<<

Cavity Optomechanics with Feedback and Fluids

Glen Ivor Harris B.A./B.Sc. (Hons.), University of Queensland, 2009

A thesis submitted for the degree of Doctor of Philosophy at The University of Queensland in 2014 School of Mathematics and Abstract

Mechanical oscillators are extensively used in applications ranging from chronometry using quartz os- cillators to ultra fast sensors and actuators as in atomic force microscopy and spin/mass/force sensing. The continuous push towards miniaturized high precision sensors has motivated the development of increasingly high quality mechanical oscillators at length scales as small as nanometers, with integra- tion into cryogenic environments used to avoid thermal noise. This raises the prospect of observing mechanical oscillators in a new regime where their dynamics are dictated by quantum, rather than classical, mechanics. Utilizing the quantum behaviour of mechanical oscillators promises to enhance applications in sensing and metrology. In this thesis experimental techniques are reported that en- hance optomechanical sensors and enable fundamental experiments in quantum optomechanics. In particular, there is a strong emphasis towards experimentally developing detection based feedback control techniques, for example to enable feedback cooling towards the mechanical ground state or to stabilize parasitic instabilities. In addition to these experimental advances, I detail a real-time es- timation strategy that invalidates the use of linear feedback to enhanced the performance of linear optomechanical sensors. Finally, a unique and promising optomechanical systems is developed and characterized based on surface waves of a thin film of superfluid helium-4. The first topic considered in this thesis is feedback cooling of a generalized optomechanical sys- tem. In that chapter a detailed mathematical treatment is derived that highlights the importance of using a dual probe position measurement to accurately characterize a feedback cooled oscillator. The theoretical results are then experimentally verified using a microtoroidal resonator controlled via elec- trostatic actuation. This chapter provides a brief introduction into feedback cooling and serves to in- troduce the fundamental optomechanical system that underscores the majority of the work presented in this thesis. In the following chapter, the problem of estimating an unknown force driving a linear oscillator is considered. In this context it is well known that linear feedback control can improve the performance of nonlinear mechanical sensors. However, for completely linear systems, feedback is often cited as a mechanism to enhance bandwidth, sensitivity or resolution. For such systems it is shown that as long as the oscillator dynamics are known, there exists a real-time estimation strategy that reproduces the same measurement record as any arbitrary feedback protocol. Consequently, some form of nonlinearity is required in the controller, plant, or sensor, to gain any advantage beyond esti- mation alone. This result holds true in both quantum and classical systems, with non-stationary forces and feedback, and in the general case of non-Gaussian and correlated noise. As a proof-of-principle, a specific case of feedback enhanced incoherent force resolution is experimentally reproduced using straightforward filtering on the position measurement record from an optomechanical sensor. The fundamental limits to optomechanical force sensing are defined by the well-known standard quantum limit (SQL) which is characterized by an equal contribution from two fundamental noise sources, imprecision noise in the form of optical shot noise and quantum back-action noise due to the stochastic arrival of . However, well before the SQL is reached, dynamical back-action may become sufficiently strong to severely alter the dynamics of the sensor and hence degrade the sensi- tivity. In this thesis a technique based on opto-electromechanical feedback control is proposed and

i experimentally demonstrated. The parametric-instability-induced degradation in mechanical trans- duction sensitivity is experimentally characterized, then it is shown that application of the proposed feedback technique suppresses the instability, enabling higher optical powers to be used, resulting in enhanced transduction sensitivity. In addition to dynamical backaction effects, the optical power incident on a mechanical system is typically bounded from above, for example from the thermal damage threshold characterized by intrinsic absorption. If this bound prevents the SQL from being reached, one may enhance the im- precision by surpassing the shot noise limit using quadrature squeezed light. In such a proposal, the reduced noise on one quadrature permits enhanced measurement sensitivity, for example, in a phase sensitive interferometer. While implementation of squeezed light in enhanced sensing has been demonstrated in spectroscopy, and atomic, polarization and gravitational wave interferometry its use for position measurement of mesoscopic mechanical oscillators is still largely unexplored. Here, phase squeezed light is experimentally shown to enhance the transduction sensitivity of a room tem- perature microtoroidal resonator; extending the applicability of non-classical light to the regime of micro-mechanical oscillators and potentially leading to quantum enhanced feedback. Undoubtedly, the most exciting prospects of mesoscopic mechanical systems lie in fundamen- tal quantum research where such systems could enable new quantum information technologies and experimental tests of quantum nonlinear mechanics. Among the most successful optomechanical sys- tems are the collective modes of ultra cold atoms where ponderomotive squeezing and ground state cooling have been observed. Therefore the collective motion of thin films of superfluid helium-4 also appears to be a promising candidate for quantum optomechanics given its zero viscosity flow and low effective mass. To this end, for the first time, precise optical readout and control of superfluid helium- 4 surface waves is demonstrated. Readout is achieved by evanescent coupling of the film to an optical whispering gallery mode resonator, allowing the intrinsic Brownian motion to be directly observed. Laser control is achieved by detuning from the optical cavity, resulting in a modified spring constant and dynamical heating and cooling of the superfluid mode. By weakly modulating the injected optical field the mechanical mode is also shown to exhibit a strong Duffing nonlinearity. This system, with the prospect of extremely low dissipation rates and strong mechanical nonlinearities, is a promising candidate for fundamental research into highly non-classical mechanical systems.

ii Declaration by author

This thesis is composed of my original work, and contains no material previously published or written by another person except where due reference has been made in the text. I have clearly stated the contribution by others to jointly-authored works that I have included in my thesis. I have clearly stated the contribution of others to my thesis as a whole, including statistical as- sistance, survey design, data analysis, significant technical procedures, professional editorial advice, and any other original research work used or reported in my thesis. The content of my thesis is the result of work I have carried out since the commencement of my research higher degree candidature and does not include a substantial part of work that has been submitted to qualify for the award of any other degree or diploma in any university or other tertiary institution. I have clearly stated which parts of my thesis, if any, have been submitted to qualify for another award. I acknowledge that an electronic copy of my thesis must be lodged with the University Library and, subject to the General Award Rules of The University of Queensland, immediately made available for research and study in accordance with the Copyright Act 1968. I acknowledge that copyright of all material contained in my thesis resides with the copyright holder(s) of that material. Where appropriate I have obtained copyright permission from the copyright holder to reproduce material in this thesis.

iii Publications during candidature

1. Glen I. Harris, David L. McAuslan, Eoin Sheridan, Zhenglu Duan, Warwick P. Bowen, “Laser cooling and control of excitations in superfluid helium,” submitted to Nature (2014)

2. Glen I. Harris, David L. McAuslan, Thomas M. Stace, Andrew C. Doherty, Warwick P. Bowen, “Minimum requirements for feedback enhanced force sensing,” Physical Review Letters 111 103603 (2013)

3. Glen I. Harris, Ulrik L. Andersen, Joachim Knittel, Warwick P. Bowen, “Feedback-enhanced sensitivity in optomechanics: Surpassing the parametric instability barrier,” Physical Review A 85 061802 (2012)

4. Shan Zheng Ang, Glen I. Harris, Warwick P. Bowen, Mankei Tsang, “Optomechanical param- eter estimation,” New Journal of Physics 15 103028 (2013)

5. Ulrich B. Hoff, Glen I. Harris, Lars S. Madsen, Hugo Kerdoncuff, Mikael Lassen, Bo M. Nielsen, Warwick P. Bowen, Ulrik L. Andersen, “Quantum-enhanced micromechanical dis- placement sensitivity,” letters 38 1413-1415 (2013)

6. Terry G. McRae, Kwan H. Lee, Glen I. Harris, Joachim Knittel, Warwick P. Bowen, “Cavity optoelectromechanical system combining strong electrical actuation with ultrasensitive trans- duction,” Physical Review A 82 023825 (2010)

7. Hugo Kerdoncuff, Ulrich B. Hoff, Glen I. Harris, Warwick P. Bowen and Ulrik L. Andersen, “Squeezing-enhanced measurement sensitivity in a cavity optomechanical system”, Accepted into Annalen der Physik

The following are publications from my candidature which are not included in this thesis

1. Alex Szorkovszky, Andrew C Doherty, Glen I Harris, Warwick P Bowen, “Position estimation of a parametrically driven optomechanical system,” New Journal of Physics 14 (2012)

2. Andrew C Doherty, A Szorkovszky, GI Harris, WP Bowen, “The quantum trajectory approach to quantum feedback control of an oscillator revisited,” Philosophical Transactions of the Royal Society A 370 5338-5353 (2012)

3. Alex Szorkovszky, Andrew C Doherty, Glen I Harris, Warwick P Bowen, “Mechanical squeez- ing via parametric amplification and weak measurement,” Physical Review Letters 107 213603 (2011)

4. Stefan Forstner, Stefan Prams, Joachim Knittel, Erik D. van Ooijen, Jon D. Swaim, Glen I. Harris, Alex Szorkovszky, Warwick P. Bowen, Halina Rubinsztein-Dunlop, “Cavity optome- chanical magnetometer,” Physical review letters 108 120801 (2012)

iv Publications included in this thesis

Many of the main results of this thesis have either been submitted to, or published in, peer-reviewed journals. The manuscripts that were written by myself are predominantly my own work, with sub- stantial input and improvements from my supervisor A/Prof Warwick Bowen. The contributions of coauthors to these publications are described below. In addition to these publications, the research I performed during my PhD produced a further four publications which are not included in this thesis as they are not of direct relevance. The following manuscript, and its supplementary information, was incorporated into Chapters 6: Glen I. Harris, David L. McAuslan, Eoin Sheridan, Zhenglu Duan, Warwick P. Bowen, “Laser cool- ing and control of excitations in superfluid helium,” submitted to Nature (2014) Contributor Statement of contribution Glen Harris Designed and constructed the apparatus (40%), Wrote the paper (35%), Performed experiment (40%), Performed analysis (35%) David McAuslan Designed and constructed the apparatus (40%), Wrote the paper (35%) Performed experiment (40%), Performed analysis (40%), Device fabrication (20%) Eoin Sheridan Device fabrication (80%) Zhenglu Duan Performed analysis (5%) Warwick Bowen Designed and constructed the apparatus (20%), Wrote the paper (30%) Performed experiment (20%), Performed analysis (20%)

Chapter 5 is based on the following publication: Glen I. Harris, David L. McAuslan, Thomas M. Stace, Andrew C. Doherty, Warwick P. Bowen, “Minimum requirements for feedback enhanced force sensing,” Physical Review Letters 111 103603 (2013) Contributor Statement of contribution Glen Harris Designed and constructed the apparatus (70%), Wrote the paper (60%), Performed experiment (80%), Performed analysis (50%) David McAuslan Designed and constructed the apparatus (10%), Performed experiment (10%), Performed analysis (10%) Thomas Stace Performed analysis (20%) Andrew Doherty Wrote the paper (10%) Warwick Bowen Designed and constructed the apparatus (20%), Wrote the paper (30%) Performed experiment (10%), Performed analysis (20%)

The discussion on optimal post-processing techniques in Chapter 5 is based on the following pub- lication: Shan Zheng Ang, Glen I. Harris, Warwick P. Bowen, Mankei Tsang, “Optomechanical parameter estimation,” New Journal of Physics 15 103028 (2013)

v Contributor Statement of contribution Shan Zheng Ang Performed analysis (60%), Wrote the paper (55%) Glen Harris Designed and constructed the apparatus (80%), Performed experiment (80%), Wrote the paper (5%) Warwick Bowen Designed and constructed the apparatus (20%), Performed experiment (20%) Mankei Tsang Performed analysis (40%), Wrote the paper (40%)

Chapter 4 is based on the following publication: Glen I. Harris, Ulrik L. Andersen, Joachim Knittel, Warwick P. Bowen, “Feedback-enhanced sensi- tivity in optomechanics: Surpassing the parametric instability barrier,” Physical Review A 85 061802 (2012) Contributor Statement of contribution Glen Harris Designed and constructed the apparatus (40%), Wrote the paper (60%) Performed experiment (40%), Performed analysis (70%), Device Fabrication (20%) Ulrik Andersen Designed and constructed the apparatus (20%), Performed experiment (40%), Performed analysis (10%) Joachim Knittel Device Fabrication (80%), Designed and constructed the apparatus (20%) Warwick Bowen Designed and constructed the apparatus (20%), Wrote the paper (40%) Performed experiment (20%), Performed analysis (20%)

Chapter 7 is based on the following publication: Ulrich B. Hoff, Glen I. Harris, Lars S. Madsen, Hugo Kerdoncuff, Mikael Lassen, Bo M. Nielsen, Warwick P. Bowen, Ulrik L. Andersen, “Quantum-enhanced micromechanical displacement sensitiv- ity,” Optics letters 38 1413-1415 (2013) Contributor Statement of contribution Ulrich Hoff Designed and constructed the apparatus (20%), Wrote the paper (60%) Performed experiment (40%), Performed analysis (50%), Developed the concept (10%) Glen Harris Designed and constructed the apparatus (20%), Device Fabrication (100%), Performed experiment (40%), Performed analysis (50%), Developed the concept (10%) Lars Madsen Designed and constructed the apparatus (20%), Performed experiment (5%) Hugo Kerdoncuff Wrote the paper (10%) Mikael Lassen Designed and constructed the apparatus (20%), Performed experiment (5%) Bo Nielsen Designed and constructed the apparatus (10%), Warwick Bowen Developed the concept (40%), Ulrik Andersen Designed and constructed the apparatus (10%), Developed the concept (40%), Performed experiment (10%), Wrote the paper (30%)

vi The theory discussed in Chapter 7 is based on discussions had while performing the related ex- periments. Those discussions resulted in the following publication: Hugo Kerdoncuff, Ulrich B. Hoff, Glen I. Harris, Warwick P. Bowen and Ulrik L. Andersen, “Squeezing- enhanced measurement sensitivity in a cavity optomechanical system”, Accepted into Annalen der Physik (2014) Contributor Statement of contribution Hugo Kerdoncuff Wrote the paper (70%), Developed the concept (20%) Ulrich Hoff Wrote the paper (10%), Developed the concept (20%) Glen Harris Developed the concept (20%) Warwick Bowen Developed the concept (20%), Ulrik Andersen Wrote the paper (20%), Developed the concept (20%),

The theory presented in Chapter 3 is research done early in my PhD based around the following publication: Terry G. McRae, Kwan H. Lee, Glen I. Harris, Joachim Knittel, Warwick P. Bowen, “Cavity op- toelectromechanical system combining strong electrical actuation with ultrasensitive transduction,” Physical Review A 82 023825 (2010) Contributor Statement of contribution Terry McRae Designed and constructed the apparatus (40%), Device Fabrication (100%), Performed experiment (60%), Performed analysis (70%), Wrote the paper (60%) Kwan Lee Designed and constructed the apparatus (20%), Performed experiment (20%) Glen Harris Designed and constructed the apparatus (20%), Performed experiment (20%) Joachim Knittel Designed and constructed the apparatus (10%) Warwick Bowen Designed and constructed the apparatus (10%), Performed analysis (20%), Wrote the paper (40%)

Contributions by others to the thesis

Apart from the original motivation, my supervisor A/Prof Warwick Bowen has provided support in the analysis and interpretation of data as well as technical support and revision of written material. The experiments in Chapter 7 were possible only because of the contributions of our collaborators at the Technical University of Denmark, as described above. The figures in that chapter are my own modified versions of those presented in the associated paper, which was written primarily by Ulrich Busk Hoff. The figures in Chapter 3 are derived from research I performed immediately prior to starting my PhD. However, the written work in that chapter represents research done during the early stages of my PhD.

vii Statement of parts of the thesis submitted to qualify for the award of another degree

None

viii Acknowledgements

I would like to firstly thank my supervisor, A/Prof. Warwick Bowen, for all the guidance he provided and the strong enthusiasm he maintained. Over the course of my PhD he patiently taught me a great deal about experimental and , for which I am forever grateful. The rapid growth of the Queensland Quantum Optics group over the last 5 years is a true testament to his vision, leadership and friendly nature. I was incredibly fortunate to be a part of such a brilliant group of people, and truly appreciated the time shared with Eoin Sheridan, Michael Taylor, Terry McRae, James Bennett, Alex Szorkovszky, Jon Swaim, George Brawley, Michael Vanner, Stefan Forstner, Kiran Khosla, Lars Madsen and Andreas Naesby. A special thanks to David McAuslan for helping me exorcise the cryostat demons that are so persistent in obscuring interesting physics at cryogenic temperatures. The experiments using squeezed light in this thesis were only possible with the help of our collab- orators at the Technical University of Denmark. In this regard, I would like to thank Ulrik Andersen and all the group members for giving me a warm welcome and making my stay in Denmark most enjoyable. I’m particularly grateful for the time spent with Ulrich Busk-Hoff, who has a uniquely calming influence when experiments aren’t going to plan. For the regular climbing adventures I’ve had over the years I would like to thank David McAuslan and Martin Ringbauer. You two encouraged me to push my limits on what some may call highly ques- tionable rock. In my opinion, there’s no better way to build friendships and develop an appreciation for the subtle things in life. As always, my family were a reliably constant source of unconditional support and guidance during my PhD. I consider myself very lucky to have been raised in such a loving and diverse family. Finally, a very special thank you to my delightful wife Kirsty, whose compassionate nature kept me leveled and focused throughout the years. My debt to you is far greater than you know.

ix Keywords

Quantum optomechanics, Feedback control, Superfluid Helium.

Australian and New Zealand Standard Research Classifications (ANZSRC)

ANZSRC code: 020604, Quantum Optics, 50% ANZSRC code: 020699 Quantum Physics not elsewhere classified, 50%

Fields of Research (FoR) Classification

FoR code: 0205, Optical Physics, 50% FoR code: 0206, Quantum Physics, 50%

x Science does not know its debt to imagination. Ralph Waldo Emerson

xi Contents

List of Figures xiv Glossary of terms ...... xix

1 Thesis Overview 1 1.1 Thesis Structure ...... 2

2 Introduction to Quantum Optomechanics 3 2.1 ...... 3 2.2 ...... 4 2.3 Hamiltonian Formalism ...... 6 2.3.1 Quantum Langevin Equations ...... 9

3 Feedback Cooling and Temperature Estimation 11 3.1 Introduction ...... 11 3.2 Limits to Feedback Cooling ...... 12 3.2.1 Temperature Estimate Using Feedback Photocurrent ...... 16 3.2.2 Temperature Estimate Using an Independent Photocurrent ...... 18 3.3 Experimental Design ...... 18 3.4 Experimental Results ...... 22 3.5 Conclusion ...... 23

4 Feedback Suppression of Parametric Instability 26 4.1 Instabilities in Optomechanics ...... 26 4.2 Theoretical Description of Parametric Instability ...... 28 4.3 Feedback Suppression of Parametric Instability ...... 31 4.4 Conclusion ...... 33

5 Minimum requirements for feedback enhanced force sensing 35 5.1 Introduction to Force Sensing ...... 35 5.2 Theoretical Treatment of Stationary Dynamics ...... 37 5.2.1 A Heuristic Approach to Stationary Dynamics ...... 39 5.3 Theoretical Treatment of Non-Stationary Dynamics ...... 40 5.4 Force Sensing in the Presence of Nonlinearities ...... 45

xii 5.5 Experimental Results: Stationary Dynamics ...... 46 5.6 Conclusion ...... 49

6 Superfluid Optomechanics 50 6.1 Superfluid Helium-4 ...... 50 6.1.1 Hydrodynamic Equations of Thin Films ...... 51 6.2 Superfluid Optomechanics with Thin Films ...... 52 6.2.1 Predicted Optomechanical Parameters for Thin Film Systems ...... 54 6.3 Measuring Superfluid Film Fluctuations ...... 56 6.3.1 Experimental Design: Minimizing Vibration ...... 58 6.4 Laser Control of Superfluid Film Fluctuations ...... 59 6.4.1 Identifying Optomechanical Interactions ...... 59 6.4.2 Theoretical Treatment: Photothermal and Radiation Pressure Optomechanics 61 6.4.3 Observing and Characterizing Dynamical Backaction ...... 64 6.5 Duffing Nonlinearity ...... 66 6.5.1 Theoretical Treatment ...... 66 6.5.2 Experimental Observation ...... 68 6.6 Conclusion ...... 69

7 Squeezed Light For Enhanced Transduction Sensitivity 70 7.1 Quantum Enhanced Sensing ...... 70 7.1.1 Squeezed Light in Cavity Optomechanics ...... 71 7.2 Experimental Results ...... 72 7.3 Conclusion ...... 76

8 Conclusion 77 8.0.1 Other Work ...... 78 8.1 Future work ...... 78

References 80

xiii List of Figures

2.1 Fabry-Perot cavity with a mechanically compliant mirror. The mechanical (optical)

resonance frequency is given by ωm (ω) with dissipation rate Γm (γ)...... 5

3.1 Schematic of feedback cooling model including an independent out-of-loop probe field denoted asa ˆ(2). The in-loop probe fielda ˆ(1) is used for feedback...... 13

3.2 Theoretical temperature prediction whilst feedback cooling. A Out-of-loop temper- ature prediction whilst cooling from room temperature. Dotted lines show current capabilities of SNR and gain. B In-loop temperature prediction whilst cooling from room temperature showing squashing effect corresponding to unphysical negative temperature.C Feedback cooling with cryogenic precooling to 300mK...... 17

3.3 A Scanning electron microscopy image of a microtoroidal resonator B & C Finite el- ement simulations of microtoroidal whispering gallery optical modes and mechanical modes, respectively...... 19

