]. The interest in such systems derives from Hybrid systems combining atomic ensembles and me-
the fact that the spin – a genuinely quantum-mechanical chanical oscillators have been used for sympathetic cool-
object – can be used to control, read-out, and lend new ing by coupling the mechanical vibrations of a membrane
functionality to the much more macroscopic mechanical to the center-of-mass oscillation of cold atoms in an op-
device. Recently, different spin-mechanics interfaces have tical lattice [5, 6]. In these systems the atomic motion
been realized, involving the coupling of a mechanical os- was strongly damped and did not offer the possibility for
cillator to (pseudo-)spin systems such as atomic ensem- coherent control. Furthermore, optical traps for atoms
bles [4–9], quantum dots [10, 11], superconducting qubits cannot reach MHz trapping frequencies without intro-
[12–14], or impurity spins in solids [15–18], using light-, ducing substantial photon scattering and dissipation, re-
strain-, or magnetically-mediated interactions. stricting this cooling scheme to low-frequency mechanical
Coherent feedback is an intriguing concept that can be oscillators. In contrast, collective spin states of atomic
studied with such systems [19, 20]. In coherent feedback, ensembles offer long coherence times and wide magnetic
a quantum system is controlled through its interaction tuning of the spin precession frequency across the MHz
with another one, in such a way that quantum coherence range. Crucially, a versatile quantum toolbox exists that
is preserved. In contrast to measurement-based feedback provides sophisticated techniques for ground-state cool-
[21], coherent feedback does not rely on measurements, ing and quantum control [30, 31]. This makes it possible
thus avoiding the associated backaction and decoherence. to use the atomic spin as a coherent feedback controller,
Coherent feedback can under certain conditions outper- which can be employed to efficiently cool and control the
form measurement-based feedback in tasks such as cool- mechanical oscillator [32], e.g., via a state-swap [33].
ing of resonators [22, 23], and it has been implemented
in solid state systems to enhance the coherence time of a Here, we demonstrate coherent feedback control of a
qubit [24]. In optomechanical systems, it has been theo- nanomechanical membrane oscillator with the collective
retically studied as a way to generate large nonlinearities spin of an atomic ensemble and employ it to cool the
at the single photon level [25, 26], to enhance optome- membrane. For this, we exploit the coherent control of-
chanical cooling and state transfer [27], as well as for fered by our recently demonstrated spin-membrane inter-
entanglement generation [27–29]. face, where light mediates strong coupling between the
In the context of spin-mechanics interfaces, the me- two systems [8]. Using optical pumping on an internal
chanical oscillator can act as the system to be controlled, atomic transition we can modify the spin damping rate
i.e. the plant, which is coupled to a noisy thermal bath, and study the membrane cooling performance in differ-
and the spin system as the controller, coupled to a zero- ent regimes. We show that coherent state swaps alter-
nated with spin pumping pulses allow us to extract the
noise from the mechanical system in an efficient way, pro-
viding the largest cooling rate and reaching the phonon
∗ These authors contributed equally to this work steady-state faster than for continuous cooling. Finally,
† Current address: IBM Research Europe, Zurich, Säumerstrasse we study the effect of feedback delay onto the steady-
4, CH-8803 Rüschlikon, Switzerland state temperature of the membrane in the light-mediated
‡ philipp.treutlein@unibas.ch coupling between the mechanical and spin systems. Our
2

optical static magnetic field B0 = 2.8 G perpendicular to the


pumping
propagation direction of the coupling laser. The Lar-
mor frequency Ωs ∝ B0 is tuned into resonance with the
membrane frequency Ωm . The spin precession is mea-
sured after the first interaction with the coupling laser
by picking up a small fraction of the light (calibration
shown in Appendix A 1). The small-amplitude dynamics
of the transverse spin components can be described by a
harmonic oscillator of frequency Ωs using the Holstein-
Primakoff approximation [30].
A coupling laser beam interacts first with the spin,
then with the membrane, and once again with the spin, as
sketched in Fig. 1 and detailed in [8]. The coupling beam
with 1 mW optical power is slightly red-detuned with re-
spect to the membrane cavity and −2π × 40 GHz red-
detuned from the 87 Rb D2 -line. It cools the membrane
further to n̄m,bath = 2.0 × 104 phonons, which broadens
its linewidth to γm = 2π × 262 Hz. In presence of the
FIG. 1. Sketch of the light-mediated spin-membrane cou-
pling. Light interacts first with the spin, then with the mem-
coupling beam, the spin linewidth is γs = 2π × 2.2 kHz.
brane, and then again with the spin. The propagation of the In the first spin-light interaction, the Xs quadrature of
light leads to a feedback delay τ . On the way back from the the atomic spin is imprinted onto the coupling beam via
membrane to the spin, a π-phase is imprinted on the light, the Faraday interaction [30], resulting in a modulation of
rendering the spin-membrane interaction effectively Hamilto- the radiation-pressure force on the membrane. Likewise,
nian for zero-delay τ = 0. The systems can be approximated the membrane displacement Xm modulates the light re-
by harmonic oscillators of frequencies Ωm and Ωs with damp- flected from the cavity [35] which then creates a torque
ing rates γm and γs coupling them to a bath with n̄m,bath and on the spin in the second interaction. On the way back
n̄s,bath phonons, respectively. The oscillators are coupled at a from the membrane to the spin, the optical field carry-
rate g. The spin damping rate can be increased by applying ing the spin and membrane signals is phase-shifted by π
a σ− -polarized pumping laser.
such that the effective spin-membrane interaction is pre-
dominantly Hamiltonian and the backaction of the light
on the spin is suppressed [36]. Tracing out the light
observations agree well with a theoretical model. field and neglecting the propagation delay for the mo-
ment, the resonant part of the effective spin-membrane
interaction is described by a beam splitter Hamiltonian
II. SETUP HBS = ~g(b†s bm + b†m bs ), where bm (bs ) is the annihila-
tion operator of a membrane (spin) excitation and g is
Our hybrid system consists of a mechanical oscillator the effective spin-membrane coupling rate [8].
and a collective atomic spin coupled by laser light over
a distance of 1 meter in a loop geometry (Fig. 1). The
mechanical oscillator is the (2, 2) square drum mode of III. CONTINUOUS COOLING
a silicon-nitride membrane [34], which has a vibrational
frequency Ωm = 2π × 1.957 MHz and an intrinsic qual- Recently, we demonstrated strong coupling with this
ity factor Q = 1.4 × 106 . The membrane is placed in a spin-membrane interface, i.e. 2g > (γs + γm ) ≈ γs [8].
single-sided optical cavity of linewidth κ = 2π × 77 MHz, Strong coupling is manifested by the hybridization of
which enhances the optomechanical coupling to external the membrane and spin modes which leads to a normal
fields. The cavity is driven by an auxiliary laser beam mode splitting of 2g = 2π × 6.8 kHz in the spectrum as
(not shown in Fig. 1) that is red-detuned from the cavity shown in Fig. 2(b). In the time domain, strong coupling
resonance, providing some initial cavity optomechanical gives rise to state swaps between the spin and the mem-
cooling of the membrane to 2 × 105 phonons [35]. The re- brane at the coupling rate g. In Fig. 2(a) we show the
flection of this beam is used to stabilize the cavity length time evolution of the membrane occupation number after
and read out the membrane displacement via homodyne switching on the coupling beam. For 2g > γs , the ther-
detection (detailed in Appendix A 3). mally excited membrane swaps its state with the spin,
The collective spin is realised with an ensemble of which is initially prepared close to its ground-state, in
1.3 × 107 cold 87 Rb atoms confined in an optical dipole half a period Tπ = π/g of the energy exchange oscil-
trap. Strong coupling of the atomic ensemble to the lations. After another half period, the thermal state is
light is ensured by its large optical depth d0 ≈ 300. swapped back onto the membrane but the phonon num-
The atomic spins are optically pumped into the hyper- ber is reduced due to the damping that occurred in the
fine ground state |F = 2, mF = −2i with respect to a spin system, whose linewidth is larger than that of the
3

