File name: Supplementary Information Description: Supplementary Figures, Supplementary Notes and Supplementary References

Similar documents
Day 3: Ultracold atoms from a qubit perspective

Supplementary Information for Coherent optical wavelength conversion via cavity-optomechanics

Advanced Workshop on Nanomechanics September Quantum Measurement in an Optomechanical System

SUPPLEMENTARY INFORMATION

Optomechanics and spin dynamics of cold atoms in a cavity

Theory of bifurcation amplifiers utilizing the nonlinear dynamical response of an optically damped mechanical oscillator

QND for advanced GW detectors

Quantum optics and squeezed states of light

Tailorable stimulated Brillouin scattering in nanoscale silicon waveguides.

Quantum-noise reduction techniques in a gravitational-wave detector

Single Emitter Detection with Fluorescence and Extinction Spectroscopy

A Guide to Experiments in Quantum Optics

Quantum Mechanical Noises in Gravitational Wave Detectors

Squeezed Light for Gravitational Wave Interferometers

Quantum Electronics Laser Physics PS Theory of the Laser Oscillation

Evaluation of Second Order Nonlinearity in Periodically Poled

Quantum enhanced magnetometer and squeezed state of light tunable filter

Squeezed states of light - generation and applications

Atom assisted cavity cooling of a micromechanical oscillator in the unresolved sideband regime

An Opto-Mechanical Microwave-Rate Oscillator

Study of a quantum nondemolition interferometer using ponderomotive squeezing

The Quantum Limit and Beyond in Gravitational Wave Detectors

Supplementary Figures

Quality Factor Thickness (nm) Quality Factor Thickness (nm) Quality Factor 10

Squeezed Light Techniques for Gravitational Wave Detection

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41

Supplementary information

Enhancing sensitivity of gravitational wave antennas, such as LIGO, via light-atom interaction

ECE 240a - Notes on Spontaneous Emission within a Cavity

The Gouy phase shift in nonlinear interactions of waves

Supplementary Figure 1 Level structure of a doubly charged QDM (a) PL bias map acquired under 90 nw non-resonant excitation at 860 nm.

Two-electron systems

Cavity optomechanics: Introduction to Dynamical Backaction

SUPPLEMENTARY INFORMATION

Noise in voltage-biased scaled semiconductor laser diodes

Young-Shin Park and Hailin Wang Dept. of Physics and Oregon Center for Optics, Univ. of Oregon CLEO/IQEC, June 5, Supported by NSF and ARL

Radiation pressure effects in interferometric measurements

SUPPLEMENTARY INFORMATION

Interferometry beyond the Standard Quantum Limit using a Sagnac Speedmeter Stefan Hild

Supplementary Figure 1: Reflectivity under continuous wave excitation.

RÉSONATEURS NANOMÉCANIQUES DANS LE RÉGIME QUANTIQUE NANOMECHANICAL RESONATORS IN QUANTUM REGIME

Non-linear Optics II (Modulators & Harmonic Generation)

Lecture 9. PMTs and Laser Noise. Lecture 9. Photon Counting. Photomultiplier Tubes (PMTs) Laser Phase Noise. Relative Intensity

Verification of conditional quantum state in GW detectors. LSC-VIRGO Meeting, QND workshop Hannover, October 26, 2006

S. Blair February 15,

Molecular spectroscopy

Where are the Fringes? (in a real system) Div. of Amplitude - Wedged Plates. Fringe Localisation Double Slit. Fringe Localisation Grating

opto-mechanical filtering

Measured Transmitted Intensity. Intensity 1. Hair

PART 2 : BALANCED HOMODYNE DETECTION

Supplemental material for Bound electron nonlinearity beyond the ionization threshold

Observation of the nonlinear phase shift due to single post-selected photons

arxiv:quant-ph/ v3 25 Jun 2004

How to measure a distance of one thousandth of the proton diameter? The detection of gravitational waves

Survey on Laser Spectroscopic Techniques for Condensed Matter

Nanomechanics II. Why vibrating beams become interesting at the nanoscale. Andreas Isacsson Department of Physics Chalmers University of Technology

Short Wavelength Regenerative Amplifier FELs (RAFELs)