3.4 Experimental schematic including electronic locking method for in-loop probe. The out-of-loop probe was thermally locked. FPC: fibre polarization controller. All ex- periments were carried out at room temperature (T = 300 K) and atmospheric pressure. 20

3.5 Mechanical spectrum at room temperature showing three mechanical modes and their corresponding FEM simulations. The frequency axis is centered about 6.272 MHz. The shaded region identifies the mechanical mode chosen for cooling and the inset is a magnification of this mode with transduction sensitivity (grey) and electronic dark noise (black) ...... 21

3.6 Feedback cooling with varying feedback gain (85). A and B: In-loop and out-of-loop transduction spectra respectively. C: Temperature inferred using in-and out-of-loop transduction signals. The red circles  and blue squares  respectively denote out- and in-loop temperature inferences; the solid red curve and dashed blue curve respectively denote the theoretical predictions of actual mechanical oscillator temperature and inferred in-loop temperature. Figure from Ref. (85) ...... 25

xiv 4.1 Ultra sensitive optomechanical systems exhibit parametric instability through cavity enhancement of the Stokes sideband produced by their mechanical motion. A In de- tuned micron scale optomechanical systems Stokes and anti-Stokes sidebands spec- trally overlap with the same optical mode. B In large-scale interferometers Stokes sidebands must overlap with an adjacent optical mode. Image of LIGO Livingston Laboratory courtesy of Skyview Technologies ...... 27

4.2 A Experimental schematic. Blue (light grey) indicates the optical components allow- ing transduction and parametric instability; Green (dark gray) indicate the electrical components used in the feedback stabilization. FPC: Fiber polarization controller. BPF: Bandpass filter. B Observed mechanical spectra below parametric instability threshold...... 30

4.3 Mechanical power spectra B and A show, respectively, the mechanical spectra with and without feedback stabilization of the regenerative 14MHz mode...... 31

4.4 A Mechanical spectra observed with 136 µW of input power, well above the paramet- ric instability threshold, with (Grey) and without (Black) feedback stabilization of the unstable 14 MHz mode. Inset: Single sided noise spectrum about the second har- monic 2U (solid line) showing the measured scaling of laser frequency noise (dashed line) and the suppression of noise from thermal locking (grayed region)B Signal to noise ratio of the 28.6 MHz mechanical mode versus optical power with feedback, circles, and without, squares ...... 34

5.1 A Conceptual diagram comparing optomechanical force sensing with feedback and filtering B Experimental schematic. Red (dark grey): fiber interferometer; solid green (light gray): electrical components for feedback stabilization; dashed green: electrical signal applied to the microtoroid for application of the incoherent force and feedback. FPC: fiber polarization controller. BPF: bandpass filter. MZM: Mach-Zender Modu- lator. PM: phase modulator. PI: proportional-integral controller. Figure from Ref. (66) 37

5.2 A Theoretical thermal displacement spectrum (red) with an incoherent signal (blue). B-C Stationary filtering is applied to the thermal spectra with increasing gain. Solid lines indicate the mechanical susceptibility and dashed lines indicate the noise floor subject to the same filtering process as the measurement record. The area of the shaded regions, denoting the amount of extractable information about the thermal noise and the signal, is independent of the filtering gain. The signal is greater than the combined noise floor from thermal noise and shot noise over the region between the two vertical grey lines, with the separation of these lines giving the bandwidth of the measurement. Figure from supplementary information of Ref. (66) ...... 40

5.3 General theoretical schematic outlining the emergence of forces and feedback in a cavity optomechanical system. Figure from supplementary information of Ref. (66) . 41

xv 5.4 A Displacement spectrum at room temperature [dark grey (red)] and with filtering [light grey (green)] at gain 2, 8, 34, 72 and 150. Inset top: displacement spectrum with feedback cooling [black (blue)] and filtering [light grey (green)] with gain 1. Inset bottom: percentage difference between filtered and cooled spectrum with equivalent gain. B Force resolution as a function of averaging duration for thermal [dark grey (red)], feedback cooled [black (blue)] and filtered spectra [light grey (green)] at gain 1, 2.4, 5 and 10. Solid line (red): fit to inverse quartic dependence of the force resolution without filtering or feedback. C Force resolution versus filter gain after 1 ms of averaging (green circles) showing good agreement to theory (dashed line). Figure from Ref. (66) ...... 47 5.5 A Normalised energy estimate at room temperature [light grey (red)], and with addi- tion of incoherent driving [dark grey (green)]. Solid lines: mean; dashed lines: one standard deviation error bounds. B Averaging time required to resolve incoherent sig- nal versus filter gain (green circles) showing good agreement to theory (dashed line). Figure from Ref. (66) ...... 48

6.1 Optomechanics with superfluid helium films. (a) Illustration of third sound waves of a superfluid helium film coating a microtoroid. The third sound oscillations have p −3 21 5 −2 phase velocity C3 = 3αvdwd , where αvdw = 2.65 × 10 nm s is the van der Waals coefficient for silica. (b) Shot-noise limited homodyne detection of superfluid helium optomechanics. A tapered fibre-coupled microtoroid is orientated face down inside the sample chamber of a helium-3 cryostat. Helium-4 gas is injected into the sample chamber at a pressure of 69 mTorr (at 2.8 K). AM - amplitude modulator, BS - beamsplitter, D - photodetector, NA/SA - network/spectrum analyser. The 90/10 BS probes the optical transmission through the microtoroid...... 53 6.2 Measurement of superfluid helium mechanical motion. a, Observation of superfluid oscillations as the cryostat cools to base temperature, while the laser frequency is scanned and optical resonance tracked. Oscillations that result from optomechanical parametric instability are observed with as little as 40 nW of optical power, over one hundred times lower than would be expected from a microtoroid mechanical mode (75). Blue, orange, and green traces are offset vertically and were respectively taken at 3 K, 1 K, and 0.6 K. b, Resonance frequency of the fundamental major mode versus cryostat temperature. Solid line is obtained by modelling the condensation of the helium gas, enabling the film thickness to be calculated as a function of tem- perature. c, Network analysis at 900 mK showing the low frequency mode (orange shading) that experiences instability and a set of high frequency superfluid modes (blue shading). d, Quality factor of the high frequency superfluid mode indicated by the arrow versus cryostat temperature. e, Spectrum of thermal excitations of the high frequency mode at 330 kHz, with 48 nW of power incident on the microtoroid and at 600 mK. Measurements were performed on a microtoroid with optical decay rate κ/2π = 35.6 MHz for an optical mode at λ = 1560.0 nm...... 57

xvi 6.3 On resonance behaviour. Optical power dependence of the mechanical properties of mode A (blue circles) and mode B (orange circles) when the laser is locked to the cavity resonance (i.e zero detuning). (a, b) Measured thermal energy of the superfluid mode versus input laser power. The blue line is a power-law fit to mode A with scaling ∝ P0.6. The orange line is the fit to mode B with scaling ∝ P0.69.(c) Mechanical

linewidth (Γm/2π) versus optical power. The blue line is a power-law fit given by 0.38 0.14 Γm/2π = 16.0 × P + 19.3 and the orange line is the fit Γm/2π = 49.2 × P + 23.4.

(d) Shift in mechanical resonance frequency (∆ωm/2π) versus laser power...... 60 6.4 Optomechanical heating and cooling of two third sound modes. Detuning the op- tical field results in cooling (a) or heating (d) of the superfluid excitations. The laser is red-detuned with respect to the cavity for both the 552 kHz mode and the 482 kHz mode. Whether the mode experiences heating or cooling depends on the relative magnitude of the radiation pressure forces and the sign of the photothermal constant (β), which itself depends on the optical mode. (a) Spectra of two closely spaced me- chanical modes around 552 kHz with varying optical detuning. The shaded mode experiences photothermal induced broadening (β > 0), with a mechanical linewidth

of Γm/2π = 77 Hz (∆/κ = 0), Γm/2π = 334 Hz (∆/κ = −0.25) and Γm/2π = 457 Hz (∆/κ = −0.5). (b-c) Change in mechanical damping rate and mechanical resonance frequency as a function of detuning for the 552 kHz mode. Solid line: theoretical fit to photothermal broadening. (d) Spectra of two closely spaced mechanical modes around 482 kHz with varying optical detuning. The shaded mode experiences pho-

tothermal induced narrowing (β < 0), with a mechanical linewidth of Γm/2π = 115 Hz

(∆/κ = 0), Γm/2π = 63 Hz (∆/κ = −0.25) and Γm/2π = 36 Hz (∆/κ = −0.5). (e-f) Change in mechanical damping rate and mechanical resonance frequency as a func- tion of detuning for the 482 kHz mode. Solid line: theoretical fit to photothermal narrowing. Traces in a and d are offset for clarity. All detuning measurements were taken with 200 nW of launched optical power...... 65

6.5 Variations of the potential energy landscape of a Duffing oscillator showing stable (o) and unstable (+) fixed points from the combination of quadratic (dotted line) and quartic (dashed line) terms...... 67

6.6 Duffing non-linearity. Network analyser traces as the laser is amplitude modulated at a frequency that is swept over the mechanical resonance. Solid lines: fit to data using the standard model of a Duffing oscillator...... 69

7.1 Combined sideband and phase space representation of the optomechanical transduc- tion mechanism. The sidebands are initially prepared in a composite two mode squeezed state with the individual sidebands being in a thermal state (61). The optomechanical phase modulation scatters photons from the coherent carrier field

(green) into the sidebands, displacing the initial quantum states δa± (red). Figure derived from (69) ...... 73

xvii 7.2 Schematic diagram of the experimental setup. MC: mode cleaning cavity. PZT: piezo- electric transducer. PBS: polarizing beam splitter. LO: local oscillator field. AL: aspherical lens. OPA: optical parametric amplifier. Inset: SEM micrograph of the mi- crotoroid with major diameter 60 µm and minor diameter 6 µm. Figure derived from (69) ...... 74 7.3 A Characterization of the transmitted vacuum squeezed state by modulating the local oscillator phase. All traces were recorded in zero span mode at a detection frequency of 4.9 MHz, RBW 100 kHz, and VBW 100 Hz. The detector dark noise was measured to be 25 dB below shot noise and has not been subtracted. B Power spectral den- sity showing mechanical modes of a microtoroidal resonator measured with squeezed light (green), yielding a transduction sensitivity reduced below the SNL (black). All traces are averaged over 10 samples, recorded with RBW 10 kHz and VBW 100 Hz. Figure derived from (69) ...... 75

xviii Glossary of terms

Term Definition CCD Charge-coupled device; a camera EOM Electro-optical modulator; a device which modulates the phase of light AM Amplitude modulator; a device which modulates the amplitude of light PDH Pound-Drever-Hall; a technique to lock phase and frequency PID Proportional-Integral-Differential controller used for locking LO Local oscillator; SNR Signal-to-noise ratio QNL Quantum noise limit; The limit to sensitivity which is set by quantum noise, also known as shot noise SQL Standard quantum limit; the absolute quantum limit to the achievable sensitivity given the number of photons used OPA Optical parametric amplifier; the squeezed light source PBS Polarising beam splitter RF Radio frequency hAi The expectation value of an arbitrary variable A δA Fluctuations of an arbitrary variable A FT Thermal force ηd Detection efficiency λ The optical wavelength † aˆn anda ˆn Annihilation and creation operators of the optical field † bˆn and bˆn Annihilation and creation operators of the mechanical field Xˆ + Amplitude quadrature of the field Xˆ − Phase quadrature of the field i Detected photocurrent

Term Definition Homodyne A measurement method whereby a signal field is measured by interfering it with a bright local oscillator Vacuum fluctuations The zero-point fluctuations of electromagnetic fields Coherent state The state of light produced by a laser, with amplitude and phase noise at the level of the vacuum fluctuations Squeezed state A state of light for which the fluctuations along one quadrature are smaller than the vacuum fluctuations

xix Chapter 1

Thesis Overview

This thesis reports developments towards the preparation, readout and control of micromechanical oscillators in the quantum domain. The primary focus of this thesis involves the incorporation of micromechanical systems into a helium-3 cryostat, resulting in a novel optomechanical system that utilizes the surface waves of evanescently coupled thin films of superfluid helium-4. In addition to cryogenic operation a significant portion of research was performed at room temperature concerning topics such as limitations and applications of classical feedback control and the use of non-classical light in optomechanical systems. The primary optomechanical system used during my PhD are whispering gallery mode micro- toroidal resonators. Fabrication of such devices uses a combination of photo-lithography, chemical etching and a selective reflow process to produce silica microtoroids with ultra-low optical loss. Sur- face tension induced, wafer-based resonators, are ideal due to scalability, reproducibility and the potential for parallelization of cavities. In all experiments, light is coupled into the optical mode with 99% efficiency using a tapered optical fiber placed close to the microtoriod. The research performed during my PhD begins in Chap. 3, where feedback cooling of a mechan- ical oscillator using a combination of electrical gradient forces and ultrasensitive optical transduction is presented. This chapter details the mathematical treatment of a continuous position measurement, made from two independent transducers, of a mechanical oscillator experiencing feedback cooling generated from one of the transduction signals. Theoretical predictions, such as squashing of the transduction signal and cooling limited by the signal to noise ratio, is in good agreement with experi- mental results. The observations highlight the importance of a second transduction signal for accurate characterization of the oscillator whilst feedback cooling. In Chapter 4, strong dynamical back-action effects that severely alter the dynamics of optome- chanical sensors, and hence degrade the sensitivity, are considered. In particular, a technique based on opto-electromechanical feedback control is proposed to suppress parametric instability. The theory is experimentally verified by characterizing the parametric-instability-induced degradation in, and feed- back induced revival of, mechanical transduction sensitivity. Furthermore, in Chap. 7, phase squeezed light is shown to enhance the transduction sensitivity of a room temperature microtoroidal resonator; extending the applicability of non-classical light to the regime of micro-mechanical oscillators and potentially leading to sub SQL measurements and quantum enhanced feedback.

1 Chapter 5 considers feedback and post-processing techniques for the generalized problem of es- timating an unknown force driving a linear oscillator. In this context, it is well known that linear feedback control can improve the performance of nonlinear mechanical sensors. However for stro- boscopic measurements and variance estimation it is not well known if feedback can enhance the performance of sensors beyond that achievable with estimation alone. In that chapter, a straightfor- ward theoretical approach is developed which shows that, even in the presence of non-Gaussian or correlated noise and non-stationary processes, a real-time estimation strategy always exists that repro- duces the same measurement record as any arbitrary linear feedback protocol. It ultimately provides a clear set of minimum requirements, namely some form of nonlinearity in the controller, plant or sensor, for feedback to provide any advantage to force sensing over that possible with estimation alone. Chapter 6 focuses on surfaces waves of superfluid helium-4 residing on the surface of the micro- toroid. In that chapter, for the first time, precise optical readout and control of superfluid helium-4 surface waves is experimentally demonstrated. Ultra-precise readout is achieved by evanescent cou- pling of the film to an optical whispering gallery mode resonator, allowing the intrinsic Brownian motion of the mechanical modes of the film to be directly observed. Detuned optical probing enables laser control of the mechanical susceptibility, resulting in a modified spring constant and dynamical heating and cooling of the superfluid mode. Furthermore, weakly modulating the injected optical field drives the mechanical mode into a strongly nonlinear regime which is shown to be described by the Duffing oscillator. While there are still many open questions regarding the superfluid hydrodynamics this system is a promising candidate to study macroscopic non-classical mechanical states.

1.1 Thesis Structure

The introduction chapter (Chapter 1) of this thesis provides a brief overview to the field optomechan- ics with a historical perspective, ultimately providing a motivating argument for pursuing research in quantum optomechanics. However, it is not the purpose of this chapter to provide an exhaustive review of the field. For such texts the reader is directed to (14, 109, 111). The subsequent chapters (Chapters 2-6) report the key research projects performed during my PhD. The chapters are designed to be independent and self-contained so that the reader may under- stand the aims and achievements relevant to the chapter without the need to read any other part of this thesis.

2 Chapter 2

Introduction to Quantum Optomechanics

2.1 Radiation Pressure

Radiation pressure is broadly described as the force per unit area resulting from the reflection or absorption of electromagnetic radiation. Due to the ubiquitous nature of such radiation, this simple process has been shown to mediate a range of physical phenomena spanning all length and mass scales, from the optical interaction of single atoms to the formation of interstellar bodies (14, 109, 111). While radiation pressure was first conceptualized by the famous astronomer Johannes Kepler in 1619 to explain the natural orientation of a comet’s debris trail away from the sun, for over 200 years it remained unexplored until James Maxwell theoretically predicted the momentum carried by a radiation field in 1862 (99, 115). This was shortly followed by experimental observations by Ernest Nichols and Gordon Hull 1901 using the Nichols radiometer (115) (not to be confused with the Crookes radiometer invented around the same time, which relies on gas ballistics). The advent of in the early 1900s verified Maxwell’s predictions of the momentum carried by light. However, it wasn’t until the 1970s and onwards that the exquisite control afforded by radiation pressure was fully realized in the form of trapping and cooling of single atoms and molecules (13, 105). Such experiments enhanced the understanding of atomic clocks, atomic collision processes, and diffraction and interference of atomic beams, with the most notable experiment confirming the “condensation” of a gas of weakly interacting bosons, also known as Bose-Einstein condensation (13, 45, 105). Before detailing the conceptual foundation of optomechanics, it should be noted that its primary feature is the exploitation of radiation pressure to precisely control the dynamics of macroscopic me- chanical oscillators. In this context, and similar to the advances made in atom optics, optomechanics aims to test fundamental laws of quantum mechanics at large length and mass scales (14). Further- more, owing to the widespread use of mechanical oscillators as sensors in applications, investigating the intrinsic quantum dynamics of mesoscopic systems provides direction towards the next generation of sensor technologies (106). In the next section the generalized optomechanical system is formally described. Attention is directed, not only towards radiation pressure, but also on the ultra-precise readout facilitated by the characteristics of laser light, namely short wavelength and spatial and tem- poral coherence.

3 2.2 Cavity Optomechanics

The quintessential cavity optomechanical system, as seen in Fig. 2.1, consists of a Fabry-Perot´ cavity with a movable mirror that contributes an additional degree of freedom in the form of a mechanical oscillator. This conceptually simple picture is purely for illustrative purposes and does not convey the potential breadth of relevance; where optical and mechanical components may range in characteristic size and resonance frequency over many orders of magnitude. As will be formally described in the next section, the optomechanical coupling fundamentally arises from modifications to the optical resonance frequency from motion of the mirror. While in principle an optical cavity is not strictly necessary, recycling the intracavity photons effectively boosts the typically weak interaction between mechanical and optical degrees of freedom. In this configuration, the intracavity optical field serves to both drive the mechanical component via radiation pressure and measure its motion by subsequent amplitude or phase detection. Until recently, cavity optomechanics was primarily considered in the context of gravitational wave detection, spurred by the pioneering work of V. Braginsky in 1970 (23). In gravitational wave obser- vatories, kilometer scale arms of a Michelson interferometer are designed to measure the minuscule length dilation associated to gravitational waves traversing the interferometer arms (1, 83). Owing to the intrinsic Poissonian statistics of coherent laser light, it was well known that increasing the mean intracavity number reduces the measurement imprecision; commonly known as optical shot noise (16). However, it was shown that radiation pressure from the stochastic arrival of photons re- flected from the mirror causes mechanical vibrations that could potentially obscure the gravitational signal (29). This phenomena, known as quantum backaction, opposes the enhanced measurement im- precision associated with increased intracavity photon number. Balancing these two distinct effects results in the minimum achievable measurement imprecision, referred to as the standard quantum limit (SQL) (23, 29). Critically, the standard quantum limit is only defined in the context of probing with coherent states of light (23). The use of nonlinear optical crystals enables photon statistics to be modified from the characteristic Poissonian distribution, generating non-classical light and enabling control over the level of optical shot noise and quantum backaction (16, 59). Due to experimental challenges, there are currently very few demonstrations of optomechanics with non-classical light, with the only example in micromechanical systems discussed and experimentally demonstrated in Chap. 7. In contrast, a large scale example is that of the gravitational wave interferometer GEO600, which uses non-classical light to modify the optical shot noise and enhance measurement imprecision (38). Since both experiments were operating far from the SQL, the primary aim was to surpass the experimentally relevant optical shot noise. As will be discussed later, this results in an equivalent increase of quantum backaction, hence maintaining the absolute value of the SQL. Despite this, it has been theoretically shown that in a specific configuration the SQL may be surpassed with non-classical light (117), although this is yet to be experimentally demonstrated. In addition to the research on quantum limits to measurement imprecision, significant theoreti- cal and experimental work has been focused towards the preparation of non-classical motion in op-

4 κ Γ ω ω

Figure 2.1: Fabry-Perot cavity with a mechanically compliant mirror. The mechanical (optical) resonance frequency is given by ωm (ω) with dissipation rate Γm (γ) tomechanical systems (14). A notable theoretical proposal utilizes the radiation pressure of a single intracavity photon to prepare the mechanical oscillator into a superposition state of two oscillation am- plitudes (22, 98), while other proposals consider teleportation as a means of generating non-classical states (97). If experiments could realize such exotic mechanical states, combined with measurements at or beyond the SQL, fundamental tests of quantum mechanics may be achieved that are inaccessible to other established research fields such as quantum optics, , trapped atoms and cavity quantum electrodynamics. For example, quantum optomechanics could allow probing of unconventional decoherence processes like gravitational self-interaction, whereby gravity mediates collapse of spatial superposition states defined by the mechanical wavefunction (123). While exper- iments to test gravitational decoherence processes are still far from realization, the field of quantum optomechanics continues to rapidly advance, with next-generation optomechanical systems promising greatly enhanced capabilities. In parallel with advances to fundamental science, micromechanical systems for detecting accel- eration, mass and charge (106) are being developed commercially and implemented on an industrial scale. This is most evident in current technologies employing time-keeping and/or navigation with electromechanical accelerometers and chronometers. While current applications typically prefer elec- tromechanical sensors, advances in technology and fabrication techniques over the last decade have lead to optomechanical devices approaching quantum-limited measurement imprecision. Examples include photonic-crystal accelerometry (79), cavity enhanced atomic force microscopy (94), optome- chanical force (56) and mass sensors (93), and optomechanical magnetometry (53). Furthermore, op- tomechanical systems have a fundamental advantage over their electromechanical counterparts arising from the massive discrepancy in characteristic wavelength, with micrometer/nanometer optical wave- lengths enabling diffraction limited measurements at a similar scale, in contrast to electrical systems that are limited to centimeter/millimeter wavelengths. Furthermore, thermal noise is virtually non- existent in optical fields compared to electrical systems that typically require cryogenic refrigeration. Consequently, fundamental research towards precise readout and control over the quantum dynamics of optomechanical systems may enable the next generation of ultra-precise sensors. Although cavity optomechanics was initially born from gravitational wave interferometry it is ac- tively researched in a number of microscopic systems including Fabry-Perot cavities with a movable

5 mirror (63) and membrane-in-the-middle (67), whispering gallery mode cavities (141), suspended photonic crystals (48), optically levitated beads (18, 33) and the collective motion of cold atoms (128). With recent observations of the quantum motion of mechanical resonators (25, 116, 129, 140, 148), there are many exciting possibilities to simultaneously probe fundamental physics and advance appli- cations in sensing.