(a) (b)
γs /(2g) 101 γs /(2g)
0.3 0.3
105
1.0 0.4

S̄Xm Xm (1/Hz)
11.0 0.7
35.7 100 1.0
hb†m bm i

no spin 1.4
11.0
104 35.7
10−1

200 400 600 1950 1952 1955 1958 1960 1962 1965
time (µs) frequency (kHz)

FIG. 2. (a): Time traces of the membrane occupation number after turning on the coupling to the atoms. The different traces
show measurements with different spin damping rates γs . The dashed lines correspond to the simulation described in the text
based on Eqs. (1) and (2). The dotted line shows the membrane dynamics without atoms but with the coupling beam turned
on. (b): Power spectral density of the membrane displacement. The dashed lines show a global fit to the data with the initial
phonon occupation hb†i bi i(t = 0), Ωm , τ , g, and the detector shot noise level as global fit parameters and Ωs and γs as individual
fit parameters. All other parameters were taken from independent calibrations. In (a) and (b), solid lines correspond to the
mean and shaded areas to the standard deviation of 355 measurements.

membrane. The oscillations dephase after approximately the membrane much faster than in the continuous cool-
1 ms and a steady state with a membrane occupation of ing case discussed above. In Fig. 3 we show a compari-
n̄m,ss ≈ 2.3 × 103 phonons is reached, corresponding to a son between stroboscopic and continuous cooling, where
temperature decrease by two orders of magnitude com- time traces for (a) the membrane occupation number
pared to the initial state. In this process the membrane and (b) the spin occupation number are shown. In
is predominantly cooled via its coupling to the cold and the stroboscopic sequence we perform a coherent π-pulse
damped spin, reaching a temperature one order of mag- (Tpulse = 100 µs, γs = 0.6g) to swap membrane and spin
nitude lower than in the presence of the optomechanical states. Afterwards, we apply an optical pumping pulse of
cooling beams alone. duration Tpump = 10 µs which increases the spin damp-
We now study the effect of increasing the spin damp- ing rate to γs ≈ 60g and depletes the spin occupation
ing rate γs on the coupled dynamics. To increase γs on a timescale much shorter than the state swap (gray
we apply a σ− -polarized pump laser along the polariza- pulses in Fig. 3(b)). During the pumping pulse the cou-
tion axis of the spin (calibration in Appendix A 2). As pling is kept on. Since the spin is reinitialised close to
can be seen in Fig. 2(a), increasing γs first enhances the the ground state, the next coherent state swap does not
membrane cooling, until the overdamped regime γs  2g transfer thermal energy back to the membrane but only
is reached where the membrane couples incoherently to cools it further. It takes two to three such iterations
a quasi-continuum of cold spin fluctuations. The mem- of a coherent π-pulse followed by a spin pumping pulse
brane decay is then governed by Fermi’s golden rule, with to reach the steady state (see Fig. 3). Using this sim-
the occupation number decreasing at the sympathetic ple sequence, we can reach the membrane steady state
cooling rate γsym ≈ 4g 2 /γs , i.e. the cooling becomes less temperature of 216 mK (n̄m,ss = 2.3 × 103 phonons) in
effective as γs is increased further. In this weak-coupling around 200 µs, approximately a factor of two faster than
regime, the modes decouple and the membrane spectrum for continuous cooling. This exemplarily shows the ad-
shows a single Lorentzian peak, broadened by the inter- vantage of a coherent feedback controller, which enables
action with the spin, see Fig. 2(b). faster cooling than if the membrane is coupled with a
similar rate to an incoherent, overdamped system.

IV. STROBOSCOPIC COOLING

Previous experiments, which coupled a membrane to


the motion of cold atoms [5, 6], lacked both strong cou-
pling and coherent control over the atoms. In contrast, V. THEORETICAL MODEL
our strongly coupled spin-membrane system allows us to
implement more elaborate coherent control schemes. In
particular, we can combine strong coupling and strong Further insight into the dynamics is gained by solving
spin damping in a stroboscopic fashion in order to cool the equations of motion for the coupled spin-membrane
4

(a) (a)
Continuous Cooling
105 Stroboscopic Cooling 104

hb†m bm iss
hb†m bm i

104

100 101 102


γs /(2g)
(b) 105
8
(b) 10
Sim, τ = 0 ns
104 Sim, τ = 15 ns
hb†s bs i

hb†m bm iss
106 Large γs
Data
103
104
102
100 200 300 400 500 600 −7.5 −5.0 −2.5 0.0 2.5 5.0 7.5
time (µs) δ/(2π) (kHz)

FIG. 3. (a) Membrane and (b) spin occupation numbers for FIG. 4. Steady state occupation of the membrane as a func-
continuous cooling at γs = 2g and stroboscopic cooling at tion of (a) spin damping rate γs (at resonance, δ = 0) and
γs = 0.6g. The gray shaded areas indicate the spin pump- (b) spin-membrane detuning δ = Ωs − Ωm at γs = 0.6g. The
ing pulses (where γs ≈ 60g). Solid lines and shaded areas solid (dashed) blue line shows the result of the simulation
correspond to the mean and standard deviation of 70 mea- with (without) delay. In (a), the red dashed-dotted line in-
surements and dashed lines correspond to a simulation. dicates the steady-state number given by the rate in Eq. (3)
with τ = 15 ns. The red shaded area shows the region for
which the dynamics is found to be unstable using the Routh-
system [8], Hurwitz criterion. For this measurement, n̄m,bath ≈ 4.0 × 104
phonons and γm = 2π×94 Hz (independently calibrated with-
out atoms). The data points with error bars correspond to the
Ẍm + γm Ẋm + Ω2m Xm = −2gΩm Xs (t − τ ) + Fm , (1)
mean and the standard deviation of steady state occupations
Ẍs + γs Ẋs + Ω2s Xs = −2gΩs Xm (t − τ ) + Fs , (2) of 20 (3) experimental realisations in (a) ((b)).