A few Experimental methods for optical spectroscopy Classical methods Modern methods. Remember class #1 Generating fast LASER pulses

Citation for published version (APA): Mollema, A. K. (2008). Laser cooling, trapping and spectroscopy of calcium isotopes s.n.

arxiv: v1 [cond-mat.mes-hall] 25 Mar 2018

SUPPLEMENTARY INFORMATION

Simple strategy for enhancing terahertz emission from coherent longitudinal optical phonons using undoped GaAs/n-type GaAs epitaxial layer structures

Let us consider a typical Michelson interferometer, where a broadband source is used for illumination (Fig. 1a).

Squeezed Light and Quantum Imaging with Four-Wave Mixing in Hot Atoms

Niels Bohr Institute Copenhagen University. Eugene Polzik

Photon Physics. Week 4 26/02/2013

The Compton backscattering Polarimeter of the A4 Experiment

Chapter 2. Optical FMCW Reflectometry. 2.1 Introduction

Quantum Noise in Mirror-Field Systems: Beating the Standard Quantum IARD2010 Limit 1 / 23

Optomechanically induced transparency of x-rays via optical control: Supplementary Information

Frequency dependent squeezing for quantum noise reduction in second generation Gravitational Wave detectors. Eleonora Capocasa

Progress Report on the A4 Compton Backscattering Polarimeter

Towards a quantum interface between light and microwave circuits

Measuring Laser Diode Optical Power with an Integrating Sphere

SUPPLEMENTARY INFORMATION

OPTI 511R, Spring 2018 Problem Set 10 Prof. R.J. Jones Due Thursday, April 19

Quantum Reservoir Engineering

Interferometric. Gravitational Wav. Detectors. \p World Scientific. Fundamentals of. Peter R. Sawlson. Syracuse University, USA.

Temporal modulation instabilities of counterpropagating waves in a finite dispersive Kerr medium. II. Application to Fabry Perot cavities

Performance Limits of Delay Lines Based on "Slow" Light. Robert W. Boyd

Dynamical Casimir effect in superconducting circuits

Vibration-Free Pulse Tube Cryocooler System for Gravitational Wave Detectors I

Scaling law in signal recycled laser-interferometer gravitational-wave detectors

FIG. 16: A Mach Zehnder interferometer consists of two symmetric beam splitters BS1 and BS2

ANALYSIS OF AN INJECTION-LOCKED BISTABLE SEMICONDUCTOR LASER WITH THE FREQUENCY CHIRPING

Supplemental Material to the Manuscript Radio frequency magnetometry using a single electron spin

The laser oscillator. Atoms and light. Fabry-Perot interferometer. Quiz

Quantum Noise and Quantum Measurement

gives rise to multitude of four-wave-mixing phenomena which are of great

OPTI 511, Spring 2016 Problem Set 9 Prof. R. J. Jones

The laser oscillator. Atoms and light. Fabry-Perot interferometer. Quiz

Transit time broadening contribution to the linear evanescent susceptibility

Quantum optics of many-body systems

Quantum Optics in Wavelength Scale Structures

Charge noise and spin noise in a semiconductor quantum device

In Situ Imaging of Cold Atomic Gases

Self-Phase Modulation in Optical Fiber Communications: Good or Bad?

Cavity optomechanics in new regimes and with new devices

Supplementary Figures

Cavity Optomechanics:

Transcription:

File name: Supplementary Information Description: Supplementary Figures, Supplementary Notes and Supplementary References File name: Peer Review File Description:

Optical frequency (THz) 05. 0 05. 5 05.7 1 05.6 5 05.5 PSD (dbm/hz) b 10 4 6 4 6 Power spectral density (dbm/hz) a 0 1 10 Supplementary Figure 1 Nonlinear transduction as a function of detuning. (a) Electronic spectrum analyser signal taken at room temperature, as a function of the optical frequency. (b) Cross section at cavity resonance frequency corresponding to Fig. 3a in the main text. As can be seen in (a), due to the quadrature-averaged homodyne detection, all peaks show a single-peaked detuning dependence which is maximum at resonance. Supplementary Note 1 Power dependence of transduced mechanical motion The transduction of mechanical motion depends on the optical power used, as shown in the dependence of hp i on A Pin PLO in Eq. of the main text. In particular, we expect the measured band power to be linear in both the power in the signal arm Pin, incident on the cavity, and the power in the reference arm PLO. Deviations from a linear relation indicate heating or cooling of the mechanical motion due to the additional laser power 1. Supplementary Figure 4 5 6 7 Prms (W) f1 fit f1 f fit f 1 0 1 Incident optical power (nw) Supplementary Figure Power dependence. Measured band power of the two fundamental mechanical frequencies f1 and f (area under the fitted Lorentzian peak) as a function of the incident optical power Pin, with the laser frequency on-resonance with the cavity. The power in the reference arm, PLO, was kept constant. The black dashed and solid lines show linear fits to the data for f1 and f, respectively. 1

Linewidth (GHz).0 g0, 1/ =. MHz,. MHz g0, / = 6. 1MHz, 6. 5MHz / =. 0GHz / = 5. 5GHz 1 0 Temperature (K) Supplementary Figure 3 Characterisation of a second device. Optical linewidth versus temperature data for a second device. For the dataset with blue squares a second cryostat window at 30 K was not in place, hence exposing the sample to thermal radiation from the outer window at room temperature. The solid and dashed black lines are fits with a model that assumes a constant Lorentzian intrinsic linewidth κ convolved with a Gaussian with a width L G that depends on T. The asymptotes (straight grey lines) of the fit functions allow us to extract κ and the optomechanical coupling rates g 0,i for the two mechanical modes (i = 1, ). Extracted values are shown in the figure. We assign the discrepancy in extracted values of κ to bad thermalisation of the sample in the case where the inner window was not in place. shows transduced mechanical motion as a function of the incident optical power, with the laser on-resonance with the cavity and while the device was cooled to 3 K. The data was corrected for the local oscillator power, which was mostly kept constant and did not vary by more than a factor. At higher powers (near 0 nw), fluctuations in the centre frequency and shape of the peak could be observed (similar to the effects shown off-resonance in Fig. 5 in the main text), however the peak area still follows the linear dependence as a function of the optical power incident at the structure. This shows that for all probe powers used here, there is no significant heating due to the incident laser light. Supplementary Note Characterisation of a second device and estimation of thermalisation The data labelled as device were taken on a sample with a slightly different design, which did not include the double-period modulation to increase the outcoupling, as well as a different shape of the holes. The optical linewidth as a function of temperature for this device is shown in Supplementary Figure 3, in two separate data sets. The data set shown by blue squares was taken with only one window in the cryostat, which was at room temperature, such that the sample was not shielded from thermal radiation. The other dataset was taken with the second window (cooled to around 50 K) in place. Therefore the sample temperature for the blue dataset was most likely higher than the thermometer temperature, which we used to plot and fit these data. Although this leads to a relatively large difference in the fitted optical loss rate κ, both datasets yield very similar coupling rates g 0, since at high temperatures the sample is well thermalised. We note that the different estimate of κ in the absence of