2.3 Hamiltonian Formalism

As introduced in the previous section, the most direct realization of cavity optomechanics is a de- formable Fabry-Perot cavity with a small movable mirror as seen in Fig. 2.1. The optical cavity has πcm length L and resonance frequency ωc = L , where m is the optical mode number. In this configuration the optical cavity is often called single sided, since one of the mirrors, in this case the rigid mirror, is partially reflective to allow driving of the cavity from an incident external optical field. Throughout this thesis the intracavity cavity field, described by the field operator a where [a, a†] = 1, is excited. As the smaller, mechanically compliant mirror moves from its equilibrium position by an amount x 0 the modified optical resonance frequency ωc is given by πc ω0 = m (2.1) c L + x ω ≈ ω − c x (2.2) c L where the approximate solution is only valid in the case of small displacement compared to the cavity length. At this point, it is possible to formulate an optomechanical Hamiltonian in terms of the center of mass position and momentum. However, to align the representation of mechanical and optical degrees of freedom in terms of raising and lowering operators, and to facilitate discussions on the nature of the optomechanical interaction, the mechanical motion is quantized. This quantization is well known (14, 111, 117) and given by s ~ † † x = (b + b ) = xzpm(b + b ) (2.3) 2meffωm r ~m ω p = − eff m i(b − b†) = −p i(b − b†) (2.4) 2 zpm where the mechanical field operators b have commutation relations [b, b†] = 1. The intrinsic quantum q ~ fluctuations of the mechanical oscillator are quantified by the zero point motion xzpm = 2m ω q eff m ~meff ωm and zero point momentum pzpm = 2 , where meff is the effective mass and ωm is the angular mechanical resonance frequency. The zero point motion can be interpreted as the standard deviation of the mechanical displacement when the oscillator is in its ground state. The modified optical resonance

6 frequency is then given by

ωc xzpm ω0 = ω − (b + b†) (2.5) c c L † = ωc − g0(b + b ) (2.6)

ωc xzpm where the g0 = L is the vacuum optomechanical coupling rate that gives the shift in optical reso- nance frequency given a displacement equal to the zero point motion. Additionally, g0 also quantifies the mechanical perturbation induced per photon from optical radiation pressure.

Considering the simple example of the dynamic coupling between a single optical and mechanical mode, the system Hamiltonian is given by the sum of the energy contributions from each harmonic oscillator,

H = Hcav + Hmech (2.7)  †  † † = ~ ωc − g0(b + b ) a a + ~ωmb b (2.8)

It should be noted that the complete Hamiltonian of the system and “bath” will include additional terms for dissipation, quantum fluctuations, and optical driving. However, as will be discussed in the next section, these contributions are typically accounted for when formulating the quantum Langevin equations with input-output relations (14, 55). From Eq. 2.8 it can be seen that the unperturbed † † energies of the optical and mechanical systems are Hcav = ~ωca a and Hmech = ~ωmb b respectively, † † and the optomechanical coupling arises through the interaction term HI = ~g0(b + b )a a. In general, the energy distribution for both the phononic and photonic modes must obey Bose-Einstein statistics, giving a mean occupationn ¯ of

 −1 n¯ = e~ωm/kbT − 1 (2.9) k T ≈ B (2.10) ~ω where the approximation is only valid in the high temperature limit i.e largen ¯. From this equation 7 the room temperature occupancy of the mechanical oscillator is found to ben ¯m ≈ 10 for a 1 MHz oscillator. This thermal noise acts to obscure the intrinsic quantum fluctuations that exist at the single phonon level. In contrast, the optical field is quantum limited with an occupancy ofn ¯ < 10−10, which may be utilized for precise readout or as a cold reservoir to cool the mechanics. These topics will be further discussed towards the end of this section.

In its current form, the Eq. 2.8 has three characteristic frequencies that typically differ by many orders of magnitude, corresponding to the optical and mechanical resonances and the coupling be- tween them. As a result, transforming into a frame rotating at the incident laser frequency ωL is made † using the unitary operator U = eiωLa at, which, when applied to the state ket, gives

|ψinew = U|ψiold (2.11)

7 Substituting Eq. 2.11 into the Schrodinger¨ equation then yields a new Hamiltonian given by

∂U H = UH U† + i~ (2.12) new old ∂t † † † † = ~∆a a + ~ωmb b − ~g0(b + b )a a (2.13)

where ∆ = ωc −ωL is the detuning between the incident laser field and the cavity resonance. While the transformed Hamiltonian still retains three characteristic frequencies, it can be seen from Eq. 2.13 that with ∆ = ±ωm the system Hamiltonian is equivalent to two harmonic oscillators of similar frequency that can exchange energy via the optomechanical interaction. While the nonlinear nature of the optomechanical interaction (Eq. 2.13) enables the generation of non-classical states (14), all experiments to-date have operated in the “weak coupling” regime where such nonlinear effects are negligible. As such, linearizing the interaction greatly simplifies mathematical derivations while still accurately describing the system dynamics. This is achieved by 2 making the substitution a = α + δa and neglecting fluctuations of second order, where |α| = nc is the mean intracavity photon number and δa are the fluctuations. The interaction Hamiltonian is then

†  2 ∗ †  HI = −~g0(b + b ) α + α δa + αδa (2.14)

The |α|2 term is interpreted as an average radiation pressure force that induces a static displacement, and may be omitted when considering only fluctuations of the oscillator. The α∗δa† + αδa term may be rewritten by choosing the phase reference such that α = α∗ to give,

† † HI = −~g0α(b + b )(δa + δa) (2.15) = −~g(bδa + b†δa† + b†δa + bδa†) (2.16) where the vacuum optomechanical coupling rate has been redefined to include the intracavity photon √ number, i.e g = g0α = g0 nc. By comparing this coupling rate to the mechanical and optical decay rates, it is possible to define a “strong coupling” regime where the coherent interaction between the two sub-systems is faster than all decoherence mechanisms. This condition, namely g > κ, Γm, corresponds to hybridization of the mechanical and optical degrees of freedom. It should be noted that while the linearized Hamiltonian (Eq. 2.16) is only valid in the weak vacuum coupling regime g0  κ, Γm, the coherently boosted coupling can be arbitrarily large while still retaining the validity of the linearization. With on resonance optical probing all of the terms within Eq. 2.16 are equally addressed. This + † means the interaction Hamiltonian can be written as HI ∝ x δX , where x ∝ (b + b ) is the mechanical position and δX+ ∝ δa† + δa is the optical amplitude. In the optically linearized picture presented here, the mechanical oscillator induces a frequency modulation of optical field, enabling precise read- out of the motion via subsequent homodyne detection. Furthermore, the δX+ term commutes with the full Hamiltonian, indicating a quantum non-demolition (QND) measurement of the intracavity amplitude (68). Off resonance probing, in contrast, allows selective terms in Eq. 2.16 to be enhance or suppressed.

8 This is evident when the mechanical and optical operators are redefined in the following way b = eiωmtb and δa = ei∆tδa and the rotating wave approximation (RWA) applied. Taking, for example, the case of ∆ = ωm the interaction Hamiltonian after the RWA is

[ωm]  † † HI = −~g b δa + bδa (2.17) which is equivalent to a beam splitter interaction, enabling cooling of the mechanical motion via exchange between the “cold” photon modes and the typically “hot” phonon modes (14, 109). This effect has been demonstrated in many optomechanical systems, for example see Refs. (25, 32, 133). Furthermore, the coherent nature of this interaction also allows state swapping, enabling non-classical optical states to be written onto the mechanical degree of freedom. While non-classical state transfer has yet to be achieved, the signature of coherent exchange of Gaussian states has been experimentally observed in microtoroidal resonators (162).

Detuning to the other side of resonance, ∆ = −ωm, the dominant terms in the interaction Hamilto- nian are

[−ωm]  † † HI = −~g bδa + b δa (2.18) which has been shown to be equivalent to two-mode squeezing, characterized by existence of quantum entanglement between the phononic and photonic modes (14, 109). As such, this configuration could be used as a resource to generate highly non-classical mechanical states for fundamental research or quantum information technologies (109). However, in contrast to the previous case this interaction amplifies the mechanical mode, making the observation of two-mode squeezing difficult due to satu- ration and other technical nonlinear effects. In Chap. 4 these detrimental nonlinear saturation effects are discussed and experimentally characterized with and without a proposed feedback stabilization protocol. It should be noted that the equations and discussions presented in this section equally apply to electromechanical systems. In this context, pioneering work has also been achieved, where the me- chanical motion of a capacitor plate or transmission line modulates an radio frequency field, with transduction capable of measuring the zero-point motion of the oscillator. The most notable exper- iments are presented in Refs. (116, 119, 120, 157, 158), where ground state production and readout of a mechanical oscillator coupled LC circuit has been demonstrated in addition to optomechanical entanglement and observation of quantum backaction.

2.3.1 Quantum Langevin Equations

The Hamiltonian formalism presented in the previous section provides good insight into the coherent dynamics of the optomechanical interaction. However, the effects of dissipation, quantum fluctua- tions, and optical driving are more easily considered within the Heisenberg framework of the quan- tum Langevin equations (QLE). The details of including noise into the QLE is extensively covered in many texts, for example Refs. (55, 167). For the equations of motion it is generally more intuitive to consider the dynamics of the me-

9 chanical position rather than the associated raising and lowering operators, therefore the Hamiltonian from the previous section (Eq. 2.13) is rewritten terms of mechanical position and momentum using Eq. 2.3 and Eq.2.4,

2 2 2 † p mωm x † H = ~∆a a + + − ~g0 x a a (2.19) 2meff 2

Following Ref. (60), the quantum Langevin equations can then be derived from from Eq. 2.19, giving

p(t) x˙(t) = (2.20) meff 2 † p˙(t) = −ωmmeff x(t) − ~g0a(t) a(t) − Γm p(t) + Fth(t) (2.21) √ √ a˙(t) = − (κ − ∆) a(t) − ig0 x(t)a(t) + κain(t) + κ0avac(t) (2.22)

where the introduction of Γm p(t) and κa(t) terms correspond to mechanical and optical dissipation respectively. To accurately describe Brownian motion of the mirror the mechanical driving term can be shown (60, 143) to have a correlation function of the form Z ! ! 0 dω −iω(t−t0) ~ω hFth(t)Fth(t )i = ~meffΓm e ω coth + 1 (2.23) 2π kBT ≈ ΓmmeffωmkBT (2.24) where the approximation is only valid in the high temperature limit. In contrast, the quantum noise √ √ entering via the in-coupling port ( κain(t) term) and the intrinsic cavity loss ( κ0avac(t) term) has zero mean and is delta correlated,

† 0 † 0 0 hain(t)ain(t )i = havac(t)avac(t )i = δ(t − t ) (2.25)

Equations 2.20-2.22 contain all the interesting optomechanical phenomena discussed in the previ- ous section, and will be used as a starting point for each subsequent chapter of this thesis. While some chapters may require additional terms to be added to explain the experimental results, these terms will be fully described in the introduction/theory section of the relevant chapter.

10 Chapter 3

Feedback Cooling and Temperature Estimation

In this chapter, feedback cooling of a mechanical oscillator using a combination of electrical gradient forces and ultrasensitive optical transduction is presented. The primary goal here is to investigate the effect of feedback induced correlations on the temperature estimate of a mechanical oscillator. In this context, this chapter details the mathematical treatment of a continuous position measurement, made from two independent transducers of a mechanical oscillator experiencing feedback cooling generated from one of the transduction signals, denoted as in-loop. Theoretical predictions, such as squashing of the in-loop transduction signal and cooling limited by the signal to noise ratio, is in good agreement with experimental results. The observed squashing highlights the importance of a second out-of-loop transduction signal for accurate characterization of the oscillator. This represents an enabling step towards strong control and accurate characterization of a micromechanical oscillators specifically in the quantum regime where dynamical backaction generates optomechanical correlations. This chapter provides further analyses on feedback cooling experiments that were performed im- mediately prior to commencing my PhD (85). Here, the experimental methods involved in the imple- mentation of feedback cooling and optical locking are detailed, with considerations of heating from dynamical backaction, as well as further analyses of the feedback cooling results.

3.1 Introduction

Cooling, control and characterization of macroscopic mechanical simple harmonic oscillators near their ground state are all inherently difficult tasks. Achieving these capacities would have far reaching applications, enabling fundamental science, such as tests of quantum gravity and experimental stud- ies of quantum nonlinear dynamics (138, 170), in addition to applications in quantum information systems (97) and ultra-precise sensing and metrology. Experimental development towards this goal is being led by two starkly different architectures; Nanoelectromechanical systems (NEMS) and cav- ity optomechanical systems (COMS). COMS possess ultrasensitive transduction capable of resolving the quantum fluctuations of a mechanical oscillator; but rely on inherently weak radiation pressure forces, hence limiting the strength of mechanical actuation to the available laser power. In contrast,

11 NEMS have strong electrostatically induced actuation but transduction sensitivities typically orders of magnitude lower than COMS, arising from the massive discrepancy in intracavity photon number and characteristic wavelength between electrical and optical systems. Cavity opto-electromechanical (COEMS) feedback cooling schemes, as in Ref. (85), extend strong electrical gradient force actuation to COMS with ultra high transduction sensitivities. Considering such a system, a detailed deriva- tion of the theoretical expressions for the final temperature of the oscillator through analysis of the mechanical power spectrum obtained from the probe beam is given. In particular, focusing on the observed squashing effect, which was first observed by (21, 126), that results from anti-correlations between the transduced noise and mechanical noise whilst feedback cooling with high gain. Squash- ing is manifested in the high feedback gain limit when a single probe beam is used for both feedback cooling and characterization of the oscillator, resulting in the observed mechanical spectrum dropping below the transduction noise which corresponds to an unphysical inverted Lorentzian and hence an apparent negative temperature. For mechanical resonators in the MHz regime the mechanical ground state requires micro-Kelvin temperatures which can be achieved via a combination of cryogenic refrigeration and feedback cool- ing. In other work, experiments involving mechanical resonators with physical dimensions ranging from meters to nanometers have reached milli-Kelvin temperatures, with the aid of cryogenic pre- cooling, by exploiting optomechanical or electromechanical coupling with active feedback (40, 77, 112, 126) or dynamical backaction effects (32, 58, 142, 157, 162). Here, feedback cooling of a micro- toroid is demonstrated using a combination of optical transduction and electrical actuation techniques, as detailed in (101), from room temperature to 58 K for a 6 MHz mechanical mode. The key point is the contrast to other experimental work where solely optical, or solely electrical, methods were used for both transduction and actuation. Although the final temperature achieved is modest in comparison with recent advances in micro- toroid optomechanics (162), this technique is capable of approaching the mechanical ground state, even in the unresolved sideband regime. Furthermore the microtoroid architecture allowed dual opti- cal probes for accurate mechanical characterization in a feedback cooling context, similar to resolved sideband cooling experiments (142, 162). It will be shown that having a second out-of-loop probe overcomes the problem of squashing, providing accurate mechanical characterization even in the presence of high feedback gain. This capacity will be particularly important in the quantum regime when radiation pressure backaction intrinsically generates correlations between probe beam and me- chanical motion. The dependence of the observed out-of-loop temperature, as a function of feedback gain, agrees well with theory for the actual oscillator temperature; while the in-loop prediction varies dramatically, also in good agreement with theory.

3.2 Limits to Feedback Cooling

In this section a mechanical oscillator experiencing both Brownian forces and an external feedback force is considered, following closely the work of (36, 124). As shown in Fig. 3.1, the mechanical motion is transduced by coupling to an optical field within an optical microcavity. Our schematic

12 Figure 3.1: Schematic of feedback cooling model including an independent out-of-loop probe field de- noted asa ˆ(2). The in-loop probe fielda ˆ(1) is used for feedback.

shows a probe beam evanescently coupled to a microtoroidal resonator with optical whispering gallery modes. It should be noted that this treatment is entirely compatible with feedback cooling techniques based on other forms of transduction such as electrical transduction commonly seen in NEMS. Since two probe beams are used, one for the feedback loop and another for mechanical characterization the terminology in-loop probe and out-of-loop probe is adopted, respectively. As described in Sec. 2.3.1, the motion of a mechanical oscillator exposed to Brownian FT (t) and external Ffb(t) forces is given by (36, 124)

h 2 i meff x¨(t) + Γm x˙(t) + ωm x(t) = FT (t) + Ffb(t) (3.1) with Γm and meff respectively being the damping rate and effective mass of the mechanical oscillator. Ignoring mean displacements and considering only the mechanical fluctuations, denoted δx(t), the equation of motion in the Fourier domain is

δx(ω) = χ(ω) [FT (ω) + Ffb(ω)] (3.2) where χ(ω) is the mechanical susceptibility given by

1 χ ω , ( ) = 2 2  (3.3) meff ωm − ω − iΓmω

For low mechanical loss systems, i.e low damping rate Γm, the mechanical susceptibility can be ap- proximated by a Lorentzian lineshape. The motion of the mechanical oscillator is imprinted onto the circulating field within the optical

13 ( j) resonator δaˆ which is then out-coupled at a rate γin and detected on a photodiode. The fluctuations ( j) of the out-coupled field are then a combination of laser noise δaˆin in addition to fluctuations from the mechanical oscillator δx(ω),

( j) p ( j) ( j) δaˆout(ω) = 2γinδaˆ + δaˆin β( j) = δx(ω) + δaˆ( j), (3.4) 2 in

( j) ( j) where δaˆout(ω) is the fluctuation annihilation operator describing the out-coupled field and β is a parameter describing the coupling strength between the optical and mechanical degrees of freedom. Superscripts on the optical field operators denote the in-loop ( j = 1) and out-of-loop ( j = 2) probe. In this expression radiation pressure back-action on the optical cavity from the intra-cavity field is neglected. This is the case of most relevance to the experiments reported later in Sec. 3.4. Radiation pressure back-action can be included with a relatively straightforward modification to the equation of motion, and may be done so in future work. Such a regime is particularly interesting since quantum fluctuations of the light are sufficient to drive, and hence correlate, the mechanical motion (129). To maintain consistency with the experiment detailed in Sec. 3.4, but without loss of generality, it is assumed that β( j) is real. This corresponds to the case where the mechanical oscillation is imprinted ˆ ( j) ( j)† ( j) on the amplitude quadrature δXout = δaˆout + δaˆout of the out-coupled field; and coincides with the optical probe field being detuned from the cavity resonance by a factor of order the cavity decay rate. Direct detection on a photodetector then yields the photocurrent

( j) ( j)† ( j) ( j) 2 ( j) ˆ ( j) i (ω) = aˆout aˆout = |α | + α Xout(ω) (3.5)   ( j) ( j) ˆ ( j) ( j) ˆ ( j) ( j) δi (ω) = α Xout(ω) = α δXnoise(ω) + β δx(ω) (3.6)

( j) ( j) ( j)|2 where α = haouti is the gain of the photodetection process, with |α corresponding to the mean detected photon flux. The transduced mechanical motion δx˜( j) deduced from the jth probe field, is then

δi( j)(ω) δx˜( j)(ω) = = δxˆ( j) (ω) + δx(ω), (3.7) α( j)β( j) noise

( j) ˆ ( j) δX (ω) β( j) where δxˆnoise(ω) = noise / is the uncertainty in the position estimate due to transduction noise. Throughout this chapter parameters with tilde accents indicate estimates from measurements i.e δx˜. To achieve cooling, a viscous damping force must be applied to increase the dampening rate. Since the measurement process described here provides the position of the oscillator, the viscous damping force is achieved most easily by delaying the signal by one quarter of the mechanical oscillation period. Using photocurrent i(1), the feedback force is given by

(1) Ffb(ω) = giΓmmefωmδx˜ (ω) −1 (1) = −gχ (ωm)δx˜ (ω) (3.8)

−1 where g is the feedback loop gain, made dimensionless by multiplying by χ (ωm). This feedback

14 force substituted into Eq. (3.2) modifies the mechanical motion to give

h −1  (1) i δx(ω) = χ(ω) FT (ω) − gχ (ωm) δxˆnoise(ω) + δx(ω) . (3.9)

Rearranging for δx(ω) the mechanical fluctuations in the presence of feedback are

0 h −1 (1) i δx(ω) = χ (ω) FT (ω) − gχ (ωm)δxˆnoise(ω) , (3.10)

0 −1 −1 −1 where χ (ω) = [χ (ω)+gχ (ωm)] is the modified mechanical susceptibility due to feedback which 0 remains Lorentzian, but with modified width Γ = Γm(1+g). The feedback modified mechanical noise spectrum is then

D 2E 0 2 h −1 2 (1) i S x(ω) = |δx(ω)| = |χ (ω)| S T (ω) + |gχ (ωm)| S noise(ω) (3.11)

( j) where S T and S noise(ω) are, respectively, the power spectra of the Brownian force experienced by the ( j) oscillator, and of the transduction noise. S noise(ω) will generally consist of shotnoise due to quantiza- tion of the probe field, and classical noise sources due to the non-ideality of the transduction probe e.g ( j) laser phase noise or electrical amplifier noise. Typically, both S T and S noise are flat over the frequency range relevant to a high Q mechanical oscillator; so that S x remains a Lorentzian under feedback, but with altered height and width. According to Parseval’s theorem the integral under S x(ω) is propor- tional to the total mechanical energy. Hence, the resulting modified motion is equivalent to that of an oscillator at a different temperature T given by

R ∞ 0 T −∞ S x(ω)dω S x(ωm)Γ = R ∞ = . (3.12) T0 S x,g=0(ωm)Γ0 −∞ S x,g=0(ω)dω

Substituting Eq. (3.11) into this expression after some work it is found that ! T g2 1 = 1 + , (3.13) T0 SNR 1 + g where T0 is the initial temperature and

S x,g=0(ωm) SNR = (1) (3.14) S noise(ωm) is the signal-to-noise ratio of the peak of the mechanical noise to the transduction noise on the probe used for feedback cooling, which can be directly determined from δx˜(1)(ω). It is then clear that cooling √ is achieved for positive g, with a minimum temperature when g = 1 + SNR − 1 of √ 1 + SNR − 1 T = 2T . (3.15) min 0 SNR √ It is interesting to note that at feedback gains of g > 1 + SNR − 1 heating from feedback loop noise overcomes the cooling due to damping of the oscillators intrinsic motion, and the temperature

15 increases as gain increases. Hence, as might be expected, maximising the signal-to-noise ratio is critical for achieving high levels of cooling.