where terms on the left-hand-side describe the internal


dynamics of the damped oscillators and the first term on For the stroboscopic cooling measurements, we took
the right-hand-side describes the state swap dynamics the fit parameters from the continuous cooling measure-
including a propagation delay τ between the spin and ment and ran the simulation with a time dependent spin
the membrane. We included the generalized Langevin damping rate which was taken to be γs = 0.6g during
forces Fm and Fs that capture stochastic force terms the state swaps and γs = 60g during the pumping pulses.
due to quantum fluctuations, thermal and measurement The fit is shown for membrane and spin in Fig. 3 as a
backaction noise (detailed in Appendix B 1). dashed line. The good agreement between fit and data
We used the following procedures to simulate our ex- shows that our model includes all the relevant factors
perimental results: for the continuous cooling measure- which govern the coupled dynamics.
ments, we first fitted the spectra for different γs in
Fig. 2(b) globally using a coupled-mode model (fit func-
tion given in Appendix B 2). From this fit, the extracted
τ and Ωm were used as the input parameters for the sim- VI. DELAYED FEEDBACK
ulation. We adapted the technique described in [37] to
numerically solve the equations of motion (1) and (2) Our hybrid spin-membrane system constitutes a coher-
and compare the solution to our data (more details are ent feedback network [23], in which delayed feedback can
given in Appendix B 1). To generate each time trace in give rise to instabilities [38–40]. In our experiment, such
Fig. 2(a) (dashed lines) we fitted the numerical solution instabilities show up as a spontaneous coupled oscillation
to our data with only γs and Ωs as free parameters. The of spin and membrane, which we observe for certain val-
fit results show a systematic shift of Ωs with increasing ues of the spin-membrane detuning δ = Ωs − Ωm . Even
spin pumping power, likely due to the light shift induced at resonance, we have to include the feedback delay to
by the circularly polarised pumping laser (Fig. 7), and predict the experimentally measured steady state occu-
γs was observed to be larger than in the independent pation of the membrane accurately. In Fig. 4 we plot
calibration of Appendix A 2. the measured and simulated occupation numbers of the
5

membrane in steady state as a function of γs [Fig. 4(a)]


and δ [Fig. 4(b)]. At resonance and for Ωm τ  1 (as in
our system), the effect of the feedback delay is most ap-
parent in the limit of small γs . The model without delay
(light-blue dashed line) predicts a significantly smaller
occupation number compared to both what we observe
in experiments and what is predicted by our model in-
cluding the feedback delay (blue solid line). In the large
γs limit, the sympathetic cooling rate is modified to

4g 2
γsym ≈ [γs cos(2Ωm τ ) + 2δ sin(2Ωm τ )] (3)
4δ 2+ γs2

FIG. 5. Simulated steady state occupation of membranes for


(see Appendix B 3 for derivation). In this limit, varying cryostat temperature and different mechanical Q fac-
the steady state occupation is given asymptotically by tors. Here, γs = 2g, δ = 0 and τ = 15 ns. The insets show the
hb†m bm iss = n̄m,bath γm /(γm + γsym ), shown as the red current membrane with phononic shield used in these exper-
dashed-dotted line in Fig. 4(a). The theory of coupled iments and a soft-clamped membrane for which Q ≈ 5 × 107 .
oscillators without delay predicts optimal sympathetic
cooling at the critical damping of γs = 2g (faded vertical
dotted line in Fig. 4). Including the feedback delay in the VII. DISCUSSION
model, the minimal occupation number shifts to larger γs
(dark vertical dotted line), because the self-oscillations In our experiment, the cooling rate of the membrane
have to be compensated by a higher spin damping rate. due to its coupling to the spin exceeds the cavity-
The experimental data confirms this theoretical predic- optomechanical cooling rate by more than one order of
tion. magnitude. The lowest achievable phonon occupation of
the membrane is thus given by the competition of cooling
Furthermore, we find that the presence of delay lifts
the membrane with the spin and heating due to its cou-
the symmetry in δ, as inferred theoretically from Eq. (3)
pling to the room-temperature environment. In Fig. 5 we
for large γs and shown both experimentally and theo-
show the expected membrane steady state occupation for
retically in Fig. 4(b) for small γs = 0.6g. We see that
varying environment temperature and two different mem-
the minimal steady state occupation of the membrane is
brane designs. In this calculation we include the cavity-
obtained for positive detuning δ, i.e. Ωs > Ωm , which
optomechanical cooling of the membrane (which has a
is true in general for a feedback system with a delay
negligible effect), the light-mediated coupling to the spin
of τ < π/(2Ωm ). For large enough negative δ, we ob-
including backaction of the light, as well as thermal and
serve that the coupling drives the system into limit cycle
quantum mechanical ground state fluctuations of both
oscillations, see Fig. 4(b). With our model we can at-
systems. The higher quality factors Q > 5 × 107 of soft-
tribute these self-oscillations to the feedback delay. In
clamped membranes [42, 43] would reduce the thermal
this self-driven regime, the resulting membrane occupa-
decoherence rate by a factor 25 and allow us to prepare
tion of 6.8 × 107 exceeds the spin length by around a
the mechanical oscillator close to its ground state in a 4 K
factor of three. The emergence of such instabilities can
environment. These technical improvements would real-
be characterised using the Routh-Hurwitz stability crite-
ize a mechanical oscillator whose phonon occupation is
rion [41], which indicates whether the real part of one of
limited by quantum backaction instead of thermal noise.
the normal modes of the system reverses its sign (shown
While in the current coupling scheme the double pass
in Appendix B 4). In Fig. 4 we indicate such unstable
eliminates backaction on the atomic spin, a large mem-
regions for our coupled system by a shaded area. Our
brane quantum cooperativity Cm > 1 would favor a
calculations show that the precise value of δ at which
double pass scheme with coherent cancellation of quan-
the driving due to the loop delay exceeds the damping
tum backaction on the membrane. This would lead to
of the coupled system depends on γs . Even at resonance
a higher quantum cooperativity for the spin-membrane
[Fig. 4(a)] self-oscillations are predicted for small enough
coupling [36]. Further, the feedback control of the mem-
γs .
brane could be improved by increasing the quantum co-
The propagation delay is an interesting tuning knob operativity of the spin system. This involves gaining a
for coherent feedback experiments, which gives access to better understanding of the spin decoherence sources and
Hamiltonian and dissipative dynamics: We can induce achieving a larger spin-light coupling rate.
self-oscillations of the system, tune the dependence of In this work we implemented a relatively simple coher-
the steady state on system parameters such as damping ent feedback sequence based on coherent state swaps of
rate and detuning, or even render the delay negligible by pulse area π interleaved with short spin pumping pulses.
tuning 2Ωm τ to a multiple of 2π. In the future, it would be interesting to explore more
6