the cold window is of course the direct result of the fact that the nanobeam temperature is larger than that of the nearby thermometer (which is plotted on the horizontal axis of Supplementary Figure 3). Comparing the solid and dashed curves shows that the temperature is raised by degrees from the base temperature. Since this temperature rise can be related to the presence of the nearby 300 K window, we can now estimate the rise that is expected due to the 50 K cold window that is in place during the measurements reported in the main text. Assuming a T 4 dependence of heat transfer rates results in an expected temperature rise of 0.1 K. This shows that the nanobeams are very likely thermalised well down to the lowest cryostat temperatures. We finally remark that it is not straightforward in our case to verify thermalisation using a phase modulation calibration, as described by Gorodetsky and coworkers : The strong intrinsic cavity frequency modulation means that such a calibration tone will inevitably mix with the mechanically-induced modulations, such that its magnitude is strongly altered and in fact itself depends on temperature. Supplementary Note 3 Quadratic displacement measurements The signal-to-noise ratio (SNR) of a quadratic displacement measurement, in the regime where the order-by-order approximation is valid, can be found by considering Eq. in the main text, and assuming the measurement is shot-noise limited. If the optical power in the reference arm is much larger than in the signal arm of the interferometer (P LO P in ), the power spectral density due to the optical shot noise is given by SP SN P = ω cp LO h, with h quantifying the measurement efficiency for the local oscillator light (that is, losses between the homodyne beamsplitter and the detector, and the quantum efficiency of the detector). The band power at Ω m due to thermal mechanical fluctuations is given by Eq. in the main text, with k = and δω = n th g0, leading to ( P Ωm = 51P in P LO hη g ) 4 0 n κ th, (1) where we also substituted A = P in P LO hη, where h accounts for the finite measurement efficiency for the signal beam (Supplementary Note 4). The power spectral density will show a Lorentzian peak with linewidth Γ at the frequency Ω m, which means its peak value is related to the band power (area under the peak) as SP max P = P Ωm /Γ. Finally, we take the signal-to-noise ratio SNR Smax P P SP SN P = 56 P inξη ω c Γ ( g0 ) 4 n κ th, () where we have defined ξ h/h. We set SNR to 1 to find the imprecision in terms of a phonon number n imp, leading to 1 ( g0 ) Pin ξ = 16 η n imp κ ω c Γ. (3) Supplementary Figure 4 shows additional data for the quadratic measurement presented in the main text. The top panel shows the peak around the 3

.0-1.5 1.0-0.5 0.0 3.00 b PSD (W/Hz) f f1 PSD (W/Hz) Piezo control voltage (V) a f1 3.5 3.50 6.5 6.50 f 6.75 f1 f + f1 1 1 1 3.35 3.330 1 1 6.56 6.53 1 0.10 0.16 3.00 3.05 6.400 6.405 f 6.646 6.65 Supplementary Figure 4 Selective linear and quadratic measurement. (a) Spectra showing the first and second order transduction obtained simultaneously, as a function of the piezo mirror position. (b) Maximum and minimum linear and quadratic transduction at different settings of the homodyne phase, for the fundamental frequencies f1 and f as well as all second-order peaks. Red (blue) data points: homodyne phase set to maximize linear (quadratic) transduction; black solid lines show Lorentzian fits to the peaks. Grey data points were taken with no incident optical power in the signal arm. For these measurements, the sample was at 3 K, and incident optical power was 1. nw. fundamental frequency decrease at the same piezo mirror position where the peaks at twice the frequency are strongest, further confirming the ability to suppress the linear measurement while performing the quadratic measurement. The bottom panel shows the linear and quadratic measurements at optimum positions of the piezo mirror for both fundamental frequencies and all 4 secondorder combinations. Supplementary Note 4 Coupling and measurement efficiency Supplementary Figure 5 shows the fitted centre frequency of the mechanical resonance f1 at 3 K with Pin = 0.6 nw, corresponding to the spectrogram shown in Fig. 5a in the main text. Fitting the change in frequency as a function of the optical detuning with the function describing the optical spring effect in the bad-cavity limit 3; provides us with an estimate of the incoupling efficiency η. We measure the optical power Pin before the objective lens, so aside from polarisation mismatch and mode overlap with a Gaussian beam, η also includes transmission losses from the (anti-reflection coated) objective lens and (uncoated) cryostat windows. The average over 5 measurements at different input power yielded a coupling efficiency η = 1.3 %, with a standard deviation of 0.3 %. This analysis ignores the effects of the large thermal fluctuations, which as 4