3.2.1 Temperature Estimate Using Feedback Photocurrent

The mechanical oscillator temperature can be inferred from the transduced mechanical motion signals δx˜( j)(ω) since the mechanical power spectra can be extracted directly from the transduction spectra as ( j) ˜ ( j) ( j) ˜ ( j) ( j) 2 ( j) S x (ω) = S x (ω) − S noise where S x (ω) = hδx˜ (ω) i. Assuming the transduction noise δxN is uncor- related to the mechanical motion δx then both in-loop ( j = 1) and out-of-loop ( j = 2) transduction signals can be used to calculate the oscillator temperature. This treatment follows closely to (36, 124). The temperature estimate is then simply,

T˜ ( j) S ( j)(ω ) Γ0( j) x m · = ( j) ( j) . (3.16) T0 S x,g=0(ωm) Γ0 where T˜ ( j) is the inferred temperature from the jth probe field. However, the fact that the feedback imprints transduction noise from the in-loop probe onto the mechanical oscillator brings this approx- imation into question in the case of large feedback gain. As a result one might expect that in this regime the temperature inference from an in-loop signal will no longer be valid. Substituting the known solution for δx(ω) under feedback, as in Eq. (3.10), into Eq. (3.7) for the transduced mechanical motion, the in-loop position estimate is then

(1) (1)  0 −1  0 δx˜ (ω) = δxˆnoise(ω) 1 − gχ (ω)χ (ωm) + χ (ω)FT (3.17)

(1) (1) 2 So that the in-loop mechanical spectra S˜ x (ω) = hδx˜ (ω) i is

2 ˜ (1) 0 2 (1) 0 −1 S x (ω) = |χ (ω)| S T + S noise 1 − gχ (ω)χ (ωm) 0 2 h (1) −1 2 h 2 ii (1) = |χ (ω)| S T − S noise|χ (ωm)| g + 2g + S noise. (3.18)

One sees that that all frequency dependence remains in the modified mechanical susceptibility term χ0(ω), hence feedback induced correlations do not destroy the Lorentzian dynamics of the in- trinsic mechanical motion, or the modified mechanical damping rate. However, the peak and width of the spectra, and hence the temperature estimate, is strongly effected. After some work it can sub- sequently be shown that " # g(g + 2) S (1)(ω) = S˜ (1)(ω) − S (1) = |χ0(ω)|2S 1 − , (3.19) x x noise T SNR and from Eq. (3.16) one finds

h g(g+2) i (1) | 0 |2 − T˜ χ (ωm) 1 SNR Γm(1 + g) = 2 T0 |χ(ωm)| Γm " # g(g + 2) 1 = 1 − × , (3.20) SNR 1 + g

16 Figure 3.2: Theoretical temperature prediction whilst feedback cooling. A Out-of-loop temperature pre- diction whilst cooling from room temperature. Dotted lines show current capabilities of SNR and gain. B In-loop temperature prediction whilst cooling from room temperature showing squashing effect corre- sponding to unphysical negative temperature.C Feedback cooling with cryogenic precooling to 300mK.

17 which only agrees with the mechanical oscillator temperature in Eq. (3.13) in the restrictive limit that g  SNR. Critically, for g(g + 2) = SNR the temperature estimate is 0 K, dramatically departing from the actual temperature. For g(g + 2) > SNR the temperature estimate becomes negative. This is the region where inverted Lorentzian spectra appear in expriments (36, 124, 126, 134), leading to a prediction of negative mechanical energy and hence clearly unphysical temperature predictions. This is shown in Fig. (3.2)C which clearly shows how the in-loop probe beam underestimates the final temperature when high feedback gain is applied. It is possible to correct for these errors as has been done in experiments such as Ref. (126), by noting that using Eqs. (3.13) and (3.20) the final oscillator temperature can be expressed as " # SNR + g2 T = × T˜ (1). (3.21) SNR − g(g + 2)

However, such predictions of the temperature are indirect, relying on the accurate determination of g and SNR, and on assumptions about the underlying behaviour of the system. This is of particular concern in the quantum regime where measurement backaction plays an additional role in correlating and perturbing both mechanical oscillator and optical field.

3.2.2 Temperature Estimate Using an Independent Photocurrent

To overcome this problem an independent out-of-loop probe field is used to transduce the motion of the mechanical oscillator, as shown in Fig. 3.1 which allows the elimination of feedback induced correlations between the optical field and the mechanical motion. So long as the in-loop and out-of- loop probe fields are uncorrelated prior to interaction with the mechanical oscillator, the mechanical (2) motion δx(ω) is uncorrelated to the second probe noise δxˆnoise(ω). Hence, the photocurrent spectra is given simply by

h i (2) ˜ (2) (2) 0 2 2 −1 2 (1) S x (ω) = S x (ω) − S noise = |χ (ω)| S T − g |χ (ωm)| S noise " # g2 = |χ0(ω)| S 1 + . (3.22) T SNR where S out−noise(ω) is the power spectrum of the out-of-loop probe field in the absence of the mechan- ical oscillator. The temperature estimation from the second probe is now identical to Eq. (3.13). " # T˜ (2) g2 1 = 1 + × , (3.23) T0 SNR 1 + g hence giving an unbiased estimate of the temperature, independent of feedback induced correlations.

3.3 Experimental Design

Our system consisted of a microtoroidal resonator, shown in Fig 3.3, which integrates both high 6 3 quality optical (Q ≈ 10 ) and mechanical resonances (Qm ≈ 10 ). Evanescent coupling from a

18 Figure 3.3: A Scanning electron microscopy image of a microtoroidal resonator B & C Finite element simulations of microtoroidal whispering gallery optical modes and mechanical modes, respectively.

tapered fiber allowed light to be coupled into whispering gallery modes. Mechanical control through gradient force actuation was realised by applying a voltage to a sharp electrode with 2 µm tip diameter positioned 15 µm vertically above the microtoroid. The silicon chip was placed on top of a grounded flat electrode, with the combination of electrodes forming a capacitor. An applied voltage difference generates charge build up on both the flat and sharp electrodes. The close vicinity of the microtoroid to the sharp electrode allows the field it experiences to be approximated by a point source. Owing to the strong electric field gradient generated by the sharp electrode, combined with surface charge induced static electric fields which polarize the microtoroid, large gradient forces were exerted. Details of the electrical actuation technique used here are given in (101) and shown in Fig 3.4. The transduction spectrum obtained by spectral analysis of the in-loop transduction signal without gradient force actuation or feedback is shown in Fig. 3.5 where the peaks correspond to mechanical modes. Finite-element modeling identified these modes to be the lowest order flexural mode, and the two lowest order crown modes. The absolute mechanical displacement amplitude was calibrated by observing the response of the optical transmission to a known reference phase modulation (145). The flat spectral noise background was due to laser phase noise, as confirmed by an observed square root dependence of its amplitude on probe power. Characteristic parameters of the mechanical oscillator such as the effective mass, damping rate and resonant frequency can be determined from a fit to the transduced spectral density S (ω) = g=0 S x (ω)+S noise. In our case the transduction noise S noise corresponds to laser phase noise and when no g=0 2 feedback is applied the thermal spectrum of the mechanical oscillator is given by S x (ω) = S T |χ(ω)| where S T = 2meffΓmkBT is calculated from the fluctuation dissipation theorem and represents the Langevin force responsible for Brownian motion (36, 124). The in-loop and out-of-loop transducers 1/2 −18 −1/2 1/2 −18 −1/2 were found to have sensitivities of S noise =(4±1)×10 m Hz and S noise =(5 ±1)×10 m Hz , respectively; two orders of magnitude better than any NEMS to date (78, 157). The mechanical mode selected for cooling was a radial flexural mode at 6.272 MHz and was experimentally found to have an effective mass of 30±10 µg with a 11.5 kHz damping rate. The peak zero-point spectral density 1/2 p~ −20 −1/2 for this mechanical mode is calculated to be S zp = /me f f Γmωm = 3×10 m Hz . The figure of

19 Figure 3.4: Experimental schematic including electronic locking method for in-loop probe. The out-of- loop probe was thermally locked. FPC: fibre polarization controller. All experiments were carried out at room temperature (T = 300 K) and atmospheric pressure.

20 Figure 3.5: Mechanical spectrum at room temperature showing three mechanical modes and their corre- sponding FEM simulations. The frequency axis is centered about 6.272 MHz. The shaded region identifies the mechanical mode chosen for cooling and the inset is a magnification of this mode with transduction sensitivity (grey) and electronic dark noise (black) merit in determining the suitability of systems for experiments involving the quantum behaviour of a mechanical oscillator is the ratio of the transduction sensitivity to the zero point motion. A ratio of 1/2 S zp 1/2 /S noise > 1, indicating observability of the quantum behaviour, is ideal and has only been recently demonstrated in (8, 129, 156). Our system has a transduction sensitivity that is within 2 orders of 1/2 − S xp 1/2 2 magnitude of the mechanical zero point motion /S noise ≈ 10 . Implementing recently improved fabrication techniques (9), cryogenic cooling (142) and shot noise limited homodyne detection (145) should enable sub-zero point motion transduction sensitivity in our COEMS. Control of the microtoroid COEMS can be achieved conveniently by feedback of the transduction signal to actuate the mechanical motion. This is an enabling technology, facilitating, for example, active cooling or heating (36, 124), tuning of the mechanical resonance frequency (71), and phonon lasing (64). Fig. 3.4 shows a schematic of our experimental setup. The experiment was performed in an enclosed chamber at room temperature and atmospheric pressure. The microtoroid was positioned on a piezostage (ThorLabs MAX312), and the sharp electrode was affixed to a XYZ translational stage. Two tuneable diode lasers (New Focus TLB-6320 and TLB- 6328) at 986 nm and 1560 nm acted as the in-loop and out-of-loop probe beams respectively; both of which had good optomechan- ical coupling to the mechanical mode of interest. Both beams were simultaneously coupled through a tapered fibre to optical modes in the microtoroid before being split by two consecutive 2 × 2 fibre couplers and sent to three detectors; two InGaAs detectors (New Focus 1801) used for out-of-loop transduction and locking and one Si detector (New Focus 1811) used for in-loop transduction. Due to the vastly different wavelengths of the two beams, the condition for critical coupling could not be simultaneously meet. However, the in-loop beam was successfully thermo-optically locked (28) to the FWHM of a microtoroidal optical resonance (Q ≈ 106) and then detected on the Si photodi- ode. Thermo-optical locking uses the natural response of the refractive index of silica to temperature

21 changes to vary the optical path length of the cavity, counteracting any shift in resonance frequency. Since the out-of-loop probe had a wavelength well above the bandgap of the Si photodiode no optical filter was required to eliminate crosstalk when detecting the in-loop beam. The out-of-loop beam was detected on the InGaAs photodiodes and locked to the FWHM of a microtoroidal optical resonance using a Pound-Drever-Hall (PDH) locking technique. Since the detection band of the InGaAs photodiode spans the frequency of both the in-loop and out-of-loop beams a 1550 nm filter was used to avoid in-loop crosstalk on the locking signal and the out-of-loop transduction. PDH locking allows a laser frequency to be locked to a specific point on the cavity resonance despite drift and jitter in the resonance. For this experiment the laser frequency was dithered at 24 kHz, and the detected photocurrent mixed down at the dither frequency to create a PDH error signal for locking of the laser frequency to the cavity. The photocurrent from the in-loop beam was electronically filtered and amplified then applied directly to the electrode to facilitate electrical actuation. The electronic filtering included two first order Butterworth bandpass filters to isolate the mechanical mode at 6.272 MHz and a notch filter at 4.947 MHz to provide additional attenuation of a nearby unwanted mechanical mode. This avoided saturation of the amplifiers as well as allowing maximum amplification of the frequency band of in- terest. A series of amplifiers (Mini-Circuits AMP-76, ZFL-500, and ZHL-32A) were used to achieve high gain. The filters were positioned between the amplifiers to suppress amplified noise as well as to minimize harmonic resonances generated from connecting consecutive amplifiers. Variability of the overall gain was realized by adding attenuators, and the phase delay required for cooling was achieved by adding lengths of coaxial cable. This technique for feedback allows electrical gradi- ent forces as large as 0.40 µN to be achieved, significantly higher than has been demonstrated with radiation pressure actuation and without observable heating effects. Since both in-loop and out-of-loop transduction were achieved with a detuned probe laser, dy- namical backaction heating (135) had the potential to compete with the feedback cooling process. The in-loop and out-of-loop probe powers incident on the COEMS were 2.8 mW and 1.5 mW, re- spectively. At these powers, and with the COEMS properties given in this report, the heating rate is predicted to be less than 1 % of the intrinsic mechanical damping rate Γm (135), and therefore should be insignificant. This was confirmed experimentally by varying the incident optical power over several orders of magnitude, with no observed heating effects.

3.4 Experimental Results

In-loop and out-of-loop temperature predictions were obtained from a Lorentzian fit to the spectra, extracting the peak response and the linewidth of the mechanical mode. Using Eq. (3.12) the tem- perature can then be estimated by taking the ratio of the product of these two parameters with and without feedback. Fig. 3.6A and B show the transduced in-loop and out-of-loop mechanical spectra, respectively, for different feedback gain. Note the predicted squashing effect on the in-loop spectra at high gain. Fig. 3.6C shows the mechanical oscillators temperature inferred from the transduced in-loop and out-of-loop power spectra as a function of feedback gain.

22 The out-of-loop temperature estimate is in good agreement with the expression for the actual me- chanical oscillator temperature seen in Eq. (3.13), with cooling improving with gain yet limited to T =58 K when g=8 due to the feedback loop noise, which in our case gave a signal-to-noise ratio of

SNR = 100 dB. This compares to a theoretically predicted minimum temperature of Tmin = 53 K from Eq. (3.15). As predicted in Eq. (3.12) and experimentally shown in Fig. 3.6C (circles) the tempera- ture increases with higher gain due to feedback noise being imprinted on the oscillator. In contrast, at high gain the in-loop predicted temperature diverges significantly from the actual temperature as theoretically predicted by Eq. (3.20) and experimentally shown in Fig. 3.6C (squares). The estimated temperature drops well below the theoretical limit and passes through 0 K, corresponding to an un- physical inversion of the observed mechanical response. This is strikingly illustrated in Fig. 3.6A and demonstrates that an out-of-loop probe is vital for independent mechanical characterization in any circumstance where mechanical oscillation becomes correlated to the transduction noise. Even at the relatively low gain of g = 3.2 the in-loop transduction under-estimates the temperature by as much as 10 %.

The cooling presented in this section achieves a final phonon occupation number of hni=kBT/~ωm = 19, 000. Despite this modest value, control and read out of mechanical zero-point motion of a mi- crotoroid is within reach due to recent advancements in microtoroid opto-mechanics discussed ear- lier (9, 142, 145). Incorporating these techniques into our existing system the transduction is pre- 1/2 −19 1/2 dicted to be limited by shot noise at S noise ≈ 10 m Hz and our mechanical quality factor to be 4 Qm ≈ 5 × 10 for a 24MHz mechanical mode with effective mass me f f ≈ 10ng. With these improve- ments and a cryogenic starting temperatures of 1.7 K a feedback cooled phonon occupation number of hni ≈ 0.7 could be realised.

3.5 Conclusion

In conclusion, this chapter provides detailed theoretical and experimental information regarding the previous publication (85) in which feedback cooling was achieved, limited only by noise on the transduction signal. The theoretical expressions derived in this here, along with experimental data, confirmed that feedback cooling under high gain causes anti-correlations between the mechanical oscillation and the transduction noise, resulting in a striking discrepancy between the in-loop and out-of-loop mechanical characterizations. This is particularly important in the quantum regime where dynamical backaction from radiation pressure correlates the mechanical motion to the probe noise. The transduction sensitivity, which ultimately defines a lower bound on the cooling capabilities, also needs to be able to resolve the zero point motion for control and read out of mechanical systems in the quantum regime. 1/2 −18 −1/2 Here, a transduction sensitivity of S noise = 4×10 m Hz was achieved, less than two orders of magnitude away from the mechanical zero-point motion. In a broader sense this technology represents an enabling step towards control of macroscopic quantum mechanical systems; and in particular towards quantum nonlinear mechanics where a mechanical oscillator in its ground state is strongly driven into a regime exhibiting quantum nonlinear dynamics. Furthermore, integration into ultracold

23 superconducting circuits has the potential to unify cavity opto-mechanics with the burgeoning field of circuit quantum electrodynamics (146).

24 Figure 3.6: Feedback cooling with varying feedback gain (85). A and B: In-loop and out-of-loop trans- duction spectra respectively. C: Temperature inferred using in-and out-of-loop transduction signals. The red circles  and blue squares  respectively denote out- and in-loop temperature inferences; the solid red curve and dashed blue curve respectively denote the theoretical predictions of actual mechanical oscillator temperature and inferred in-loop temperature. Figure from Ref. (85)

25 Chapter 4

Feedback Suppression of Parametric Instability

The intracavity power, and hence measurement sensitivity, of optomechanical sensors is commonly limited by mechanical parametric instability, arising from the dynamical backaction discussed in Sec. 2.3. In this chapter the degradation of measurement sensitivity induced by mechanical para- metric instability in a micron scale cavity optomechanical system is experimentally characterized. To suppress the detrimental instability, a feedback control technique is proposed and experimen- tally demonstrated based on the opto-electromechanical feedback described in Chap. 3. As a re- sult a 5.4 fold increase in mechanical motion transduction sensitivity is achieved to a final value of 1.9 × 10−18 m Hz−1/2. The content of this chapter is published in the following paper

Glen I Harris, Ulrik L Andersen, Joachim Knittel, Warwick P Bowen , “Feedback Enhanced Sen- sitivity in Optomechanics: Surpassing the Parametric Instability Barrier,” Physical Review A 85, 061802(R), (2012)

4.1 Instabilities in Optomechanics

Optical techniques are capable of ultra-precise measurements of parameters such as phase, position and refractive index. The sensitivity is typically limited by optical shot noise which can be reduced by increasing optical power. Using coherent states of light the ultimate sensitivity is fundamentally set by the Standard Quantum Limit (SQL) (31, 156). However, well before the SQL is reached radiation pressure may become sufficiently strong to severely alter the dynamics of the intrinsic mechanical motion of the sensor. This regime, called parametric instability, is characterized by violent mechani- cal oscillations and was first theoretically investigated by Braginsky (24) in the context of large scale interferometers for gravitational wave detection followed by experimental observation in electrical readout of resonant bar systems (41) and later in optical micro-cavities (136). The physical pro- cess, described graphically in Fig. 4.1, is a result of radiation pressure from asymmetric Stokes and anti-Stokes sidebands generated from the mechanical motion of the cavity. If this process, known as dynamical backaction heating (76), amplifies the motion at a rate faster than the mechanical decay

26 rate then parametric instability occurs. Due to large mechanical oscillations the optical cavity reso- nance frequency is modulated by more than the optical linewidth, leading the probe field to sample the intrinsically non-linear (Lorentzian) optical lineshape and saturate the mechanical amplification. Subsequently, the noise floor of the optomechanical sensor increases, rendering it ineffective at trans- ducing small signals. Consequently, parametric instability is predicted to be a problem in the context of ultra-precise optical sensors such as gravity wave interferometers (24).