elaborate feedback sequences to optimize the cooling in Appendix A: Calibrations


a specific situation. For example, the duty cycle of the
stroboscopic cooling sequence could be changed over time 1. Calibration of the spin signal
to cool a mechanical oscillator with a high initial occupa-
tion that exceeds the spin length. Initially, short coupling The spin occupation was calibrated by using the off-
pulses of pulse area  π could remove excitations with- resonant Faraday interaction [30] between the atoms and
out saturating the spin, and once the phonon number is the light. In contrast to the looped coupling scheme in
sufficiently reduced, the pulse area could be increased to which light interacts twice with the spin, the calibration
minimize the final temperature. is performed by measuring the light directly after the
Our coherent feedback cooling scheme is a rather gen- first interaction with the spin. Hence, the light does not
eral technique that can be applied to any physical system interact with the membrane nor with the atoms a second
with a strong light-matter interface. This includes cavity time.
optomechanical systems or mechanical oscillators with- For a single spin-light interaction, the Hamiltonian of
out an optical cavity. Moreover, similar cooling schemes the Faraday interaction is given by [44]
could be implemented in the microwave domain with elec-
tromechanical oscillators [14] coupled to solid-state spin Hint = ~α1 Sz Fz (A1)
systems. The macroscopic distance between the feed-
back controller and the target system enables modular where α1 is the vector polarizability of the atoms, Sz is
control schemes in analogy to classical feedback in elec- the circularly polarized component of the Stokes vector
trical engineering. This opens up the new possibility to of the light, and Fz is the collective spin component along
use coherent feedback control in quantum networks. the propagation direction of the probe laser. The input-
The coherent control and bidirectional Hamiltonian output relation for the Sy Stokes vector component of
coupling employed in this work pave the way towards the probe light yields
more elaborate quantum protocols such as the genera- Sy(out) = Sy(in) + α1 Sx(in) Fz . (A2)
tion of non-classical mechanical states via state swaps
[33] as well as further studies of coherent feedback in the For the experiments in this paper, the probe laser was
quantum regime [19–22]. linearly polarized with an angle of 55◦ with respect
to the magnetic field in order to minimise frequency
shifts due to the tensor interaction with the light [44],
ACKNOWLEDGMENTS which otherwise would give rise to inhomogeneous
broadening of the spin. In the following we define Sx as
We thank Christoph Bruder for a careful reading of the difference between the flux of light linearly polarized
the manuscript. This work was supported by the project at 55◦ and the flux in the orthogonal linear polarization,
“Modular mechanical-atomic quantum systems” (MOD- such that we can approximate the Sx operator as a
(in) (out)
ULAR) of the European Research Council (ERC) and classical quantity Sx ≈ Sx ≈ hSx i and define the
by the Swiss Nanoscience Institute (SNI). MBA acknowl- (out)
Faraday angle θF = Sy /(2hSx i) = α1 Fz /2. For the
edges funding from the European Union’s Horizon 2020 calibration measurement, we slowly rotate the spin to
research and innovation programme under the Marie align it to the propagation direction of light. Thus,
Sklodowska-Curie grant agreement N°101023088. the collective atomic spin points along the z-axis, and
the Fz component of the spin can be approximated by
Fz ≈ hFz i = F Na = 2Na where the number of atoms
in the dipole trap Na was measured independently by
(out)
absorption imaging. The Sy Stokes vector component
of the out-going field was determined by a polarization
homodyne measurement. Measuring hSx i independently
allows us to determine the Faraday angle (shown in
Fig. 6 for different atom numbers). By knowing the
number of atoms from absorption imaging and the Fara-
day angle from polarization homodyne measurement,
the ensemble-averaged vector polarizability α1 = θF /Na
can be calculated.

With this calibrated value for the vector polarizability


α1 , a measurement of the Faraday angle θF yields directly
the Fz component of an arbitrary spin state. If the collec-
tive spin is aligned along a magnetic field perpendicular
to the propagation direction of the light field, the Fz com-
ponent oscillates at the Larmor frequency. In this case
7

fit
2. Calibration of the spin damping rate
6 data
One of the main parameters in the experiments is the
θF (mrad)

spin damping rate γs . In order to measure the spin damp-


4 ing rate in the presence of all lasers but without coupling
to the membrane, we detuned the coupling laser from the
cavity resonance (|∆|  κ). The laser thus is reflected
2 from the incoupling mirror of the cavity and only the
spin is probed. For the calibration measurements, the
4 6 8 10 spin was coherently excited by a weak RF-pulse. The
Na (106 ) spin signal was measured by detecting the remaining sig-
nal on the light after the second pass. It is normalised to
FIG. 6. DC Faraday rotation signal of the atomic ensemble occupation numbers [shown in Fig. 7(a)]. The damping
spin-polarized along the optical propagation axis at a probe rate γs is extracted from the exponential fit to the tempo-
laser detuning of −2π × 40 GHz. In the experiment, first the ral dynamics [Fig. 7(b)] and the frequency Ωs is extracted
Faraday rotation angle θF was measured with a weak, far- from a Lorentzian fit to the spectrum [Fig. 7(c)]. For op-
detuned probe pulse and then the number of atoms Na was tical pumping power larger than Ppump > 0.7 µW, the
determined by absorption imaging. Each data point with er- spectra were too broad to provide reasonable fit results
ror bar corresponds to the mean and the standard deviation [and are therefore not shown in Fig. 7(c)]. In Fig. 7(b)
of five experimental runs at the same MOT loading time.
and (c), fit parameters for the coupled dynamics are
shown.

of an oscillating signal, the polarization homodyne mea-


surement of the out-going field is demodulated by a lock- 3. Calibration of the membrane signal
in amplifier which returns the root mean square (rms)
amplitude VS50 Ω
y ,rms
∝ S̄y (where S̄y is the slowly varying
The vibrations of the membrane are detected via their
amplitude of Sy ). In order to determine the Faraday an-
effect on the phase of the beam reflected from the cavity
gle from this rms amplitude and the DC measurement of
[35]. In particular, the membrane vibrations modulate
VSx ∝ hSx i using an oscilloscope, one√has to first mul-
the cavity resonance frequency ωc . For small membrane
tiply the rms amplitude by a factor 2 to get a peak displacements, we can write ωc (xm ) ≈ ωc + Gxm , where
amplitude voltage and further by a factor of 2 to com- G = −dωc /dxm is the cavity frequency shift per mem-
pensate for the impedance mismatch between the 50 Ω brane displacement. For a single-sided cavity, the phase
input impedance of the lock-in amplifier and the high φc of the beam reflected from the cavity with respect
input impedance of the oscilloscope. Including these fac- to the incoming beam is related to the cavity detuning
tors, we get √
the slowly varying amplitude of the Faraday ∆ = ωL − ωc by [35]
angle θ̄F = 2VS50 Ω
y ,rms
/VSx . By normalising the spin sig-
nal by the square-root of the total spin length we obtain 
κ∆

the slowly varying amplitude of the Xs -quadrature of the φc = arctan , (A5)
(κ/2)2 + ∆2
spin
where κ is the cavity linewidth. A change in the cav-
F̄z 2θ̄F 2VS50 Ω
y ,rms ity frequency δωc = Gxm thus leads to a change in
X̄s = p = √ = √ . (A3)
hFx i α1 2Na α1 VSx Na the phase of the reflected beam by an amount δφc =
−(dφc /d∆)δωc ≈ −4Gxm /κ, where the approximation
From this result, we can use the equipartition theorem to holds for small detunings |∆|  κ. We can write the
calculate the number of excitations of the spin oscillator: previous expression in terms of the vacuum optomechan-
ical coupling strength g0 = GxZPF as
1 hXs (t)2 + Ps (t)2 i 2θ̄2 4g0 xm
n̄s + = = X̄s2 = 2 F . (A4) δφc = − , (A6)
2 2 α1 Na κ xZPF

where Xs (t) and Ps (t) are the fast rotating quadratures where we have introduced the zero-point fluctuation am-
of the spin oscillator. In the looped experiment, only a plitude of the membrane xZPF = (~/2meff Ωm )1/2 , with
small fraction of the light was measured in between the meff the effective mass of the vibration mode. These
first interaction of the spin and the interaction of the light phase variations δφc can now be read interferometrically
with the membrane. This in-loop measurement was cal- by means of balanced homodyne detection. For this, the
ibrated using a coherent spin excitation and comparing beam reflected from the cavity is combined with a strong
it to the polarization homodyne measurement presented local oscillator in a 50:50 beam splitter. The output
in this section. beams are subsequently photodetected and the output
8