Centre frequency (MHz) 3.177 3.176 3.175 3.174 fitted incoupling: 1.37% 05.6 05.7 05.75 05.7 Optical frequency (THz) Supplementary Figure 5 Optical spring effect at low temperature. Fitted centre frequency of the mechanical resonance f 1 with the structure at 3 K and P in = 0.6 nw, as a function of the laser frequency (corresponding to the data set shown in Fig. 5a in the main text). The solid line shows a fit of the data according to the linearised model for the optical spring effect, where the magnitude of the effect depends on the incoupling efficiency η. The extracted value is shown in the figure. we show in the main text broadens and reduces the strength of the optical spring effect. This means the obtained coupling efficiency provides a lower limit for the actual coupling efficiency. However, the numerical model we used to generate the simulated spectra in Fig. 5 in the main text shows that the fluctuations at 3 K mainly broaden the mechanical spectra but do not yet lead to strong reduction in the strength of the optical spring effect. In order to characterise the full measurement efficiency ζ = hη, we also need to consider the quantum efficiency of our homodyne detection (h). This includes mainly the overlap of the signal and local oscillator beams ( 40 % in the measurements presented), and the quantum efficiency of our photodiode ( 0 %). We also lose some photons via losses in the optical path and by using a variable aperture to balance the powers reaching the two photodiodes of our balanced photodetector. Overall, we estimate h 5 % in experiments in this paper. We note that we did not try extensively to optimise h, and we expect it could be improved significantly with modest effort. This would be essential for the feedback schemes mentioned in the main text. Supplementary Note 5 Radiation pressure force in a cavity with large thermal fluctuations We express the radiation pressure force in the cavity as F rad = ( ω c/ x)n max c 1 + u, (4) where u κ ( + ( ω c/ x)x) and n max c is the maximum number of photons in the cavity when it is driven at resonance. We assume the resonator moves harmonically: x(t) = x 0 sin Ωt. This results in a time-dependent normalized detuning u(t) = u( ) + u 0 sin Ωt, where u 0 = ( ω c / x)x 0 /κ. Given an amplitude x 0, we extract the first Fourier coefficient of the force 5

a 100 00 00 600 400 00 0 Linewidth (Hz) Fitted value median 3 K 05.70 05.7 05.74 05.76 05.7 Optical frequency (THz) b 00 00 00 700 600 500 400 300 00 Linewidth (Hz) Fitted value median 5 K 05.45 05.60 05.75 05.0 06.05 Optical frequency (THz) Supplementary Figure 6 Mechanical linewidth. Fitted mechanical linewidth as a function of laser frequency at (a) 3 K and (b) 5 K, corresponding to the detuning sweeps shown in the main text in Fig. 5a & d. The solid horizontal lines show the median value of the fitted linewidth, which was used as the intrinsic linewidth for the simulated spectrograms shown in Fig. 5b & e in the main text. (at the resulting harmonic frequency Ω): a rad 1 = Ω π π/ω 0 F rad (t ) sin Ωt dt. (5) In analogy with the mechanical restoring force, we can calculate an effective spring constant from the Fourier coefficient: with F = kx and x(t) = x 0 sin Ωt, F = n=1 a n sin nωt implies that a 1 = kx 0. Finally, a correction on the spring constant results in a correction on the frequency Ω: Ω = k + krad k m = 1 + k rad m k, (6) which we can approximate using 1 + x 1 + x/ +...: k Ω m (1 + k rad ), (7) k therefore δω k k rad m k = k rad mω. () To obtain the results presented in Fig. 5 in the main text, we perform this calculation for 4 amplitudes x 0. These amplitudes are sampled from a Rayleigh distribution, which describes the thermal motion (equivalently, we take x 1 + x with x 1, x independently sampled from a normal distribution). The mean amplitude is set such that the variance of u corresponds to the total variance we expect from both mechanical modes, using the measured values of κ, g 0,i, Ω m,i and the temperature. To generate spectra that can be compared to the experimental data, we then average together Lorentzian lineshapes with centre frequencies given by the resulting set of frequency shifts. Since the mechanical linewidth changes with temperature, we choose the widths of these individual Lorentzian peaks based on the median value of the measured linewidth. As Supplementary Figure 6 shows, this is a good approximation of the intrinsic mechanical linewidth, which we measure for large detuning. 6

Supplementary References [1] Aspelmeyer, M., Kippenberg, T. J. & Marquardt, F. Cavity optomechanics. Rev. Mod. Phys. 6, 131 145 (014). [] Gorodetsky, M. L., Schliesser, A., Anetsberger, G., Deleglise, S. & Kippenberg, T. J. Determination of the vacuum optomechanical coupling rate using frequency noise calibration. Opt. Express 1, 336 46 (0). [3] Leijssen, R. & Verhagen, E. Strong optomechanical interactions in a sliced photonic crystal nanobeam. Sci. Rep. 5, 1574 (015). 7