ω−Ω ωΩ ω ω−Ω ω

Figure 4.1: Ultra sensitive optomechanical systems exhibit parametric instability through cavity enhance- ment of the Stokes sideband produced by their mechanical motion. A In detuned micron scale optome- chanical systems Stokes and anti-Stokes sidebands spectrally overlap with the same optical mode. B In large-scale interferometers Stokes sidebands must overlap with an adjacent optical mode. Image of LIGO Livingston Laboratory courtesy of Skyview Technologies

Parallel to the development of gravitational wave detectors, there has been a recent push towards real time read out and control of mesoscopic mechanical oscillators in the quantum regime(116, 150, 157); facilitating new quantum information technologies(97), and experimental tests of quantum non- linear mechanics(92, 138, 170) and even potentially quantum gravity(98). However, parametric in- stability limits both the strength of entangled and squeezed states which may be produced via the optomechanical interaction(164), and - even when employing strategies such as back-action-evasion (BAE) - the ability to transduce the mechanical motion at the quantum level(149). Radiation pressure mediated optomechanical interactions are also of increasing importance to photonic circuits and sensors, where many applications are facilitated by miniaturized integrated ar- chitectures (4, 172). Miniaturization is often accompanied by high mechanical compliance. This has the adverse consequence of increasing exposure to parametric instability but provides the possibility to introduce new functionality via mechanical elements such as optomechanical switches (139) and memories (17), and ultra-precise gyroscopes, magnetometers (53), and mass and force sensors (90). Even hybrid optomechanical circuits have been proposed, where fully integrated phononic and pho-

27 tonic circuits are interfaced via the optomechanical interaction(6); with phononic elements used for memories, filters, and processing elements, and photonic elements used for communication. Para- metric instability can severely adversely effect the performance of such integrated optomechanical devices. The growing role of complex optomechanical systems in both fundamental science and appli- cations means that techniques capable of individually addressing and suppressing instabilities in optomechanical elements are of crucial importance. In this chapter I propose and experimentally demonstrate such a technique based on opto-electromechanical feedback control, characterizing the parametric instability induced degradation in, and feedback induced revival of, mechanical trans- duction sensitivity in a cavity optomechanical transducer. Our cavity opto-electromechanical system (COEMS), seen in Fig. 4.1, consists of a silica microtoroid integrating high Q mechanical and opti- cal modes with strong electrical actuation (85). Parametric instability of a 14MHz mechanical mode is found to occur at optical powers of 60 µW resulting in a drastic loss in broadband sensitivity for higher power levels, and a maximum optomechanical sensitivity still a factor of three hundred higher than the SQL. Stabilization of the parametric instability is achieved via a viscous damping force ap- plied to the unstable mechanical mode using electric feedback. Narrow-band filtering ensures that the feedback force is only applied at frequencies in close proximity to the unstable mode, allowing broadband sensitivities at the level of 1.9 × 10−18 m Hz−1/2 limited by available optical power.

4.2 Theoretical Description of Parametric Instability

The dynamical interaction between light and mechanical motion including radiation pressure Frad, feedback Ffb and thermal forces FT can be described through the equations of motion (76)

h 2 i meff x¨ + Γm x˙ + ωm x = Frad + FT + Ffb (4.1)   p a˙ = − γ − i(∆0 + gx) a + 2γinain (4.2)

The first equation describes the motion of the mechanical oscillator where meff, Γm and ωm are its effective mass, damping rate and resonance frequency, respectively; F = ~g|a(t)|2, and F = √ rad T 2ΓmkBTmξ(t) where ξ(t) is a unit white noise Wiener process. The second equation describes 2 the intra-cavity optical field where ∆0 is the optical detuning, |a| is the intra cavity photon num- 2 ber and |ain| is the input photon flux, coupled into the cavity at rate γin. The total optical decay rate is γ = γin + γ0 where γ0 is the intrinsic decay rate. The equations are coupled via the optome- chanical coupling parameter, g, which gives rise to both static and dynamic effects such as radiation pressure bistability (44), the optical spring effect (151), and dynamical backaction cooling and am- plification (12, 36, 58, 75). Due to the nonlinear nature of the equations of motion, linearization is required to reach an analytic solution where a separation of each variable into its mean value and fluc- tuations is performed; a = a¯ + δa and x = x¯ + δx. Taking the linearized equations into the frequency

28 domain yields p 2γinδa˜in + iga¯δx(ω) δa(ω) = (4.3) γ − i (∆ − ω) −1 h † ∗ i χ0 δx(ω) = ~g a¯δa (−ω) + a¯ δa(ω) +FT(ω)+Ffb(ω) (4.4)

−1 −1 h 2 2 i where χ0 = meff ωm − ω + iΓmω is the mechanical susceptibility and ∆ = ∆0 + gx¯ is the static detuning of the cavity in the presence of radiation pressure. As seen in Eq. (4.3) the mechanical p fluctuations are imprinted onto the field δa which, in turn, is out-coupled with aout = ain − 2γina † and detected on a photodiode giving a photocurrent i = aoutaout. After some work the resulting ∗ † photocurrent fluctuation, δi(ω) = a¯outδaout + a¯outδaout, is found to be ! 2ig|a¯|2∆ ω − i2γ  δi = δx 0 + δi (4.5) γ2 + ∆2 − ω2 + i2γω a where δia contains all noise terms associated with the input field. This signal is then applied back onto the oscillator via the feedback force Ffb(ω) = Gδi where G is a complex feedback gain. Substituting this feedback force and the optical fluctuations, given by Eq. (4.3), into Eq. (4.4), an analytic form can be obtained for the modification of mechanical motion due to radiation pressure and feedback forces combined. The terms contributing to δx modify the mechanical susceptibility, χ, such that

2g|a¯|2∆ ~g + G (iω − 2γ ) χ−1 = χ−1 + 0 , (4.6) 0 γ2 + ∆2 − ω2 + i2γω

If no feedback is applied, corresponding to G = 0, the mechanical susceptibility is modified purely by radiation pressure (75). As is well known the phase of the modulating radiation pressure depends on the sign of the detuning ∆ resulting in either mechanical linewidth narrowing or broadening. Para- metric instability occurs when the modified mechanical linewidth is negative, with correspondingly exponential amplification of the mechanical oscillations. This amplification process eventually satu- rates to give a steady-state linewidth close to zero but positive. Similar to mode competition in a laser this limits the parametric instability to one mechanical mode. Due to the large shifts of the optical resonance from the amplified mechanical motion the average intra-cavity power, and hence radiation pressure, decreases resulting in saturation of the mechanical amplification. Since the mechanism for transduction is equivalent to actuation, the nonlinear saturation comes together with nonlinear trans- duction. This nonlinearity acts to mix different frequency components in the spectrum resulting in broadband noise and hence severely degrading the transduction sensitivity. From Eq. (4.6) it can be seen that this degradation can be canceled and the original mechanical susceptibility recovered if the gain is chosen to be

~g Gcrit = , (4.7) 2γ0 − iω such that all modifications to the mechanical susceptibility from radiation pressure are canceled by the electrical feedback. This simple expression is completely insensitive to flucuations in the detuning

29 2 ∆, the coupling rate γin and the input power |ain| , making the feedback system very robust against external noise sources. This robustness necessarily translates to simplicity of implementation, which is all important for optomechanical systems pushing towards the SQL. Many techniques have been proposed for the stabilization of parametric instabilities such as the addition of acoustic dampers, thermal control and active feedback from a transduction signal using optical, electrical or mechani- cal actuation (74); but feedback stabilization has to date been demonstrated only in large scale low frequency systems (39) where the parametric instability typically occurs due to the presence of many optical modes (Fig. 4.1B) rather than in one optical mode as is the case here (Fig. 4.1A), and with no characterisation of the degradation in sensitivity due to parametric instability, or the enhancement achieved via active feedback. It should be noted that the physical mechanism differs for parametric instability in micron sized systems to that of large scale interferometers. As illustrated in Fig. 4.1, LIGO is in the resolved side- band regime where the mechanical resonance frequency Ω is much higher than the optical linewidth γ. This means that the Stokes sideband must have spatial and spectral overlap with higher order optical modes to generate parametric instability. This process, often called three mode parametric interaction, has been observed on large optical cavities (173) supporting the argument that parametric instability in kilogram scale optomechanical systems will, in future, limit the ultimate sensitivity.

z

6

U

U

z 4 z46

Figure 4.2: A Experimental schematic. Blue (light grey) indicates the optical components allowing trans- duction and parametric instability; Green (dark gray) indicate the electrical components used in the feed- back stabilization. FPC: Fiber polarization controller. BPF: Bandpass filter. B Observed mechanical spectra below parametric instability threshold.

30 4.3 Feedback Suppression of Parametric Instability

I

o

o. oo.I

oo.I Figure 4.3: Mechanical power spectra B and A show, respectively, the mechanical spectra with and without feedback stabilization of the regenerative 14MHz mode.

Our experimental setup is shown in Fig. 4.2A. A tunable diode laser at 780nm was evanescently coupled into a microtoroidal whispering gallery mode using a tapered optical fiber. The microtoroid had major and minor diameters of 60 µm and 6 µm respectively with a 26 µm undercut. The toroid- taper separation was controlled by a piezo stage to allow critical coupling into the optical cavity. The laser was thermo-optically locked (28, 102) to the full width half maximum (FWHM) of the optical mode which had an intrinsic quality factor of Q ≈ 107. This optical detuning allowed simultane- ous radiation pressure induced mechanical amplification and transduction of the mechanical motion. The absolute mechanical displacement amplitude was calibrated via the optical response to a known reference phase modulation (141). The mechanical motion, which modulates the optical resonance

31 frequency, was detected via fluctuations in the transmitted power incident on an InGaAs photodiode. Fourier analysis of the photocurrent reveals a mechanical power spectra with peaks corresponding to mechanical resonances. A typical spectra containing many mechanical modes can be seen in Fig. 4.2B at optical powers below the threshold for parametric instability. At higher optical powers the unstable mode at 14 MHz, labelled U, experiences parametric instability. Focusing on the effect of this process, with and without feedback stabilization, on the measurement of a mode at 28.6MHz, labeled M. This mode was characterized experimentally to have an effective mass and linewidth of 0.3 µ g and 90 kHz q respectively, resulting in a standard quantum limit of S 1/2 = ~ = 8 × 10−21 m Hz−1/2. SQL 2me f f ω0Γm

Fig. 4.3A shows that at low power the measured mode is easily resolved with the sensitivity improving with increasing optical power as expected from photon shot noise statistics as a function of power. However, as the input power is increased above 60 µW the unstable mode experiences parametric instability, generating harmonics on the transduction signal at 28, 42 and 56 MHz due to the nonlinear process involved in saturation. This is evident by the emergence of a dark narrow band, amongst broadband noise, at 28MHz. The noise spectra at a fixed power of 136µW is shown in Fig. 4.4A (dark line) and reveals the narrowed harmonic at 28MHz, and beat frequencies from mixing with the groups of mechanical modes, labelled G1 and G2 in Fig. 4.2B. The three peaks in the region labelled U + G2 arise from sum frequency generation between the group G2 and the unstable mode U; while the peaks in the regions 2U ± G1 arise from sum and difference frequency generation, respectively, between group G1 and the second harmonic of the unstable mode 2U. In addition to these harmonics and beats, low frequency laser phase noise, seen in the inset to Fig. 4.4A, is also mixed up via sum and difference frequency generation with the second harmonic causing broadband noise. This extra noise is clearly illustrated in Fig. 4.4A (dark trace) where mixed up laser noise completely obscures the motion of the measured mode M (grey trace). The SNR of the measured mode is shown as a function of power in Fig. 4.4B (black squares), with severe degradation apparent once threshold is reached.

Feedback to suppress the parametric instability was implemented by electrically filtering and am- plifying the photocurrent and applying it directly to a sharp electrode placed close to the micro- toroid. This facilitated strong electrical actuation of the mechanical motion through electrical gradi- ent forces (101). Consecutive bandpass filters were used to isolate the unstable mechanical mode at 14 MHz allowing maximum amplification of the feedback signal while eliminating the effect of feed- back on the measured mode M. The feedback phase and gain were controlled inside the feedback loop by an voltage variable phase shifter and attenuator respectively. With correct gain and phase as given in Eq. (4.7), the viscous damping force applied by feedback fully suppressed the parametric instabil- ity and eliminated the harmonics and associated noise from the unstable 14MHz mode, as shown in Fig. 4.3B and Fig. 4.4A (light line). Consequently, the transduction sensitivity of the measured mode was found to improve with optical power, even above threshold, as shown in Fig. 4.4B (circles).

32 4.4 Conclusion

Parametric instability can in principle be circumvented without any feedback by operating in the zero or red detuned regime, rather than the blue detuned regime used here. However, blue detuning is unavoidable in many circumstances. In room temperature experiments with silica microtoroids such as those reported here, for example, the thermal response of silica causes intrinsic locking to the blue detuned side of resonance (28). To overcome this effect using electronic locking would require impractically high gain and bandwidth for the optical powers used in our experiments. Moreover, blue detuning is a requirement for many optomechanical protocols, such as optomechanical amplifi- cation (12, 75, 100), back-action-evasion (149), entanglement and mechanical squeezing (164). It is also expected to be difficult to avoid in complex devices with many optomechanical elements such as gravity wave detectors and photonic-phononic circuits (24). Our scheme provides a pathway to suppress instabilities in each of these circumstances. In complex optomechanical devices, for exam- ple, the feedback forces could be applied by electrodes integrated onto each optomechanical element, while the feedback signals could be obtained either directly from the optical output of the device with spectral filtering to select the appropriate unstable modes, or if there are degenerate unstable modes, from weak optical tap-offs integrated throughout the device. This work represents an important step for stabilization in mesoscopic quantum optomechanical systems, particularly when involving BAE or optomechanical entanglement. Extending into appli- cations, this technique could be used for stabilizing mechanical instabilities in miniaturized photon- ic/phononic circuits and mechanical sensors.

33 Pbf bC Pb s bC d s

Pb bCPb

PbWd bC Cwb b

dPb Rd dRb Pb bC

Pb bC sgss o

Pb bC d dw d dw d dw d dw kC sgo bC sss b

dCfC C C bCC bdC bfC bC ssgo

Figure 4.4: A Mechanical spectra observed with 136 µW of input power, well above the parametric in- stability threshold, with (Grey) and without (Black) feedback stabilization of the unstable 14 MHz mode. Inset: Single sided noise spectrum about the second harmonic 2U (solid line) showing the measured scaling of laser frequency noise (dashed line) and the suppression of noise from thermal locking (grayed region)B Signal to noise ratio of the 28.6 MHz mechanical mode versus optical power with feedback, circles, and without, squares

34 Chapter 5

Minimum requirements for feedback enhanced force sensing

When using linear measurement to estimate an unknown force driving a linear oscillator, feedback is often cited as a mechanism to enhance bandwidth, sensitivity or resolution. In this chapter it is shown that as long as the oscillator dynamics are known, there exists a real-time estimation strategy that reproduces the same measurement record as any arbitrary feedback protocol. Consequently some form of nonlinearity is required to gain any advantage beyond estimation alone. This result holds true in both quantum and classical systems, with non-stationary forces and feedback, and in the general case of non-Gaussian and correlated noise. Recently, feedback enhanced incoherent force resolution has been demonstrated [Nat. Nano. 7, 509 (2012)], with the enhancement attributed to a feedback induced modification of the mechanical susceptibility. As a proof-of-principle this is experimentally reproduced using straightforward filtering. The content of this chapter is published in the following paper

Glen I Harris, David L McAuslan, Thomas M Stace, Andrew C Doherty, Warwick P Bowen , “Mini- mum requirements for feedback enhanced force sensing,” Physical Review Letters 111, 103603 (2013)

5.1 Introduction to Force Sensing

Micro and nano-mechanical oscillators are capable of ultra-sensitive force measurement, allowing precision spin, charge, acceleration, and field sensing (35, 52, 96, 137). It is well known that lin- ear feedback control can improve the performance of nonlinear mechanical sensors (65, 84). For example, in non-contact atomic force microscopy linear feedback is commonly used to stabilise the tip-surface separation, thereby avoiding collisions and suppressing frequency drifts due to short range forces such as van der Waals forces (20, 86). Since feedback control modifies the response of an os- cillator to environmental forces it also appears attractive as a technique to enhance the performance of linear sensors. For instance feedback cooling allows the suppression of thermal noise (36, 104), while feedback tuning of the spring constant can provide increased mechanical response at the signal fre- quency (37). However, such precision enhancement is prohibited for linear processes with stationary linear feedback and uncorrelated Gaussian noise by the well-known principle of neutrality in linear

35 control theory, which shows that the accuracy with which the oscillator position can be determined is independent of feedback (72, 121). Non-stationary processes, non-Gaussian noise and non-linear estimation strategies are each found in a range of linear oscillator-based force sensors. Stroboscopic measurement of impulse forces (165), and variance estimation of incoherent forces as applied in bolometry (88, 161), are two relevant ex- amples. Linear feedback cooling has been proposed as a means to enhance the performance of linear sensors in both cases (56, 165), and experimentally demonstrated in the latter (56). However, neither proposal identifies an optimal estimation strategy. This leaves unresolved the important question of whether the same, or improved, sensing enhancement might be achieved without feedback by apply- ing a better estimation strategy. As described in Sec. 2.3.1, the evolution of a mechanical oscillator 1 can be described, in both classical and quantum regimes (95) by the equation of motion

h 2 i meff x¨ + Γm x˙ + ωm x = Fm(t, x) + Fact(t, x˜) (5.1) where m, Γm, and ωm are the mass, damping rate, and resonance frequency, respectively. For com- pactness the combined force Fm(t, x) = FT(t, x) + Fs(t, x) + Fba(t, x) is used, where FT(t, x) is a thermomechanical force due to the coupling of the oscillator to its environment, Fs(t, x) is the signal force and Fba(t, x) is a backaction force due to the act of measurement. Fact(t, x˜) is an actuation force used for feedback wherex ˜(t) = x(t) + N(t, x) is the instantaneous measurement record of the oscilla- tor position x, and N(t, x) is the measurement noise which maybe correlated to the backaction noise. In general, the forces and measurement noise can all have non-stationary dynamics, non-Gaussian noise and non-linear dependence on the oscillator position. They can each be linearized by Taylor expanding around the mean position of the oscillatorx ¯ and retaining only zeroth and first order terms. The zeroth order terms are independent of fluctuations in the oscillator position, forcing, and mea- surement noise; and only serve to deterministically shift the mean position of the oscillator. The first order terms are each linearly dependent on only one source of fluctuation; and either act to modify the mechanical susceptibility or introduce incoherent driving. Higher order terms introduce nonlin- earities and instabilities which can give rise to detrimental effects such as saturation and nonlinear dissipation (49, 65). By neglecting these higher order terms, the analysis is restricted to the most gen- eral linear oscillator experiencing linear feedback. It is important to note that the deterministic shift in mean oscillator position due to the zeroth order terms can affect the force resolution; for example, by shifting a cavity optomechanical system onto optical resonance. However, since this is deterministic, and known a priori, an equivalent displacement may be made to the oscillator without feedback by applying a known external force, as depicted in Fig. 5.1A. In this chapter a straightforward theoretical approach is presented which shows that, even in the presence of non-Gaussian or correlated noise and non-stationary processes, a real-time estimation strategy always exists that reproduces the same measurement record as any arbitrary linear feedback protocol (see Fig. 5.1A). The theory applies to both quantum and classical oscillators, and to the

1The term “oscillator” within the engineering community sometimes implies self-oscillation. Here, it is only referring to harmonic motion

36 intrinsically non-linear problem of variance estimation (56). It ultimately provides a clear set of minimum requirements for feedback to provide any advantage to force sensing over that possible with estimation alone. Essentially, some form of nonlinearity is required in the physical system, which may arise from the measurement process, feedback loop, signal, or from the oscillator itself. In the absence of non-linearities or uncertainty about the oscillator dynamics, the theory yields a filter which allows the estimate that would be obtained with feedback to be determined causally from the measurement record without feedback. This precludes the possibility of any additional sensitivity or resolution enhancement from feedback in either of the examples discussed above (56, 165), or indeed feedback improved bandwidth (110) in linear force sensors.

Ω Ω

Figure 5.1: A Conceptual diagram comparing optomechanical force sensing with feedback and filtering B Experimental schematic. Red (dark grey): fiber interferometer; solid green (light gray): electrical compo- nents for feedback stabilization; dashed green: electrical signal applied to the microtoroid for application of the incoherent force and feedback. FPC: fiber polarization controller. BPF: bandpass filter. MZM: Mach-Zender Modulator. PM: phase modulator. PI: proportional-integral controller. Figure from Ref. (66)

To validate the theoretical results, the effect of feedback enhanced force resolution achieved in Ref. (56) is experimentally reproduced, but replacing feedback with causal filtering. This demon- strates that the resolution enhancement achieved in that experiment was not due to a feedback-induced change in coupling of the oscillator to its the environment. Rather it arises from using an estimator that performs better for low quality, and thus feedback cooled, oscillators. By clearly demarcating the circumstances in which feedback may advantage force sensing over estimation alone, our results both clarify a significant ambiguity in the optomechanics and force sensing communities, and con- tribute towards simplifying experimental implementations of ultra-precise force sensing with linear oscillators.