(a) (b) (c)


linear fit
102 fit simulation
1960
104

Ωs /(2π) (kHz)
γs /(2π) (kHz)
data
hb†s bs i

103 1958
101

102
1956
100
0.25 0.50 0.75 1.00 10−2 10−1 100 0.0 0.2 0.4 0.6
time (ms) Ppump (µW) Ppump (µW)

FIG. 7. Measurement of the spin in the absence of coupling to the membrane after it is excited by a weak RF-pulse: (a) Time
trace of double pass measurement of the spin with different pumping powers (range from 0 to 10 µW). The dashed lines show
fits with an exponential decay. The spin linewidth (b) and spin frequency (c) are plotted as a function of the pump power.
The dashed lines in (b) and (c) show a linear fit to the spin linewidth and resonance frequency. The crosses show the fit
parameters extracted from Fig. 2(a) which were used as input for the simulations. The data shown in (a) is an average over
seven experimental realisations and was used to fit the exponential decay [for (b)] and the Lorentzian peak [for (c)]. The error
bars in (b) and (c) show the fit-error of the corresponding quantity.

signals subtracted. The recorded balanced voltage can its rms value δV (t)50 Ω
rms , which we further need to multiply
be written as by a factor of 2 due to impedance mismatch of our mea-
suring instrument. To convert the measured rms √ value to
V = V0 cos(∆φ), (A7) amplitude variations, we thus need an overall 2 2 factor.
with V0 the modulation amplitude, proportional to the This finally yields
square-root of the power of the beam reflected from the 1
cavity and of the local oscillator beam and with ∆φ = n̄m (t) + = hXm (t)2 it
2
φc − φLO where φLO is the phase of the local oscillator. Ω 2
2
ηc δV (t)50
  
The modulation amplitude V0 is inferred by modulat- rms κ
= . (A10)
ing φLO , thanks to a movable mirror in the local oscillator V0 2g0
path which allows to generate path differences of a few The values of κ = 2π × 77 MHz and g0 = 2π × 224 Hz,
wavelengths. V0 can be extracted from the contrast of have been independently calibrated from the width of
the observed interference fringes. In order to detect the the Pound-Drever-Hall signal and by measuring the op-
phase fluctuations δφc of φc induced by the membrane tomechanical response to an optical amplitude modula-
motion, we lock the relative phase ∆φ to π/2, i.e., the tion tone, respectively.
point where the slope of the fringes is maximal. For small
shifts δφc  π/2, the recorded voltage variation δV (t) is
directly proportional to δφc (t), and thus to xm (t). In Appendix B: Theoretical model for the
practice, δV (t) is effectively reduced by a factor ηc due membrane-spin coupling
to imperfect cavity coupling, such that
ηc δV (t) xZPF κ We modelled our membrane-spin coupling by two cou-
xm (t) = . (A8) pled harmonic oscillators as shown in Eqs. (1) and (2). In
V0 4g0
the following we show how we characterised, simulated,
In order to determine membrane phonon occupation and approximated the system starting from these equa-
n̄m (t) we first define the dimensionless
√ membrane tions. In section B 1 the stochasic simulation of the sys-
quadrature
√ operators X m = x m /( 2x ZPF ) and Pm = tem is presented. In section B 2 we show the derivation
2xZPF pm /~, defined so that [Xm , Pm ] = i. We can now of the fit function for the spectra. From the spectrum, we
write calculate the sympathetic cooling rate and the resonance
hXm (t)2 + Pm (t)2 i frequency shift in the weak coupling limit in section B 3.
hH(t)i= ~Ωm Finally, we show the Routh-Hurwitz stability analysis of
2 
 the coupled dynamics with delay in section B 4.
1
= ~Ωm n̄m (t) + . (A9)
2
By means of the equipartition theorem, we can write 1. Simulation of the spin-membrane dynamics
hXm (t)2 it = hPm (t)2 it and thus relate the measured volt-
age variations to the membrane phonon occupation num- In this section, we provide some details on the simula-
ber. In practice, we do not measure the voltage δV (t) but tion method we used to solve the stochastic equations of
9

motion Eqs. (1) and (2) for the spin-membrane system. be neglected in the rotating frame i.e. X̃j (t) ≈ X̃j (t − τ )
This simulation follows closely the algorithm presented and P̃j (t) ≈ P̃j (t − τ ). The equations of motion then
in [37]. For the simulation, we rewrite the equations of read
motion as four coupled first-order differential equations
for X̃j and P̃j , with j ∈ (m, s) in a frame rotating at the     
membrane frequency Ωm (operators in the rotating frame X̃m (t) X̃m (t) − sin (Ωm t)Fm (t)

are denoted with a tilde) and apply the rotating wave d  P̃m (t)  = −M  P̃m (t)  + 
   cos (Ωm t)Fm (t) 
approximation (RWA). In the limit where the propaga- dt  X̃s (t)   X̃s (t)   − sin (Ωm t)Fs (t)  ,
tion delay is small compared to other timescales involved P̃s (t) P̃s (t) cos (Ωm t)Fs (t)
in the coupled dynamics (i.e. τ  γj−1 , g −1 , δ −1 ), the (B1)
change of the oscillator quadratures during the time τ can where we have split the dynamics into the 4×4 dynamical
matrix

γm /2 0 −g sin (Ωm τ ) −g cos (Ωm τ )


 
0 γm /2 g cos (Ωm τ ) −g sin (Ωm τ ) 
M= , (B2)

−g sin (Ωm τ ) −g cos (Ωm τ ) γs /2 −δ
g cos (Ωm τ ) −g sin (Ωm τ ) δ γs /2

and a stochastic part, given by the generalized noise and the calculated room temperature occupation of the
p (tot) membrane. We assumed the spin pumping to be perfect
forces Fj (t) = 2γj Fj (t). The total force noise
(tot) (th) such that the spin oscillator environment is in its quan-
Fj (t) includes the thermal noise Fj (t) and the back-
(ba)
tum mechanical ground state (i.e. n̄s,bath = 0).
action noise Fj (t) which itself depends on the optical
(in) The approach given in [37] allows for an exact simu-
vacuum noise Fj (t). Thus, it is given by
lation of the stochastic dynamics for a single oscillator
(tot) (th) (ba) for arbitrary time steps, which we extend to the case of
Fj (t)= Fj (t) + Fj (t)
s two coupled oscillators with delay. This is done by calcu-
(th) 2Γj (in) lating for each time step the coherent evolution and the
= Fj (t) + F (t), (B3)
γj j noise separately:

where Γj is the measurement rate of the individual sys-


(ν)
tem. The noise terms Fj (t), ν ∈ (th, in) can be ex-
t →t
∆X̃mi i+1
     
pressed explicitly in terms of the product of a noise am- X̃m (ti+1 ) X̃m (ti )
(ν)  P̃m (ti+1 )  ti →ti+1 
plitude and a zero mean, delta correlated noise fj (t):  = e−M∆t  P̃m (ti )  +  ∆P̃m
  
 X̃s (ti )  ∆X̃sti →ti+1  ,
 
 X̃s (ti+1 ) 
r
1 (th) t →ti+1
(th)
Fj (t) = n̄j,bath + fj (t),
P̃s (ti+1 ) P̃s (ti ) ∆P̃s i
2 (B6)
where ∆t = ti+1 − ti is one simulation time step, and
r
η 2
(in)
Fm (t) = f (in) (t), (B4) t →t t →t
∆X̃j i i+1 , ∆P̃j i i+1 are terms for the stochastic noise
2 m
r which enters the system in between time ti and ti+1 . We
1 − η 4 (in) performed the simulation at time steps comparable to the
Fs(in) (t) = fs (t), (B5) ti →ti+1
2 oscillation period Ω−1
m . Thus, the noise terms ∆X̃j
t →t
where η 2 ≈ 0.8 is the power transmission coefficient of and ∆P̃j i i+1 are correlated which is taken into account
the light between the spin and the membrane and n̄j,bath by following the calculation of noise variances and co-
is the number of thermal phonons in the individual sys- variances in [37]. Because the coupling between the two
tem. The thermal noise amplitude is calculated from the oscillators is much slower than the simulation time step
fluctuation dissipation theorem while for the derivation g  Ωm ≈ ∆t−1 we neglect the correlation of noise build-
of the backaction noise we refer to [36]. The number of ing up between the oscillators during one simulation step.
thermal phonons of the membrane n̄m,bath was measured Thus, we can treat the noise of both oscillators sepa-
by homodyne detection in presence of all laser beams but rately. In order to simulate the system more efficiently,
without loading the atoms. This calibrated value agrees we perform the simulation in time steps of multiples of
very well with an estimation from comparing the spectral one frame rotation ∆t = k·2π/Ωm , k = 1, 2, 3... such that
linewidth in presence of the cooling and coupling beams the noise amplitudes [proportional to sin(Ωm t), cos(Ωm t),
with the spectral linewidth of the uncooled membrane see Eq. (B1)] are the same for each step of the simulation.
10

2. Fit function for the power spectral density of Here, we have defined an effective frequency shift δΩshift
mechanical displacement and the sympathetic cooling rate γsym , which for ω = Ωm
read
In this section, we provide some details on the coupled-
mode model used for fitting the power spectral density 4g 2 Ωm Ωs
δΩ2shift = 2 2
of the mechanical displacement shown in Fig. 2(b). For (Ω2s − Ω2m ) + (Ωm γs )
this, we first Fourier transform the equations of motion × Ω2s − Ω2m cos (2Ωm τ ) − Ωm γs sin (2Ωm τ ) ,
  
Eqs. (1) and (2), which allows us to derive the following (B16)
effective susceptibilities

χm,0 (ω)−1 Xm (ω) + 2g eiωτ Xs (ω) = − 2γm Fm


p (tot)
(ω), and
(B7) 4g 2 Ωm Ωs
γsym =
χs,0 (ω)−1 Xs (ω) + 2g eiωτ Xm (ω) = − 2γs Fs(tot) (ω),
p
2 2
(Ω2s
− Ω2m ) + (Ωm γs )
(B8)
Ω2s − Ω2m
 
× γs cos (2Ωm τ ) + sin (2Ωm τ ) .
where we have defined the individual oscillator suscepti- Ωm
blities as (B17)
Ωi
χi,0 (ω) = . (B9) For Ωs ≈ Ωm and large spin damping γs > g, we get a
Ω2i − ω 2 − iωγi
simplified expression for the frequency shift and sympa-
Solving for Xm and Xs yields thetic cooling rate [Eq. (3)]
 p (tot)
Xm (ω) =χm,eff (ω) − 2γm Fm (ω) 4g 2 Ωm
δΩ2shift ≈ [2δ cos (2Ωm τ ) − γs sin (2Ωm τ )] ,
4δ 2 + γs2
iωτ
p (tot)

+ 2g e 2γs χs,0 (ω)Fs (ω) , (B10)
(B18)
 p 4g 2
Xs (ω) =χs,eff (ω) − 2γs Fs(tot) (ω) γsym ≈ [γs cos (2Ωm τ ) + 2δ sin (2Ωm τ )] ,
p 4δ 2 + γs2
+ 2g eiωτ 2γm χm,0 (ω)Fm (tot)

(ω) , (B11) (B19)
where we have introduced the effective susceptibilities of
where δ = Ωs − Ωm .
the membrane and spin oscillators as
χm,eff (ω)−1 = χm,0 (ω)−1 − 4g 2 ei2ωτ χs,0 (ω), (B12)
χs,eff (ω) −1
= χs,0 (ω) −1 2 i2ωτ
− 4g e χm,0 (ω). (B13) 4. Routh-Hurwitz stability criterion of the coupled
system
We used this model to fit the power spectral densities of
the mechanical displacement spectra [see Fig. 2(b)] using In this section we present a stability analysis in which
as fit function a2 |χm,eff (ω)|2 where a is a global scaling the Routh-Hurwitz criterion [41] from control theory is
factor accounting for the noise terms driving the system. applied to our linearly coupled spin-membrane oscilla-
tors. The criterion provides a convenient means to assess
the stability of our linear systems without solving the
3. Derivation of the sympathetic cooling rate equations of motion. In this treatment, we exclude the
Langevin noise, as we are interested to see if the delayed
Here, we derive the sympathetic cooling rate for the coupled oscillator dynamics is stable by itself. We then
mechanical oscillator given in Eq. (3). For this, let us explore the experimental parameter space to see under
first write Eq. (B12) explicitly which conditions the coupled system becomes unstable.
We take the equations of motion for the delayed coupled
1 system Eqs. (1) and (2) neglecting the noise terms
χm,eff (ω)−1 = Ω2m − ω 2 − iωγm
Ωm
2 2
! Ẍm + γm Ẋm + Ω2m Xm = −2gΩm Xs (t − τ ), (B20)
i2ωτ Ωm Ωs Ωs − ω + iωγs
− 4g 2 e 2 2 , Ẍs + γs Ẋs + Ω2s Xs = −2gΩs Xm (t − τ ). (B21)
(Ω2s − ω 2 ) + (ωγs )
(B14) Substituting the ansatz Xj (t) = Xj (s) est where s ∈ C
which can be written in the form of yields
1  2 s2 + sγm + Ω2m Xm (s) = −2gΩm e−sτ Xs (s), (B22)

χm,eff (ω)−1 = Ω − δΩ2shift − ω 2 − iω (γm + γsym ) .

Ωm m
s2 + sγs + Ω2s Xs (s) = −2gΩs e−sτ Xm (s).