5.2 Theoretical Treatment of Stationary Dynamics

To simplify the analysis and present results most relevant to our experiments, in this section a com- mon scenario is considered, where the mechanical oscillator’s susceptibility is only modified by the feedback force, and therefore drop the first order susceptibility modifying terms in the other forces.

37 Furthermore, all processes involved are assumed to be stationary. These specific assumptions are not necessary for the overall conclusions of this chapter (see general case in Sec. 5.3), which hold for the most general linear case, including non-stationary processes and first order terms. Under these assumptions, the combined force is given by Fm(t, x) = Fm(t, x¯) and the feedback force is

Z t Fact(t, x˜) = g(t − τ)x ˜(τ)dτ (5.2) −∞ wherex ˜(τ) = x(τ) + N(t, x¯) and g(t − τ) is the stationary feedback kernel describing the filter applied to the measurement record. Enforcing causality, namely g(t−τ) = 0 for τ > t, simplifies Eq. (5.2) into the Fourier convolution Fact(t, x˜) = g(t) ∗ x˜(t). Substitution into Eq. (5.1) and Fourier transforming then gives

  x(ω) = χ(ω) Fm(ω, x¯) + g(ω)x ˜(ω) (5.3)

−1 h 2 2 i where χ(ω) = m ωm − ω + iΓmω is the intrinsic mechanical susceptibility. The oscillator position without feedback can be simply obtained by omitting the feedback force g(ω)x ˜(ω) from Eq. (5.3), x0(ω) = χ(ω)Fm(ω, x¯0), where the subscript “0” is used to distinguish from the feedback case.

As discussed earlier, an external force may be applied to equate the mean positions of the oscillator with and without feedback. In the case of cavity optomechanics this amounts to ensuring identical cavity detunings when experiments are initiated. Withx ¯ = x¯0, the measurement noise and the common forcing terms with and without feedback are identical. Substituting for x(ω) and x0(ω) in terms of their respective measurement records (eg. x(ω) = x˜(ω) − N(ω, x¯)) then gives a completely deterministic equation relating the time domain measurement records that would be achieved with and without feedback " # 1 x˜(ω)= x˜ (ω)=h(ω)x ˜ (ω) (5.4) 1 − χ(ω)g(ω) 0 0 where h(ω) = χ0/χ is the ratio of the modified mechanical susceptibility χ0 to the intrinsic mechanical susceptibility. Therefore the exact position record that would be obtained using stationary linear feedback can be retrieved straightfowardly by applying the filter h(ω) to the position record without feedback. This precludes enhancements of resolution (56), bandwidth (110), or sensitivity beyond that achievable with estimation alone. Since no constraints are placed on the statistics of the driving forces or measurement noise, this result is valid even for non-Gaussian noise or if correlations exist between measurement and process noise, such as those induced by quantum backaction. Furthermore, since it applies directly to the measurement records, rather than a specific parameter estimation process based on them, it holds for both linear and non-linear estimation processes. In Sec. 5.3 the theory is generalized to include linear non-stationary forcing and feedback as well as modifications to the mechanical susceptibility due to effects such as optomechanical dynamical backaction (12, 36, 75). In this case, the required filter is more complex and is, in general, non-stationary, but remains causal. Our results are valid in both the quantum and classical regime and rigorously prove that no force sensing advantage is provided by linear feedback onto a linear oscillator with known dynamics.

As clearly outlined, the results derived above do not apply if nonlinearities are present in the

38 system. This includes, for example, measurement nonlinearities (110) and nonlinear interactions with the environment which result in oscillator frequency drifts (57). Feedback enhanced transduction sensitivity has, in fact, recently been demonstrated in a cavity opto-electromechanical system with measurement nonlinearities introduced by radiation pressure backaction (65). It should also be noted that enhanced force sensing can be achieved via coherent control combined with measurement. In optomechanics, for example, coherent control is predicted to allow the standard quantum limit of force sensing to be surpassed (11).

5.2.1 A Heuristic Approach to Stationary Dynamics

In the case of a classical linear oscillator with stationary system dynamics and stationary feedback, the result of the previous section can be straightforwardly understood by considering the measurement record in the Fourier domain. An estimate of the force applied to the oscillator can be obtained by taking a weighted sum of all frequency components that contain information about the signal. The resolution of the sensor, with or without feedback, is ultimately determined by both the weighting function and the signal-to-noise ratio of each frequency component, henceforth termed the spectral SNR. As shown in Ref. (56) using feedback, and in this chapter using filtering, improved sensitivity can be achieved by modifying the weighting function. For a classical linear oscillator with stationary feedback the spectral SNR is independent of linear feedback (see Fig. 5.2 which illustrates this for white force sensing). Consequently, since any weighting function that can be achieved with feedback, can also be achieved using filtering, it is clear that feedback offers no advantage in stationary classical force sensing. The effect of filtering, or equivalently perfect noiseless feedback, on the spectral SNR and spectral weighting is shown for the case of white force sensing in Fig. 5.2A-C. The thermal noise (red), incoherent signal (blue), and shot noise floor (dashed line) are all equivalently suppressed resulting in an unchanged spectral SNR with feedback. Note also, that the bandwidth over which the signal is greater than the combined noise floor is also unchanged with feedback and filtering. Since the spectral SNR is constant across the region of frequency space where shot-noise is negligible, it is apparent that flattening the weighting function across this region will enhance the sensitivity of incoherent force measurement. Feedback cooling as demonstrated in Ref. (56), or the equivalent filter demonstrated in Sec. 5.2, achieves this goal, with its effect shown in Fig. 5.2B. However, this simple approach also enhances the contribution of shot noise, as can be seen in the wings of Fig. 5.2C, where the shot- noise (dashed line) exceeds even the signal amplitude at the mechanical resonance frequency. As a result of this relative “squashing” of the signal on resonance, even though the weighting function becomes increasingly flat with increasing gain, at sufficiently high feedback/filter gain the sensitivity of the measurement begins to degrade (as shown later in Fig. 5.4). In such circumstances, when using filtering, further sensitivity enhancements could be gained by modifying the weighting function away from that naturally arrived upon with broadband feedback, and instead using a weighting function that suppresses the wings of the spectrum. This case is not considered here, allowing a direct comparison between the enhancement with filtering and that achieved with feedback in Ref. (56).

39 htfebc nue nacmnsi estvt,bnwdho eouinaeacrtl attributed accurately are resolution or bandwidth sensitivity, in ensures enhancements This nonlinearity. induced of feedback form that some experiencing included sensors e is of equally discussion performance is the brief that characterizing a when estimator results, is an general kind using these of second of importance light the the In of emphasizing more section. equation this a Fredholm in Rather, undertaken a as of required, noise. discretization temporal measurement involving or approach experience coupling, technical that environmental systems non-stationary feedback, to changing straightforwardly case in dynamically extended record the be measurement in cannot the taken domain, considering approach Fourier by simple estimation the to The feedback comparing non- 75). dynamics, 36, linear stationary (11, of include e backaction to to due dynamical susceptibility generalized optomechanical mechanical is the as to such section modifications as previous well as the feedback and in forcing presented stationary result theoretical the Here, Dynamics Non-Stationary of Treatment Theoretical 5.3 ihtesprto fteelnsgvn h adit ftemaueet iuefo supplementary from lines, Figure grey vertical measurement. two com- the the the of between than bandwidth region greater the (66) the is giving Ref. over signal lines of noise The information these about shot of gain. information and separation filtering mea- extractable noise the the the of thermal with of as amount from independent process the floor is filtering denoting noise signal, bined regions, same the shaded the and noise the to mechanical thermal of subject the the area floor indicate The noise lines Solid the record. gain. indicate surement increasing lines with dashed spectra thermal and the susceptibility to applied is filtering tionary A 5.2: Figure hoeia hra ipaeetsetu rd iha noeetsga (blue). signal incoherent an with (red) spectrum displacement thermal Theoretical

δωπ 40 ff ciewt n ihu feedback without and with ective B-C Sta- ff ects to the control of nonlinearities and beyond that attainable with estimation alone.

The generalized optomechanical system considered here is shown schematically in Fig. 5.3. As was defined in Sec. 5.1, the evolution of the mechanical oscillator can be described, in both classical and quantum regimes (95) by the following equation of motion

h 2 i meff x¨ + Γm x˙ + ωm x = Fm(t, x) + Fact(t, x˜) (5.5) where meff, Γm, and ωm are the mass, damping rate, and resonance frequency of the mechanical oscil- lator, respectively. For compactness, the force Fm = FT(t, x) + Fs(t, x) + Fba(t, x) combines the ther- momechanical force FT(t, x), a signal force Fs(t, x) and measurement backaction Fba(t, x). Fact(t, x˜) is a linear actuation force wherex ˜(t) = x(t) + N(t, x) is the instantaneous measurement record of the oscillator position x, and N(t, x) the measurement noise. In general, the forces and measurement noise can all have non-stationary dynamics, arbitrary correlations, non-Gaussian noise and non-linear dependence on the oscillator position.

Figure 5.3: General theoretical schematic outlining the emergence of forces and feedback in a cavity optomechanical system. Figure from supplementary information of Ref. (66)

The combined thermal, signal, and backaction force may be expanded without loss of generality,

41 as

Z t   Fm(t, x) = dτ gm(t, τ, x) + nm(t, τ, x)ξm(τ) (5.6) 0 Z t  0  = dτ gm(t, τ, x¯) + gm(t, τ, x¯)(x(τ) − x¯) + nm(t, τ, x¯)ξm(τ) (5.7) 0 Z t  0  = dτ Gm(t, τ, x¯) + gm(t, τ, x¯)(x(τ) − x¯) (5.8) 0 where the first term in Eq. (5.6), gm(t, τ, x), is a generalized gain kernel that allows an arbitrary coher- ent response at time t, dependent on the full prior history of the oscillator position x. The second term includes all incoherent forcing with ξm(τ) being a unitless white-noise Wiener process and nm(t, τ, x) a memory kernel that permits non-Markovian, non-stationary and position dependent noise. The ac- tuation force can be expressed similarly by replacing x withx ˜ in Eq. (5.6). In general, both terms in Eq. (5.6) can introduce nonlinearities and instabilities which can give rise to detrimental effects like saturation, nonlinear dissipation (49, 65) or frequency drifts of the oscillator, for example, from spatially varying short-range forces in atomic force microscopy (20, 86). Here, the goal is to show that no advantage is possible from feedback in the case of a linear oscillator and linear feedback.

Consequently, both memory kernels gm(t, τ, x) and nm(t, τ, x) are expanded about the mean position 0 of the oscillatorx ¯ and neglect nonlinear terms. This gives gm(t, τ, x) = gm(t, τ, x¯) + gm(t, τ, x¯)(x − x¯) 0 where gm(t, τ, x¯) = ∂gm(t, τ, x)/∂x|x=x¯ for the coherent memory kernel. Since the incoherent mem- ory kernel is multiplied by the noise term ξm(τ), nonlinearity occurs even in the second term of the expansion, so that for a linear oscillator nm(t, τ, x) = nm(t, τ, x¯). Substituting these expressions into Eq. (5.6) results in Eq. (5.7), which can be further simplified into Eq. (5.8) by grouping all terms without dependence on the instantaneous displacement of the oscillator from its mean position into a single term Gm(t, τ, x¯)=gm(t, τ, x¯)+nm(t, τ, x¯)ξm(τ) containing both noise and deterministic forcing.

The actuation force Fact can be similarly Taylor expanded into its component force as

Z t 0 Fact(t, x˜) = dτGact(t, τ, x¯) + gact(t, τ, x¯)(x ˜(τ) − x¯). (5.9) 0

0 wherex ˜(τ) = x(τ)−N(t, x¯). Here, the term gact(t, τ, x¯)(x(τ)−x¯) provides linear feedback modifying the susceptibility of the oscillator and driving its motion with white measurement noise N(t, x¯). Taking the case of ideal feedback where no noise is introduced by the feedback loop itself, Gact(t, τ, x¯) = gact(t, τ, x¯), acts only to shift the mean position of the oscillator independent of the measurement record. Substituting the expressions for the actuation and combined forces into Eq. (5.5) and Fourier transforming gives

( Z t 0 x(ω) = χ(ω)Ft→ω dτGm(t, τ, x¯) + gm(t, τ, x¯)(x(τ) − x¯) 0 ) 0 + Gact(t, τ, x¯) + gact(t, τ, x¯)(x ˜(τ) − x¯) (5.10)

−1 h 2 2 i where Ft→ω represents the Fourier transform and χ(ω) = meff ωm − ω + iΓmω is the intrinsic me-

42 chanical susceptibility. It should be noted that the Fourier transform can only be taken if the oscillator is intrinsically stable and the energy is finite. Consequently, scenarios such as regenerative ampli- fication where feedback gives rise to exponential growth in displacement and eventually triggers a nonlinear response (75, 155) are excluded from the analysis here. However, even in unstable situa- tions, in the physically realistic case where the measurement is performed over a finite time interval the energy does remain finite, and the Fourier transform can be applied.

If feedback is not applied, both the linear feedback term and the feedback induced deterministic displacement are removed. In principle, the resulting change in mean oscillator displacement could modify the force resolution. For example, in cavity optomechanics, the displacement could move the optical cavity toward resonance, modifying the optical power level in the cavity and through this the resolution. However, since the feedback induced displacement is deterministic and known, an equivalent displacement may be achieved without feedback by applying an external force. In the case of cavity optomechanics, operationally, this corresponds to ensuring the cavity is detuned from resonance equivalently in both the feedback and non-feedback scenarios at the beginning of the measurement. Applying this external force so thatx ¯ = x¯0, the oscillator position without feedback 0 can then be simply obtained by omitting the feedback force gact(t, τ, x¯)(x(τ) − x¯) from Eq. (5.10)

(Z t ) 0 x0(ω) = χ(ω)Ft→ω dτGm(t, τ, x¯)+gm(t, τ, x¯)(x0(τ) − x¯)+Gact(t, τ, x¯) (5.11) 0 where the subscript 0 is used to distinguish from the feedback case.

Subtracting Eq. (5.10) from Eq. (5.11) eliminates the common terms Gact(t, τ, x¯) and Gm(t, τ, x¯) and, after substituting x(ω) = x˜(ω) − N(ω, x¯), results in a completely deterministic equation relating the time domain measurement record with and without feedback

( (Z t0 )) −1  0 0 0 0  0 x˜(t)−Fω→t χ(ω)Ft →ω dτ gm(t,τ,x¯)+gact(t,τ,x¯) (x ˜(τ)−x¯) 0 ( (Z t0 )) −1 0 0 0 = x˜0(t)−Fω→t χ(ω)Ft →ω dτgm(t,τ,x¯)(x ˜0(τ)−x¯) (5.12) 0 where the dummy variable t0 has been introduced to distinguish the Fourier transform from its in- verse. Eq. (5.12) gives a non-trivial relationship between the measurement records with and without feedback in the presence of non-stationary forcing and feedback, non-Gaussian noise and correla- tions between measurement and process noise. However, for feedback and filtering to be equivalent, a causal filter must exist that mapsx ˜0 7→ x˜. To determine the existence of such a filter it is necessary solve forx ˜(τ) as a function ofx ˜0(τ) in Eq. (5.12).

Without loss of generality Eq. (5.12) may be simplified by choosing our position coordinate such that the mean displacement is zerox ¯ = 0. The second term on the LHS of Eq. (5.12) may then be

43 expanded as

( (Z t0 )) −1  0 0 0 0  Fω→t χ(ω)Ft0→ω dτ gm(t , τ, x¯)+gact(t , τ, x¯) x˜(τ) 0 Z ∞ Z t0 0  0 0 0 0  0 = dt dτ gm(t , τ, x¯)+gact(t , τ, x¯) x˜(τ)χ(t − t ) (5.13) −∞ 0 Z ∞ "Z ∞ # 0  0 0 0 0  0 = dτ dt gm(t , τ, x¯)+gact(t , τ, x¯) χ(t − t ) x˜(τ) (5.14) 0 −∞ Z ∞ = dτ (hm(t, τ, x¯) + hact(t, τ, x¯)) x˜(τ) (5.15) 0

R ∞ 0 iω(t−t ) − 0 where the time shift property of Fourier transforms −∞ dωe χ(ω) = χ(t t ) has been used to 0 0 0 obtain Eq. (5.13). Eq. (5.14) is found by enforcing causality, namely g j(t , τ) = 0 when τ > t , allowing the upper bound of the integral over τ to be extended to infinity and the order of integration to be rearranged. The term contained in the large square brackets of Eq. (5.14) is a combined transfer function for the system which can be consolidated into terms denoted hm(t, τ, x¯) and hact(t, τ, x¯) that are zero when τ > t, resulting in Eq. (5.15). Applying the same procedure to the RHS of Eq. (5.12) simplifies the equation into a Fredholm equation of the second kind

Z ∞ Z ∞ x˜(t) − dτ (hm(t, τ, x¯) + hact(t, τ, x¯)) x˜(τ) = x˜0(t) − dτhm(t, τ, x¯)x ˜0(τ). (5.16) 0 0

The kernel of the LHS is bounded and zero for τ > t, so it is square integrable. Fredholm’s theorem therefore guarantees the existence of solutions forx ˜. In practice, without exploiting symmetries or as- sumptions about the kernel such equations are typically solved using numerical techniques (73, 127). Subsequent discretization transforms both integral equations into matrix form that can be inverted to give the causal filter. Performing the temporal discretization on Eq. (5.16) gives

[I − (Hm + Hact)] x˜ = [I − Hm] x˜0 (5.17) where I is the identity matrix, x˜ is a measurement record vector and the matrices H are the discretized kernels of Eq. (5.16). The inner product between H andx ˜ effects the integration. Consequently it is possible to solve for x˜ to finally give

−1 x˜ = [I − (Hm + Hact)] [I − Hm] x˜0 (5.18)

This expression shows that, for a linear oscillator, it is possible to exactly reproduce the measurement record that would be obtained with from a system with non-stationary processes by causally filter- ing the measurement record without feedback. This result is valid in both the quantum and classical regime and rigorously proves that no force sensing advantage is provided by linear feedback onto lin- ear oscillators over that possible with estimation alone. This precludes enhancements from feedback of both force resolution (56) and bandwidth (110) beyond that achievable with estimation alone.

44 5.4 Force Sensing in the Presence of Nonlinearities

This section provides further discussion on characterizing the enhancement provided by feedback, particularly when nonlinearities are present in the system. A common form of nonlinearity is satura- tion in the measurement process, which can restrict the dynamic range of sensors without feedback. Such a nonlinearity is reported in Ref. (110), where measurement nonlinearities are exhibited in a mi- croelectromechanical force sensor with optical readout, even when the mechanical motion is due only to thermal excitation of the mechanical mode. In that experiment, stationary feedback cooling was used to suppress the nonlinearity, with a three order of magnitude enhancement in bandwidth reported. However, this enhancement was arrived at by comparing the full-width-half-maximum (FWHM) of the mechanical susceptibility with and without feedback. This approach, in general, significantly over-estimates the enhancement. This can be seen, in the light of the results reported herein, by con- sidering the case of feedback cooling on a purely linear oscillator. In this case, our results show that the feedback provides no information advantage over estimation alone. However, the cooling can, in principle, arbitrarily increase the FWHM, leading to the inaccurate conclusion of a greatly increased bandwidth. The bandwidth would be better defined as the frequency range over which the signal-to- noise ratio is larger than unity. Using this definition, it can be seen that no bandwidth improvement is possible if the system is linear, consistent with our results. Any improvement in bandwidth obtained, may then be accurately attributed to the action of feedback in suppressing the nonlinearity. Nonlinear interactions with the environment may also degrade the function of force sensors. Such interactions are especially important for ultraprecise sensors, which have high mechanical quality factors and low oscillator mass. Ref. (56) is a case in point, achieving impressive absolute force reso- lution at the level of 15 aN Hz−1/2, limited by fluctuations in the intracavity power which constrained the measurement duration. Such fluctuations have a number of consequences. Most obviously, they cause a linear and dynamical change in transduction sensitivity. In Sec. 5.3 it is shown that feedback to the oscillator does not offer any advantage over estimation in accounting for such linear fluctua- tions in transduction sensitivity. It should be noted, however, that if a separate direct measurement of a parameter proportional to the intracavity power (such as the laser output power) was made, that measurement might allow the fluctuations to be stabilized (for example, via feedback to the laser intensity) and therefore an improved measurement to be achieved. In addition to linear changes in transduction sensitivity, fluctuations in intracavity power may cause nonlinear effects such as heating induced changes to the mechanical resonance frequency. It is conceivable that feedback to the oscil- lator may be advantageous if such nonlinear effects act to limit the measurement resolution. Although in Ref. (56) the reported feedback enhanced resolution is fully consistent with linear feedback upon a linear oscillator, and may be understood as a consequence of using an estimator that performs better for low quality oscillators; it might be expected that feedback could play a crucial role in future ex- periments with these devices. It should be emphasised that, similar to the experiments in Ref. (110), an estimator that is equally effective with and without feedback would be crucial to draw accurate conclusions regarding the information gain provided by feedback in such experiments. Due to their ultrahigh quality factor, ultrasensitive nanomechanical force sensors typically exhibit

45 greatly enhanced motional response to applied forces at the mechanical resonance frequency. In addition to increased sensitivity to environmental fluctuations, this presents the requirement, when feedback is not applied, of very high measurement dynamic range (in the case of Ref. (56), 50dB). Feedback cooling can be used to greatly reduce dynamic range requirements. Therefore it not only provides the possibility to enhance the information gain in nonlinear force sensing, but also, in some circumstances, allows the use of less sophisticated detectors thereby simplifying the experimental apparatus.