(B15) (B23)
11

2.00 Having our dynamics in this polynomial form, we can


define the polynomial coefficients of a fourth order poly-
1.75 nomial by

1.50 p(s) = a4 s4 + a3 s3 + a2 s2 + a1 s + a0 = 0, a4 > 0. (B26)

1.25 τ = 80 ns
In order to apply the Routh-Hurwitz criterion, the so-
γs /(2g)

τ = 30 ns called Hurwitz matrix containing the polynomial coeffi-


1.00
τ = 15 ns cients has to be defined. For a fourth order polynomial
this matrix reads
0.75
a3 a1 0 0
 
0.50 τ = 8 ns a a a 0 
H4 =  4 2 0 . (B27)
0 a3 a1 0 
τ = 4 ns 0 a4 a2 a0
0.25

According to the Routh-Hurwitz criterion, the system


0.00
−2 −1 0 1 2 dynamics is asymptotically stable if all the principal mi-
δ/(2g) nors of the Hurwitz matrix are non-zero and positive.
Application of the Hurwitz criterion leads to the follow-
FIG. 8. Evaluation of the stability of the coupled system
ing stability criteria for a fourth order polynomial system:
using the Routh-Hurwitz criterion: The colored regions (i.e.
region above each solid line) show the sets of parameters for ∆1 = |a3 | > 0, (B28)

which the coupled dynamics is stable for a given value of a3 a1
∆2 = = a2 a3 − a4 a1 > 0, (B29)
the feedback delay. Without propagation delay, every set of

a4 a2
detunings and spin damping leads to stable dynamics. For
τ = 80 ns we have Ωm τ ≈ 1 thus the validity of the Taylor
a3 a1 0
∆3 = a4 a2 a0 = a1 ∆2 − a23 a0 > 0,

expansion of the exponential function in presence of small (B30)
delays reaches its limit. For the stability estimations shown 0 a3 a1
here we used 2g = 2π × 6.8 kHz, γm = 2π × 262 Hz, and
Ωm = 2π × 1.957 MHz
∆4 = det(H4 ) = a0 · ∆3 > 0. (B31)

In our system, the coefficients are given explicitly by


Solving the simultaneous equations Eqs. (B22) and a4 = 1, (B32)
(B23), we obtain the characteristic equation for non-
trivial solutions Xm 6= 0, a3 = γs + γm , (B33)
a2 = Ω2s + Ω2m + γs γm , (B34)
s2 + sγm + Ω2m s2 + sγs + Ω2s − 4g 2 Ωm Ωs e−2sτ = 0.
 
a1 = γm Ω2s + γs Ω2m 2
+ 8g Ωm Ωs τ, (B35)
(B24)
Ωs Ωm − 4g 2 .

For clarity, we consider here small propagation delays a0 = Ωs Ωm (B36)
τ  1/Ωj and apply a first order Taylor expansion
exp(−2sτ ) ≈ (1 − 2sτ ) (in the actual simulation we keep Since Ωs Ωm  4g 2 , all coefficients are positive. Thus,
terms up to 4th order). We then obtain the criterion ∆1 is fulfilled and the criterion ∆4 depends
directly on the criterion ∆3 . Therefore, only ∆2 and ∆3
0 =s4 + (γs + γm )s3 + (Ω2m + Ω2s + γm γs )s2 are left to be checked. In order to get an intuition on
the stability for different parameters, Fig. 8 shows the
+ (Ω2s γm + Ω2m γs + 8g 2 Ωm Ωs τ )s
stable regions as a function of spin damping, detuning
+ Ωm Ωs (Ωm Ωs − 4g 2 ). (B25) and delay.

[1] Philipp Treutlein, C Genes, Klemens Hammerer, and Schmiedmayer, “Quantum technologies with hybrid sys-
M Poggio, Hybrid mechanical systems, edited by Markus tems,” PNAS 112, 3866–3873 (2015).
Aspelmeyer, Tobias J. Kippenberg, and Florian Mar- [3] Yiwen Chu and Simon Gröblacher, “A perspective on
quardt (Springer Berlin Heidelberg, Berlin, Heidelberg, hybrid quantum opto- and electromechanical systems,”
2014) pp. 327–351. Appl. Phys. Lett. 117, 150503 (2020).
[2] Gershon Kurizki, Patrice Bertet, Yuimaru Kubo, Klaus [4] Stephan Camerer, Maria Korppi, Andreas Jöckel, David
Mølmer, David Petrosyan, Peter Rabl, and Jörg Hunger, Theodor W. Hänsch, and Philipp Treutlein,
12