5.5 Experimental Results: Stationary Dynamics

Recently, enhanced incoherent force resolution was experimentally demonstrated (56) by stationary feedback cooling of a linear oscillator. However, as shown here, no force resolution enhancement is obtained from this method over estimation alone. The exact filter equivalent to the feedback cooling used in Ref. (56) is obtained by substituting g(ω) = −imeffΓmωgf into Eq. (5.4), where gf represents the filter’s unitless gain. This filter, denoted hc(ω), effects the causal mapx ˜0 7→ x˜. Here, this is experimentally demonstrated in a similar system to that of Ref. (56) consisting of a microtoroidal cavity optomechanical system. An intrinsic mechanical mode of the microtoroid is used to transduce an incoherent electrostatic gradient force applied by a nearby electrode (85). A whispering gallery optical mode of the microtoroid is used to read out the mechanical motion and thereby determine the variance of the incoherent force. Our experimental setup is shown in Fig. 5.1B. A shot-noise limited fiber laser at 1550 nm was evanescently coupled into the whispering gallery mode of the microtoroid using a tapered optical fiber. The microtoroid had major and minor diameters of 60 µm and 6 µm respectively with a 26 µm undercut. The mechanical motion of the microtoroid, which induces phase fluctuations on the trans- mitted light, was detected interferometrically by beating with a bright 3.5 mW optical phase reference followed by shot-noise limited homodyne detection. The toroid-taper separation is actively stabilized using an amplitude modulation technique (34) that maintains a constant coupling rate into the opti- cal cavity. Pound-Drever-Hall locking was used to lock the laser frequency to the optical resonance, 7 which had an intrinsic quality factor of Q0 =2.6 × 10 . A 50/50 tap-off after the microtoroid was used to detect the cavity transformed amplitude and phase modulation, providing the error signal for the optical frequency and taper-toroid separation locks. The interferometer was locked midfringe via a piezo actuated fiber stretcher that precisely controls the optical path length in one arm. The measurement record is acquired from the homodyne signal by electronic lock-in detection where demodulation of the photocurrent at the mechanical resonance frequency allows real time measurement of the slowly evolving quadratures of motion, denoted I(t) and Q(t) where x(t) =

I(t) cos(ωmt) + Q(t) sin(ωmt). Fourier analysis reveals a mechanical power spectra with peaks corre- sponding to microtoroid mechanical resonances. Fig. 5.4A (red) shows the room temperature Brown- ian motion of a mechanical mode with a signal-to-noise ratio of 37 dB and a fundamental frequency, damping rate and effective mass of ωm = 40.33 MHz, Γm = 23 kHz and meff = 7 ng respectively. The absolute mechanical displacement amplitude was calibrated via the optical response to a known ref-

46

D δωπ

4 δωπ

4

4

44 4

Figure 5.4: A Displacement spectrum at room temperature [dark grey (red)] and with filtering [light grey (green)] at gain 2, 8, 34, 72 and 150. Inset top: displacement spectrum with feedback cooling [black (blue)] and filtering [light grey (green)] with gain 1. Inset bottom: percentage difference between filtered and cooled spectrum with equivalent gain. B Force resolution as a function of averaging duration for thermal [dark grey (red)], feedback cooled [black (blue)] and filtered spectra [light grey (green)] at gain 1, 2.4, 5 and 10. Solid line (red): fit to inverse quartic dependence of the force resolution without filtering or feedback. C Force resolution versus filter gain after 1 ms of averaging (green circles) showing good agreement to theory (dashed line). Figure from Ref. (66)

erence phase modulation (141).

As shown earlier, applying the filter hc(ω) to the measurement record without feedback should retrieve an identical measurement record to that obtained with feedback. Applying this filter to the measurement record it is possible to mimic feedback cooling as shown in Fig. 5.4A (green) where the

filter gain gf is varied from 2 to 150. Extending the gain beyond gf > 30 the mechanical spectrum in- verts and exhibits squashing, a well known characteristic of high gain feedback cooling (85). To con- firm the equivalence of the measurement record obtained via feedback and filtering, feedback cooling is implemented by applying the homodyne photocurrent to the toroid through an electrode which gen- erates strong electrical actuation of the mechanical motion through electrical gradient forces (101). This allows the mechanical mode to be cooled from room temperature by a factor of 2. The upper inset in Fig. 5.4A shows feedback cooling (blue) and equivalent gain filtering (shaded green); with the fractional difference between the feedback and filtering spectra showing no statistically significant difference (lower inset). An estimate of the variance of an incoherent force applied to an oscillator may be obtained by

47 determining the oscillator’s energy (56). After averaging time, τ, the estimate of the energy is given R τ by E = 1/τ dtI(t)2 + Q(t)2. To calculate the ensemble average hE i and the standard deviation τ 0 τ σE(τ) of the energy estimate multiple independent measurements are made for each τ. Following Ref. (56) the energy estimate can be translated into an estimate of the magnitude of the force with a R ∞ resolution given by δF(τ)2 =σ (τ)/ dω|χ0(ω)|2. It is important to note that this estimation process E 0 −1/4 is not necessarily optimal. In the case where τ > 1/Γm the force resolution scales as (Γmτ) which appears to motivate the use of feedback cooling to increase the mechanical decay rate Γm (56). Figure 5.4B (red points) shows the inverse power-law dependence of the force resolution on av- eraging duration, τ, for our experiments with only thermal driving. As predicted by our theory, by applying the filter hc(ω) to the thermal data it is possible to enhance the force resolution in the same way as feedback cooling. This is shown in Fig. 5.4B (green) where increasing the filter gain, gf, provides a clear improvement in force resolution while consistently maintaining the predicted power- law dependence with respect to averaging time. Figure 5.4C (circles) shows the force resolution as a function of filter gain taken for a fixed averaging duration of τ = 1 ms. For gains below gf = 20 the measured force resolution agrees with the theoretical fit. At higher gains the degradation in sen- sitivity arises from squashing of the mechanical power spectrum. As discussed in Sec. 5.2, this acts to suppress spectral components close to the mechanical resonance frequency relative to components further from resonance where shot noise dominates.

Figure 5.5: A Normalised energy estimate at room temperature [light grey (red)], and with addition of incoherent driving [dark grey (green)]. Solid lines: mean; dashed lines: one standard deviation error bounds. B Averaging time required to resolve incoherent signal versus filter gain (green circles) showing good agreement to theory (dashed line). Figure from Ref. (66)

To demonstrate the improvement in force resolution achievable via filtering a small incoherent electrostatic gradient force is applied to the microtoroid with a magnitude of approximately 7% of the thermal energy. The ability to resolve this force against the thermal noise depends on the averaging time. Only when the standard deviation of the energy estimate is smaller than the strength of the signal can the incoherent force be resolved, or equivalently the time-integrated sensing noise power must be less than the signal force noise. The convergence of the thermal energy estimate with increasing averaging time is shown in Fig. 5.5A (red). With the addition of the incoherent signal the ensemble average is increased without affecting the error bounds as shown in Fig. 5.5B (green). At 3 ms the

48 error becomes comparable to the energy separation and the applied incoherent force is resolved. If the filter hc(ω) is applied the force resolution is improved and the time taken to resolve the applied force decreases as shown in Fig. 5.5B. The required averaging time decreases as the estimation gain is increased in good agreement with theory. At gains g f > 20 the averaging time increases again due to inversion of the mechanical spectrum and suppression of the signal relative to shot noise. The inflection point in Fig. 5.5B and Fig. 5.4C shows that even though thermal noise dominates shot noise by orders of magnitude at the peak of the mechanical susceptibility, in incoherent force sensing it is shot noise that determines the ultimate force resolution limit.

5.6 Conclusion

The experiments presented here show that feedback and filtering allow equivalent enhancement in in- coherent force resolution. In this context, the results of Ref. (56), can be naturally understood to result from using an estimator that performs better for low quality oscillators. Near-resonant spectral com- ponents of the incoherent force drive the oscillator more strongly, and are therefore over-represented in the measurement. As a result, even though feedback only applies a reversible transformation to the measurement record, the force resolution appears to improve with increasing oscillator linewidth as the estimation becomes more balanced. The stroboscopic feedback enhanced force sensing sce- nario proposed in Ref. (165) is similarly biased to low quality oscillators. In that case, measurements prior to application of the signal force are used to pre-cool the oscillator. This improves its initial localization in phase space and, thereby, the capacity to resolve displacements due to external forces. However, the existence of this prior measurement record is not taken into account when calculating the signal to noise ratio of force measurements without feedback. Filtering it appropriately allows equivalent localisation to feedback cooling, though offset from the origin, and achieves an identical signal to noise ratio. These examples illustrate the main result of this chapter that, for a linear os- cillator, any advantage in force measurement arises not through the action of feedback, but rather through measurement and estimation alone. During review of the paper this chapter is based on, a comment was published (163) on Ref. (56) which reaches similar conclusions but limited to the case of stationary dynamics driven by uncorrelated white noise.

49 Chapter 6

Superfluid Optomechanics

This chapter reports the use of cavity optomechanics to directly probe the of super- fluid excitations in real time. Specifically, the thermomechanical fluctuations of third sound modes in a nanoscale superfluid helium-4 film are observed. Detuned laser driving allows laser cooling and heating, while amplitude-modulated driving allows the non-linear Duffing interactions between phonons to be probed with as few as thirty seven intracavity sideband photons. The results presented here demonstrate some of the unique prospects of superfluid films for cavity optomechanics, includ- ing strong optomechanical coupling, femto- to pico-gram effective mass, high mechanical quality factor and strong phonon-phonon interactions; potentially enabling the realization of macroscopic non-classical states (131), superfluid phononic circuits (46, 168), optomechanics with quantized vor- tices (62), and applications in superfluid force and inertial sensing (43). The content of this chapter has been submitted and is currently under review. Figures 6.1, 6.2 and 6.3 were made by David L. McAuslan.

Glen I. Harris, David L. McAuslan, Eoin Sheridan, Zhenglu Duan and Warwick P Bowen , “Laser Cooling and Control of Excitations in Superfluid Helium,” Submitted

6.1 Superfluid Helium-4

Emergent quantum phenomena such as superconductivity, quantum magnetism, and superfluidity arise due to strong interactions between elementary excitations in condensed matter systems (159). In this context the observation of zero viscosity flow of liquid helium-4, coined superfluidity in 1937 by P. Kapitsa, J. Allen and D. Miesner, provided one of the first experimental observation of many-body quantum physics; initiating over 70 years of fundamental research towards developing microscopic models of interacting condensates (87, 159). In superfluid helium-4, the elementary excitations are phonons and rotons; with techniques such as neutron and light scattering used to probe their behaviour since the 1960s (26, 159). However, quite generally, such techniques have been limited to measure- ments of average properties of bulk superfluid (159); or to observations of the driven response far out of thermal equilibrium (43, 50, 70). Consequently, the microscopic behaviour of strongly inter- acting condensates, like superfluid helium-4, is not fully understood. In this context, the ability to probe excitations in real time may provide a new approach towards understanding the microscopic

50 behaviour of superfluids, including phonon-phonon interactions (159), quantum vortices (122) and two dimensional quantum phenomena such as the Berezinskii-Kosterlitz-Thouless transition (159).

6.1.1 Hydrodynamic Equations of Thin Films

Here the classical hydrodynamic equations that govern superfluid thin films will be derived, allowing resonance frequencies to be calculated for a given geometry, film thickness and substrate. Despite the success of this simplified linear model, the possibility of deriving generalized equations that describe nonlinear classical and quantum effects remain an open question. This is primarily due to the com- plexity of the Navier-Stokes equations in addition to the intrinsic nonlinear interaction of superfluid helium, most notably expressed in the nonlinear Gross-Piteavski equation. First, a low temperature film of liquid helium-4 is considered in the context of the two-fluid model, comprised of both normal and superfluid phases. In general for any fluid to exhibit third sound q 2η oscillations the thickness of the film must be larger than the viscous penetration depth ρω where η, ρ and ω are the viscosity, density and oscillation frequency respectively. The superfluid component, with near zero viscosity, has been shown to have third sound oscillations with films only a few atom q 2η layers thick (147). However the normal fluid component has a viscous penetration depth of ρω ≈ 1cm to 10µm (for ω ≈ 1Hz to 1MHz) which is much larger than the typical film thickness considered here h0 ≈ 1 − 100nm. Consequently the third sound modes are purely superfluid oscillations with the normal fluid component viscously clamped to the substrate. In this regime numerous hydrodynamic equations of varying complexity have been derived over the last 50 years (5, 15, 19, 81). Here, the simple method outlined in (166) is followed, considering a continuity equation where a perturbation in film height must be accompanied by mass flow for mass conservation,

∂δd ρ + dρ ∇ · v = 0 (6.1) ∂t s where δd are the thickness fluctuations of a film that has an equilibrium thickness d. The total mass density, ρ, combines the superfluid and normal fluid components ρs and ρn respectively. To retrieve a standard wave equation the velocity of the superfluid, v, must be expressed in terms of the film height fluctuations. The velocity is defined, from the quantum mechanical definition of mo- mentum, as pψ(r) = −i~∇ψ(r) where the superfluid is described by a ground state wavefunction 2 ψ(r) = ψ0 exp(iφ(r)). In this formalism the square magnitude of the wavefunction, |ψ0| , can be inter- preted as the superfluid density which maybe integrated over the volume to retrieve the total number of helium atoms residing in the film. The spatially varing phase φ(r) manifests as flow of the superfluid resulting in an expression for the velocity in terms of a phase gradient of the condensate

 pψ0 exp(iφ(r)) = −i~∇ ψ0 exp(iφ(r)) (6.2) ~ v = ∇φ(r) (6.3) m where m is the mass of the helium atom. This expression for the velocity leads to the interesting result that the flow of superfluid is irrotational i.e. ∇ × v = 0. To preserve the condition of irrotational flow,

51 superfluid vorticies must have a velocity profile that scales as 1/r, where r is the distance from the center, with a normal fluid core to prevent infinite flow at r = 0. Furthermore, since the wavefunction is single valued, namely the phase is uniquely defined everywhere, then by Cauchy’s residue theorem the contour integral about the vortex core leads to quantized circulation. The well known Josephson-Anderson equation (118), commonly used to describe superconduct- ing circuits, can be used to equate the evolution of the condensate phase to the chemical potential µ.

∂∇φ(r) ~ = −∇µ (6.4) ∂t = −( f ∇δh − s∇T) (6.5) where the definition of the chemical potential, ∇µ = f ∇δh − s∇T from Ref. (19), contains the gradi- ent of the van der Waals force from fluctuations in the film thickness f ∇δh, and a term quantifying the entropy and transverse temperature gradient s∇T. The latter term is commonly neglected in cal- culations due to the large thermal conductivity of superfluid helium which serves to thermalize any transverse temperature gradients. It has been shown that including temperature gradients primarily contributes to the dissipation of the film oscillation, leaving the resonance frequencies relatively un- affected (19, 166). Substituting the expression for the velocity Eq. (6.3) into the Josephson-Anderson equation Eq. (6.5) then yields,

∂v m = − f ∇δd (6.6) ∂t

With an expression for the velocity in terms of fluctuations in the film height the continuity equa- tion Eq. (6.1) can be differentiated and rearranged to give a standard wave equation.

∂2δd dρ f 0 = s ∇2δd (6.7) ∂t2 ρ where f 0 is now the van der Waals force per unit mass f 0 = f /m. The propagation speed of the q 0 q dρs f ρs αvdw surface wave is then C3 = ρ = ρ d3 , where αvdw is the van der Waals coefficient. It should be noted that the majority of results presented in this chapter are obtained well below the superfluid condensation temperature, where the density of superfluid is approximately equal to the total density

ρs ρ ≈ 1.

6.2 Superfluid Optomechanics with Thin Films

In cavity optomechanics, the coupling between optical and mechanical degrees-of-freedom is greatly enhanced by the presence of high quality optical and mechanical resonances. This has enabled the demonstration of quantum behaviour such as ponderomotive squeezing (27), coherent state trans- fer (162) and optomechanical entanglement (120); measurement precision approaching the funda- mental limit set by the Heisenberg uncertainty principle (148); and cavity optomechanical sensors of external stimuli such as mass, acceleration and magnetic fields (106). Excitations in superfluids have

52 BS a b LO Laser 1550 nm 3He Cryostat Probe Microtoroid 90/10 BS BS AM D Tapered Fibre D SA Scope D _

NA 70 μm

Figure 6.1: Optomechanics with superfluid helium films. (a) Illustration of third sound waves of a superfluid helium film coating a microtoroid. The third sound oscillations have phase velocity C3 = p −3 21 5 −2 3αvdwd , where αvdw = 2.65 × 10 nm s is the van der Waals coefficient for silica. (b) Shot-noise limited homodyne detection of superfluid helium optomechanics. A tapered fibre-coupled microtoroid is orientated face down inside the sample chamber of a helium-3 cryostat. Helium-4 gas is injected into the sample chamber at a pressure of 69 mTorr (at 2.8 K). AM - amplitude modulator, BS - beamsplitter, D - photodetector, NA/SA - network/spectrum analyser. The 90/10 BS probes the optical transmission through the microtoroid.

recently been identified as an attractive mechanical component for cavity optomechanics (2, 43); in- troducing unique features such as viscosity that approaches zero at absolute zero, quantized rotational motion and vortices (122), and strong phonon-phonon interactions (159); and providing a platform for quantum microfluidics (46), quantum computing with electrons (125), and nonclassical state gen- eration using the phonon-phonon interactions (153). In Ref. (43) a pressure wave in bulk helium acts as a gram-scale resonator, with the combination of high mechanical quality factor and mass provid- ing a path towards ultra-precise inertial sensors. However, the comparatively large mass prevents the observation or control of thermal excitations, and the manifestation of quantum effects. Another promising candidate for superfluid optomechanics has been proposed in Ref. (51), consisting of a fiber cavity immersed in superfluid helium, with the bulk modes of the superfluid confined by the cavity and co-localized with the optical mode. This system has predicted optomechanical coupling rate and effective mass of g0 = 1.8 kHz and meff = 1 ng respectively (2).

Here, an alternative approach to superfluid optomechanics is proposed and demonstrated based on femto- to pico-gram films of superfluid helium condensed on the surface of a microscale whispering gallery mode resonator (Fig. 6.1a). Superfluid films form naturally on surfaces due to the combination of ultra-low viscosity and attractive van der Waals forces. Excitations in such films, known as third sound (50, 70, 122, 152), manifest as perturbations to their thickness with the restoring force provided by the van der Waals interaction. The physical structure of the resonator provides a template for the self-assembling film, acting to confine third sound modes at microscale in two dimensions, while their radial dimension is determined by the film thickness. Third sound modes have been observed in helium films thinner than a single atomic layer (152), providing the prospect for extremely low effective mass.

53 6.2.1 Predicted Optomechanical Parameters for Thin Film Systems

While superfluid film optomechanical devices have the prospect to span many orders of magnitude in both cavity dimensions and film thickness; here, the specific case of a superfluid-coated micro- toroid (Fig. 6.1a) with major (minor) diameter D = 70 µm (d = 7 µm) is considered. The phase velocity C3 of third sound defines the mechanical resonance frequencies of the film, in concert with the microresonator boundary conditions. For example, low frequency major modes will likely ex- ist due to confinement around the major circumference (Fig. 6.1b), with periodic boundary condi- tions defining a fundamental major mode frequency ωD/2π = C3/πD = 13 kHz, for a 10 nm thick film. Minor modes may also be anticipated due to confinement with length-scale of approximately

Ld ≈ πd around the minor circumference. Here, intersection with the supporting disk results in a non-periodic boundary condition, giving a fundamental minor mode frequency of approximately

ωd/2π = C3/2Ld ≈ C3/2πd = 65 kHz. Furthermore, it is likely that the supporting disk will also de- fine drumhead modes that may weakly couple to the optical field around the periphery. For simplicity, henceforth I will only consider minor and major modes of the superfluid, which maybe visualized by considering a 2D membrane stretched over the microtoroid.

Assuming a film of uniform thickness, and considering the intersection of the 2 µm supporting disk, the total mass of helium surrounding the microtoroid is estimated to be mfilm = ρsV ≈ 7 pg −3 where the superfluid density is ρs = 145.1 kg m and the volume is V = 2πD × (2πd − 2 µm) × 10−8 = 4.4 × 10−17 m3. To estimate the effective mass a reducing factor is applied that is given by R 2 2 the integral over the normalized displacement xn, given by dV|xn| /max{|xn| }. First considering an infinitesimally thin shell of the superfluid film, it is known that the mode shape of a two dimensional membrane has approximately a sinusoidal shape, with an effective mass of 25% of the total mass. Furthermore, the amplitude of oscillation is known to linearly decrease as a function of depth through the superfluid film, leading to an effective mass equal to 12.5% of the total mass of the film. In the case considered here, this results in an effective mass of approximately one picogram, a twelve order-of-magnitude reduction compared with Ref. (43).