“Realization of an optomechanical interface between ul- [17] A. Barfuss, J. Teissier, E. Neu, A. Nunnenkamp, and
tracold atoms and a membrane,” Phys. Rev. Lett. 107, P. Maletinsky, “Strong mechanical driving of a single
223001 (2011). electron spin,” Nat. Phys. 11, 820–824 (2015).
[5] Andreas Jöckel, Aline Faber, Tobias Kampschulte, Maria [18] Donghun Lee, Kenneth W Lee, Jeffrey V Cady, Preeti
Korppi, Matthew T. Rakher, and Philipp Treutlein, Ovartchaiyapong, and Ania C Bleszynski Jayich, “Top-
“Sympathetic cooling of a membrane oscillator in a hy- ical review: spins and mechanics in diamond,” J. Opt.
brid mechanical–atomic system,” Nat. Nanotechnol. 10, 19, 033001 (2017).
55–59 (2015). [19] Seth Lloyd, “Coherent quantum feedback,” Phys. Rev. A
[6] Philipp Christoph, Tobias Wagner, Hai Zhong, Roland 62, 022108 (2000).
Wiesendanger, Klaus Sengstock, Alexander Schwarz, [20] Jing Zhang, Yu xi Liu, Re-Bing Wu, Kurt Jacobs, and
and Christoph Becker, “Combined feedback and sympa- Franco Nori, “Quantum feedback: Theory, experiments,
thetic cooling of a mechanical oscillator coupled to ultra- and applications,” Phys. Rep. 679, 1–60 (2017).
cold atoms,” New J. Phys. 20, 093020 (2018). [21] Howard M. Wiseman and Gerard J. Milburn,
[7] Christoffer B. Møller, Rodrigo A. Thomas, Georgios Vasi- Quantum Measurement and Control (Cambridge Uni-
lakis, Emil Zeuthen, Yeghishe Tsaturyan, Mikhail Bal- versity Press, 2009).
abas, Kasper Jensen, Albert Schliesser, Klemens Ham- [22] Ryan Hamerly and Hideo Mabuchi, “Advantages of co-
merer, and Eugene S. Polzik, “Quantum back-action- herent feedback for cooling quantum oscillators,” Phys.
evading measurement of motion in a negative mass ref- Rev. Lett. 109, 173602 (2012).
erence frame,” Nature 547, 191–195 (2017). [23] James S Bennett, Lars S Madsen, Mark Baker, Halina
[8] Thomas M. Karg, Baptiste Gouraud, Chun Tat Ngai, Rubinsztein-Dunlop, and Warwick P Bowen, “Coherent
Gian-Luca Schmid, Klemens Hammerer, and Philipp control and feedback cooling in a remotely coupled hybrid
Treutlein, “Light-mediated strong coupling between a atom–optomechanical system,” New J. Phys. 16, 083036
mechanical oscillator and atomic spins 1 meter apart,” (2014).
Science 369, 174–179 (2020). [24] Masashi Hirose and Paola Cappellaro, “Coherent feed-
[9] Rodrigo A. Thomas, Michal Parniak, Christoffer back control of a single qubit in diamond,” Nature 532,
Østfeldt, Christoffer B. Møller, Christian Bærentsen, 77–80 (2016).
Yeghishe Tsaturyan, Albert Schliesser, Jürgen Appel, [25] J. Zhang, R. Wu, Y. Liu, C. Li, and T. Tarn, “Quantum
Emil Zeuthen, and Eugene S. Polzik, “Entanglement be- coherent nonlinear feedback with applications to quan-
tween distant macroscopic mechanical and spin systems,” tum optics on chip,” IEEE Trans. Autom. Control 57,
Nat. Phys. 17, 228–233 (2021). 1997–2008 (2012).
[10] I. Yeo, P.-L. de Assis, A. Gloppe, E. Dupont-Ferrier, [26] Zhaoyou Wang and Amir H. Safavi-Naeini, “Enhancing a
P. Verlot, N. S. Malik, E. Dupuy, J. Claudon, J.-M. slow and weak optomechanical nonlinearity with delayed
Gérard, A. Auffèves, G. Nogues, S. Seidelin, J.-Ph Poizat, quantum feedback,” Nature Communications 8, 15886
O. Arcizet, and M. Richard, “Strain-mediated coupling (2017).
in a quantum dot–mechanical oscillator hybrid system,” [27] Alfred Harwood, Matteo Brunelli, and Alessio Serafini,
Nat. Nanotechnol. 9, 106–110 (2014). “Cavity optomechanics assisted by optical coherent feed-
[11] Michele Montinaro, Gunter Wüst, Mathieu Munsch, back,” Phys. Rev. A 103, 023509 (2021).
Yannik Fontana, Eleonora Russo-Averchi, Martin Heiss, [28] M. J. Woolley and A. A. Clerk, “Two-mode squeezed
Anna Fontcuberta i Morral, Richard J. Warburton, and states in cavity optomechanics via engineering of a single
Martino Poggio, “Quantum Dot Opto-Mechanics in a reservoir,” Phys. Rev. A 89, 063805 (2014).
Fully Self-Assembled Nanowire,” Nano Lett. 14, 4454– [29] Jie Li, Gang Li, Stefano Zippilli, David Vitali, and Tian-
4460 (2014). cai Zhang, “Enhanced entanglement of two different me-
[12] A. D. O’Connell, M. Hofheinz, M. Ansmann, Ra- chanical resonators via coherent feedback,” Phys. Rev. A
doslaw C. Bialczak, M. Lenander, Erik Lucero, M. Nee- 95, 043819 (2017).
ley, D. Sank, H. Wang, M. Weides, J. Wenner, John M. [30] Klemens Hammerer, Anders S. Sørensen, and Eugene S.
Martinis, and A. N. Cleland, “Quantum ground state Polzik, “Quantum interface between light and atomic en-
and single-phonon control of a mechanical resonator,” sembles,” Rev. Mod. Phys. 82, 1041–1093 (2010).
Nature 464, 697–703 (2010). [31] Luca Pezzè, Augusto Smerzi, Markus K. Oberthaler, Ro-
[13] Patricio Arrangoiz-Arriola, E. Alex Wollack, Zhaoyou man Schmied, and Philipp Treutlein, “Quantum metrol-
Wang, Marek Pechal, Wentao Jiang, Timothy P. ogy with nonclassical states of atomic ensembles,” Rev.
McKenna, Jeremy D. Witmer, Raphaël Van Laer, and Mod. Phys. 90, 035005 (2018).
Amir H. Safavi-Naeini, “Resolving the energy levels of a [32] B. Vogell, T. Kampschulte, M. T. Rakher, A. Faber,
nanomechanical oscillator,” Nature 571, 537–540 (2019). P. Treutlein, K. Hammerer, and P. Zoller, “Long dis-
[14] A. A. Clerk, K. W. Lehnert, P. Bertet, J. R. Petta, tance coupling of a quantum mechanical oscillator to the
and Y. Nakamura, “Hybrid quantum systems with cir- internal states of an atomic ensemble,” New J. Phys. 17,
cuit quantum electrodynamics,” Nat. Phys. 16, 257–267 043044 (2015).
(2020). [33] M. Wallquist, K. Hammerer, P. Zoller, C. Genes, M. Lud-
[15] D. Rugar, R. Budakian, H. J. Mamin, and B. W. Chui, wig, F. Marquardt, P. Treutlein, J. Ye, and H. J. Kim-
“Single spin detection by magnetic resonance force mi- ble, “Single-atom cavity QED and optomicromechanics,”
croscopy,” Nature 430, 329–332 (2004). Phys. Rev. A 81, 023816 (2010).
[16] O. Arcizet, V. Jacques, A. Siria, P. Poncharal, P. Vin- [34] J. D. Thompson, B. M. Zwickl, A. M. Jayich, Florian
cent, and S. Seidelin, “A single nitrogen-vacancy defect Marquardt, S. M. Girvin, and J. G. E. Harris, “Strong
coupled to a nanomechanical oscillator,” Nat. Phys. 7, dispersive coupling of a high-finesse cavity to a microme-
879–883 (2011). chanical membrane,” Nature 452, 72–75 (2008).
13

[35] Markus Aspelmeyer, Tobias J. Kippenberg, and Florian atomic motion in an optical lattice coupled to a mem-
Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. brane,” Phys. Rev. Lett. 120, 073602 (2018).
86, 1391–1452 (2014). [41] Eberhard P. Hofer, Grundlagen der Regelungstechnik
[36] Thomas M. Karg, Baptiste Gouraud, Philipp Treutlein, (Open Access Repositorium der Universität Ulm, 2008)
and Klemens Hammerer, “Remote Hamiltonian interac- pp. 40–51.
tions mediated by light,” Phys. Rev. A 99, 063829 (2019). [42] Y. Tsaturyan, A. Barg, E. S. Polzik, and A. Schliesser,
[37] Simon F. Nørrelykke and Henrik Flyvbjerg, “Harmonic “Ultracoherent nanomechanical resonators via soft
oscillator in heat bath: Exact simulation of time-lapse- clamping and dissipation dilution,” Nat. Nanotechnol.
recorded data and exact analytical benchmark statis- 12, 776–783 (2017).
tics,” Phys. Rev. E 83, 041103 (2011). [43] C. Reetz, R. Fischer, G. G. T. Assumpção, D. P. McNally,
[38] D. V. Ramana Reddy, A. Sen, and G. L. Johnston, P. S. Burns, J. C. Sankey, and C. A. Regal, “Analysis of
“Time delay induced death in coupled limit cycle oscil- Membrane Phononic Crystals with Wide Band Gaps and
lators,” Phys. Rev. Lett. 80, 5109–5112 (1998). Low-Mass Defects,” Phys. Rev. Appl. 12, 044027 (2019).
[39] D. V. Ramana Reddy, A. Sen, and G. L. Johnston, “Ex- [44] J. M. Geremia, John K. Stockton, and Hideo Mabuchi,
perimental evidence of time-delay-induced death in cou- “Tensor polarizability and dispersive quantum measure-
pled limit-cycle oscillators,” Phys. Rev. Lett. 85, 3381– ment of multilevel atoms,” Phys. Rev. A 73, 042112
3384 (2000). (2006).
[40] Aline Vochezer, Tobias Kampschulte, Klemens Ham-
merer, and Philipp Treutlein, “Light-mediated collective

You might also like