The close proximity of the film to the microtoroid surface enables strong near-field coupling (7) of the thickness perturbations to the optical mode. To calculate the single photon optomechanical coupling rate to the superfluid oscillation, the radial breathing mode of a microtoroidal resonator is first considered, with an optomechanical coupling rate given by g = ω x /R (162), where ω is the √ 0 L zp L laser frequency, R is the major microtoroid radius and xzp = ~/2meffωm are the mechanical zero point fluctuations. When the fundamental optical mode is located at the anti-node of the superfluid mode the optomechanical coupling rate will have the same form, however, reduced by the lower refractive index of the superfluid. This reducing factor can be derived using perturbation theory, calculating the shift in optical resonance frequency induced by the presence of a nearby dielectric. From Ref. (10) the fractional frequency shift of an optical mode E0(~r) arising from the introduction of a localized

54 dielectric n1(~r) is given by

R  2  2 ∆ω 1 n1(~r) − 1 |E0(~r)| d~r = − (6.8) 2 2 ω 2 n0Vmode|Emax| Z Ω υ  2  2 = − n1 − 1 |E0(~r)| d~r (6.9) 2Vmode where the geometrically dependent refractive indices (n0(~r) and n1(~r)) have been simplified to a con- stant value due to the reduction of the integral from the entire optical mode to just the volume of the 2 |Eout| localized dielectric. The optical mode volume is given by Vmode and the unitless parameter υ = 2 |Emax| is the ratio of the electric field at the surface of the WGM to the maximum electric field. In the case considered here the localized dielectric is a layer of superfluid helium with a refractive index nHe and a mechanical component that modulates the thickness of the layer. The optomechanical coupling rate ∝ d∆ω can then be expressed in terms of the rate of change of ∆ω with the depth of the film, namely g dd . Since a radial breathing mode of the microtoroid is equivalent to a modulated “film” of dielectric with the same refractive index as the resonator, then the reducing factor ζ can be shown to be given by

g n2 − 1 ζ SF He = = 2 (6.10) gRBM ns − 1 where gSF = ζgRBW is the optomechanical coupling rate of the superfluid mode relative to a radial breathing mode. Introducing this factor, for a 330 kHz mechanical mode, gives a superfluid single photon optomechanical coupling rate (denoted g0) of

g n2 − 1 ω 0 He L x = 2 zp (6.11) 2π ns − 1 R = 69 kHz (6.12)

where nHe = 1.028 and ns = 1.458 is the refractive index of superfluid helium and silicon, respectively. While this estimate na¨ıvely assumes perfect overlap between the optical and mechanical modes, a simple geometry could be conceivably designed to maximize such overlap. For example, a high order radial drumhead mode will have good optomechanical overlap to the optical modes of a silicon disk resonator, providing a promising direction for future work. Such a system would competitive with state-of-the-art cavity optomechanical systems (32).Furthermore, third sound dissipation rates as low as Γm/2π = 100 Hz have been observed in our experiments. Combined with experimentally realistic optical decay of κ/2π = 1 MHz (42), a single photon optomechanical cooperativity as high 2 as C0 = g0/κΓm ≈ 50 may be possible. By comparison, state-of-art optomechanical systems such as −1 ultra-cold atoms and phononic-photonic crystal cavities have C0 . 100 (27) and C0 ≈ 10 (32), respectively.

55 6.3 Measuring Superfluid Film Fluctuations

Techniques to probe thermal excitations have been crucial to our understanding of superfluids since the 1960s (26, 159). For instance, neutron and light scattering (159) allow the dynamic structure factor to be determined, which quantifies the dispersion relation, as well as the mean occupancy and correlations between modes. However, such techniques are slow compared to the mechanical decay rate, constraining them to measurements of average properties of the superfluid and prohibiting real-time measurement and control. Real-time measurements have previously only been performed by driving the superfluid far out of equilibrium, and have been constrained to coherent dynamics. The ability to resolve thermodynamical fluctuations in real-time, as will be demonstrated here, may allow new insights into less well-understood superfluid helium phenomena such as dissipation in thin films (70), quantized vortices (122), and the Berezinskii-Kosterlitz-Thouless transition (159). To experimentally realize thin-film superfluid optomechanics, a fiber coupled microtoroidal res- onator is placed in low a pressure helium-4 gas environment within a helium-3 cryostat (Fig. 6.1d). A typical optical resonance lineshape is seen at temperatures above the 1 K gas-to-superfluid transition (Fig. 6.2a (blue)). Below the superfluid transition, unstable oscillations appear on the blue detuned side of resonance (Fig. 6.2a (orange, green)), characteristic of optomechanical parametric instabil- ity (75). As the cryostat cools, helium condenses and the superfluid film thickens. This causes a reduction in the oscillation frequency which saturates at 10 kHz near base temperature, in agreement with the expected functional form (Fig. 6.2b). Identifying the observed parametric oscillations as the 1 −1 fundamental major mode yields the phase velocity C3 = 2 DωD = 2.2 ms at base temperature and a 12 nm film thickness, consistent with estimates based on the injected helium gas concentration and sample chamber surface area and volume. Network analysis is performed to investigate the superfluid mode structure, with an optical am- plitude modulation used to drive superfluid excitations (Fig. 6.2c). Both low (orange) and high (blue) frequency mechanical modes are observed, with the three high frequency modes exhibiting the har- monic spacing expected of third sound modes. Assuming the set of high frequency modes are associ- ated to confinement about the minor axis, the characteristic length scale can be determined from the frequency of the fundamental high frequency mode. The result is Ld = πC3/ωd = 7.3 µm, substan- tially smaller than the minor circumference but close to microtoroid minor diameter. At this point, it should be noted that the exact mode structure of the superfluid oscillations are unknown, primarily due to relatively large uncertainty in the thickness in addition to the complexity of modeling a 10 nm thick film on a non-trivial geometry that is orders of magnitude larger in size. It should also be noted that superfluid oscillations have been proposed previously as a possible mechanism for features observed in the driven response of a cryogenic microtoroidal resonator (132). In that experiment, upon cryogenic cooling of a microtoroid in a helium gas environment, high fre- quency (∼ MHz) periodic features were observed in the signal of an optical pump-probe setup aimed at quantifying the thermal response of the cryogenic system. However, no further investigations were performed to identify and characterize the source of the periodic features. Henceforth I focus on the high frequency mode set of modes in Fig. 6.2c, which exhibits par-

56 − − − 45KKS −− K K K 445S 445KKS 445S Q π 5S

ω

4445KKS −−− 45S 4445S

Figure 6.2: Measurement of superfluid helium mechanical motion. a, Observation of superfluid oscilla- tions as the cryostat cools to base temperature, while the laser frequency is scanned and optical resonance tracked. Oscillations that result from optomechanical parametric instability are observed with as little as 40 nW of optical power, over one hundred times lower than would be expected from a microtoroid me- chanical mode (75). Blue, orange, and green traces are offset vertically and were respectively taken at 3 K, 1 K, and 0.6 K. b, Resonance frequency of the fundamental major mode versus cryostat temperature. Solid line is obtained by modelling the condensation of the helium gas, enabling the film thickness to be calculated as a function of temperature. c, Network analysis at 900 mK showing the low frequency mode (orange shading) that experiences instability and a set of high frequency superfluid modes (blue shading). d, Quality factor of the high frequency superfluid mode indicated by the arrow versus cryostat temperature. e, Spectrum of thermal excitations of the high frequency mode at 330 kHz, with 48 nW of power incident on the microtoroid and at 600 mK. Measurements were performed on a microtoroid with optical decay rate κ/2π = 35.6 MHz for an optical mode at λ = 1560.0 nm.

57 ticularly high quality factor and signal-to-noise. The mechanical quality factor increases dramat- ically with decreasing temperature (Fig. 6.2d), with the dissipation rate reaching a minimum of

Γm/2π = 55 Hz at 600 mK. While lower dissipation rates of 3 mHz have been reported (50) in a third sound resonator of resonance frequency Ωm = 1.2 kHz and quality factor Q= 400000, this was achieved at significantly lower temperature and with a resonator four orders-of-magnitude larger in size. Our smaller resonator increases the mechanical mode confinement and results in higher mode frequencies and a superior Q. f product of 3.3 × 109 Hz. This quantity is relevant for quantum appli- cations since it quantifies the decoherence within a mechanical period, with kT/~ < Q. f indicating less than one thermal phonon enters the oscillator per period. The dissipation mechanisms in superfluid films remain poorly understood, with current theories suggesting that vortex dynamics and surface roughness play important roles (70, 122). The increased mode confinement arising from the micron scale geometry achieved here compared to previous third sound experiments (50, 70), combined with atomically smooth surfaces and greatly enhanced mea- surement precision, provides the prospect to explore dissipation mechanisms in new regimes where surface roughness is negligible, phonon-phonon and phonon-vortex interactions are enhanced via confinement, and driving far out of equilibrium is not required.

6.3.1 Experimental Design: Minimizing Vibration

Our experiment is contained inside a sealed sample chamber which is mounted within an Oxford Instruments closed cycle helium-3 cryocooler with a base temperature of 580 mK. Inside the sample chamber the microtoroid is mounted on Attocube translation stages for linear positioning relative to the tapered optical fibre, which itself is mounted on a custom made glass taper-holder to match the thermal contraction/expansion during thermal cycling. Imaging of the microtoroid and taper is achieved via mounted microscopes directed through a set of windows fixed into the bottom of the cryostat. A stainless steel access tube connected to the sample chamber is used to introduce helium-4 gas into the sample chamber. The amount of helium-4 gas must be carefully chosen to optimize the amount of superfluid helium while minimizing the detrimental heat link arising from helium-4 gas and superfluid “leaking” up the access tube. Vibrations from the continuously running pulse-tube cooler (PTC) are transmitted through the cryostat to the sample chamber, preventing stable positioning of the taper-toroid separation. To cir- cumvent this issue, on-chip stabilization beams are added using photolithography during the fabrica- tion of the microtoroid. Located on either side of the microtoroid the taper may “rest” on the beams, making the relative vibration common-mode and stabilizing the separation to a high precision. This makes it possible to maintain critical coupling indefinitely, even while the PTC is operational. To minimize light scattering out of the taper mode the beams are fabricated with a width of 5 µm and selectively thinned from 2 µm to less than 500 nm. This selective thinning results in no observable loss in taper transmission while contacting the beams. While the addition of stabilization beams allow excellent suppression of relative motion between the taper and microtoroid, the thickness of the su- perfluid is extremely sensitive to vibration and does not benefit from the use of such beams. Vibration of the superfluid manifests as low frequency noise which is parametrically mixed with the high fre-

58 quency mechanical modes to produce broadband noise. Therefore, to make precision measurements on the superfluid mode the PTC is switched off when the cryostat reaches the desired temperature. This severely limits the hold-time of the cryostat, and combined with the heat link from the access tube, results in a reduction in the hold-time of the cryostat from 8hrs to as little as 10min. During the measurement process, while the PTC is switched off, the temperature of the sample chamber could not be stabilized to a high precision. This resulted in significant thermal drifts and a subsequent shift of the mechanical resonance frequency. As such post-processing techniques were implemented to track the mechanical mode throughout the measurement time. Ultra-sensitive readout of the phase fluctuations imprinted by the motion of the superfluid film is achieved via homodyne detection using a fiber interferometer. To lock the relative phase angle between the local oscillator and signal a 200 MHz amplitude modulation is generated before the mi- crotoroid. Mixing down the AC component of the photocurrent at the modulation frequency then provides a phase dependent error signal for the interferometer that is filtered and applied to a piezo- electric fiber stretcher. In this configuration the DC component provides a dispersive error signal for the cavity lock which is enhanced by the local oscillator, enabling stable operation even with nanowatts of optical power in the signal arm of the interferometer.

6.4 Laser Control of Superfluid Film Fluctuations

Many applications in quantum optomechanics require initialization into the mechanical ground state (27, 32, 120, 148, 162), typically achieved via dynamical backaction cooling (32, 120, 162). Here, a series of experiments are reported that investigate dynamical backaction in superfluid film optomechanics, varying the incident optical power, optical mode, and optical detuning. Furthermore, the dynamical backaction on the superfluid mode includes both a radiation pressure and a photothermal component, with the latter thought to be due to the well-known superfluid “fountain effect”. The combination of radiation pressure and photothermal forces provides a mechanism to both cool and heat the super- fluid excitations. Such mechanisms have enabled precise optical control in a wide range of appli- cations including molecular and , quantum optomechanics and quantum information with trapped ions (14, 105). Here, for the first time, optical cooling and heating of the mechanical modes of a superfluid film is presented, dominated primarily by photothermal forces. While radia- tion pressure forces are typically preferred due to their intrinsically lower bulk heating, photothermal backaction it has been shown theoretically to enable ground state cooling without requiring sideband resolution (130).

6.4.1 Identifying Optomechanical Interactions

The ability to resolve the thermal motion of third sound allows thermometry and dynamical back- action analysis to be performed. Here we focus on two high frequency modes denoted mode A

(Ωm = 482 kHz at 530 mK) and mode B (Ωm = 552 kHz at 530 mK). For both modes additional dynamical backaction effects were observed that lead to the conclusion that dispersive radiation pres- sure was not the only optomechanical coupling present in the system. Firstly, spectra of the superfluid

59 150 0 c 10 a 140 d 130 -100 120 -200 110 (Hz) (Hz) π

10 π 100 -300 /2 /2 m

10 b m 90 Γ 80 ∆ω -400 70 -500 60 Normalized area under peak (a.u.) 10 50 -600 10 10 10 10 10 10 Laser power (nW) Laser power (nW) Laser power (nW)

Figure 6.3: On resonance behaviour. Optical power dependence of the mechanical properties of mode A (blue circles) and mode B (orange circles) when the laser is locked to the cavity resonance (i.e zero detuning). (a, b) Measured thermal energy of the superfluid mode versus input laser power. The blue line is a power-law fit to mode A with scaling ∝ P0.6. The orange line is the fit to mode B with scaling 0.69 ∝ P .(c) Mechanical linewidth (Γm/2π) versus optical power. The blue line is a power-law fit given 0.38 0.14 by Γm/2π = 16.0 × P + 19.3 and the orange line is the fit Γm/2π = 49.2 × P + 23.4. (d) Shift in mechanical resonance frequency (∆ωm/2π) versus laser power. motion were acquired over a range of laser powers with zero optical detuning (Fig. 6.3a-d). It was found that the thermal energy of the superfluid mode was dependent on the launched laser power, with a scaling law ∝P0.6 (Fig. 6.3a) and ∝P0.69 (Fig. 6.3b). While laser driving is known to increase the thermal occupation at zero detuning by either bulk heating or quantum backaction, both effects have an integer power law ∝P1, in contrast to the behaviour seen for superfluid modes. To completely elim- inate the possibility of quantum backaction as the source of heating we applied calibrated amplitude noise to the injected optical field and performed thermometry on the superfluid mode. It was found that the applied amplitude noise needed to be many orders of magnitude larger than the quantum noise level, confirming that as expected the superfluid mode is far from the quantum backaction dominated regime. To verify no bulk heating was occurring over the relevant power range, measurements were per- formed on a microtoroid mechanical mode at 1.35 MHz. In this case the thermal energy of the mode, mechanical linewidth, and resonance frequency were all found to be independent of the laser power whilst at zero detuning. Furthermore we confirmed the temperature of the microtoroid mode is in- deed at the temperature of the cryostat by monitoring the power spectral density from 11 K down to 530 mK with fixed optical power. It should also be noted that we have observed behaviour consistent with boil-off of the superfluid layer, namely the disappearance of all superfluid modes along with drastically increased thermal broadening of the optical lineshape; however this occurs at laser powers greater than 5 µW, over an order of magnitude higher than powers used in our experiment. It was also observed that for both modes the mechanical linewidth increases with increasing laser power (Fig. 6.3c) and the mechanical resonance frequency decreases with increasing laser power (Fig. 6.3d). The observation of anomalous heating (inconsistent with bulk heating or quantum backaction) and mechanical damping with increasing laser power indicates that the superfluid modes are coupled to a non-equilibrium bath of some kind (103), and that this bath is driven out of equilibrium by the optical field. While the origin of this non-equilibrium bath is still unknown, one could consider

60 quantized vortices, superfluid phonon-phonon interactions or even an interaction with the normal fluid as possible mechanisms for this behaviour. In addition to the zero detuned backaction shown in Fig. 6.3, there were several detuned phenom- ena that could not be explained by dispersive radiation pressure alone; the measured change in damp- ing rate was too high for the corresponding shift in mechanical frequency, dynamical backaction heat- ing was observed for a red-detuned laser, and the dynamical response was highly dependent on which optical mode the laser was coupled to. One mechanism that can generate strong cooling and heating in the bad cavity regime is photothermal optomechanical coupling, where absorption of photons by the mechanical oscillator thermally induces a displacement in the oscillator position (107, 108, 130, 160). Such coupling might be anticipated here since it is well-known that local heating causes a strong su- perfluid flow towards the source of heat, the so-called “fountain effect” (3). Furthermore, the sign of the photothermal force is known to be dependent on the overlap between the optical and mechanical modes (130, 169), consistent with observations made here.

6.4.2 Theoretical Treatment: Photothermal and Radiation Pressure Optome- chanics

Here I will theoretically explore the situation of generalized optomechanical coupling with radiation pressure and photothermal forces. As highlighted in Ref. (130) the photothermal effect is not well described by an energy conserving Hamiltonian formalism owing to its inherently dissipative nature. This is in stark contrast to the well known dispersive optomechanical Hamiltonian, where energy and momentum are conserved by a reversible interaction between optical and mechanical degrees of freedom. Using an approach similar to that taken in Ref. (130) the dispersive optomechanical Hamiltonian is represented in the Heisenberg picture including coupling to a “bath” of resonators that introduces both fluctuations and dissipation into the system. Similarly the photothermal effect is introduced as a fluctuating force with temporal correlations generated by the thermal response to the random absorption of photons. The quantum Langevin equation for the oscillator position x(t) is:

h 2 i meff x¨(t) + Γm x˙(t) + ωm x(t) = FRP + FPT + Fth (6.13) " Z t # † βA − t−u † = ~g a(t) a(t) + due τt a(u) a(u) (6.14) τt −∞ p + 2ΓmmeffkbTξ(t)

where A is the absorption coefficient given by the ratio of absorbed to circulating optical power and τt is the thermal response time. The dimensionless parameter β quantifies b the relative strength of the photothermal process over the radiation pressure such that FPT = βAFRP. As will be discussed later the absolute value and even the sign of β is strongly dependent on the spatial overlap of the optical and mechanical mode. The intracavity field is dispersively coupled to the mechanical motion so the quantum Langevin equation for the optical annihilation operator a(t) is :

61 h  0 i p a˙(t) = − κ − i ∆ + gx(t) a(t) + 2κiai(t) (6.15)

where the dispersive optomechanical coupling is g and the cavity decay is κ = κi+κ0. It is illustrating to note that photothermal backaction is enabled by the temporal delay of the thermal response in (6.14), and not the cavity induced delay as in standard dispersive optomechanics. As such, for photothermal cooling the condition of being “resolved” considers only the thermal response time, hence relaxing the stringent requirements on minimizing optical decay. To expand the equations of motion into linear and nonlinear components the position and intra- cavity field annihilation operator are expressed as a coherent amplitude with quantum fluctuations i.e. x = x¯ + δx(t) and a(t) = α + δa(t). Here, considering only a linear photothermal and dispersive interaction, the nonlinear terms in (6.14) and (6.15) become:

iga(t)x(t) = ig (αx¯ + δx(t)α + δa(t)x ¯) (6.16)   ~ga†(t)a(t) = ~g |α|2 + 2αδX+(t) (6.17)

+ 1  †  where δX (t) = 2 δa (t) + δa(t) are amplitude fluctuations of the intracavity field.For simplicity the phase of the intracavity field is chosen such that α = α∗. The steady state equation for the optical field is then: √ 2κ α α = i i (6.18) κ − i∆ where the cavity detuning is also modified by the static displacement, ∆ = ∆0 + gx¯. The linearised equations of motion for the fluctuations are then:

" # h i βA  t  2 + − τ + meff δx¨(t) + Γmδx˙(t) + ωmδx(t) = 2~gα δX (t) + H(t)e t ∗ δX (t) (6.19) τt p + 2ΓmmeffkbTξ(t) δa˙(t) = − (κ − i∆) δa(t) + iαδx(t)g (6.20) p + 2κiδai(t)

The photothermal integral in (6.14) has been replaced by a convolution with a Heaviside function H(t) to preserve causality. Transformation into the Fourier domain (i.e. F {δx(t)} = δx) yields:

" " # # βA δx = χ(ω) 2~gαδX+ 1 + + 2Γ m k T (6.21) 1 + iωτ m eff b √ t igαδx + 2κ δα δa = i i (6.22) D(ω) which uses the Fourier identity {H(t)et/τt } = τt , reducing the functional form of the photother- F 1+iτtω

62 mal force to a low pass filter with a corner frequency defined by the characteristic thermalization time. For brevity the mechanical and optical transfer functions have been introduced, χ(ω)−1 =  2 2  meff ωm − ω + iωΓm and D(ω) = κ + i(ω − ∆) respectively. From (6.21) it is clear that radiation pressure and photothermal effects arise from intracavity amplitude fluctuations δX+ which can be expressed as

1   δX+ = δa†(−ω) + δa(ω) (6.23) 2 αδx   − { } = ∗ ∆g0 + O δain (6.24) xzpD(ω)D (−ω) where the dispersive coupling rate has been normalized by the oscillators’ zero point motion (xzp), that is g0 = gxzp, and the fluctuations of the injected optical field have been grouped into a single term

O{δain}. In the case of a coherent optical drive this term quantifies the additional thermomechanical noise generated by optical vacuum fluctuations, commonly known as quantum backaction. Here, the term O{δain} may be discarded since dynamical instabilities and thermal noise greatly exceed quantum backaction. Substituting the resulting intracavity amplitude fluctuations into (6.21) gives

 2 2 " #  −4g0|α| ωmδxmeff∆ βA p δx = χ(ω) ∗ 1 + + 2ΓmmeffkbT (6.25) D(ω)D (−ω) 1 + iωτt 0 p = χ (ω) 2ΓmmeffkbT (6.26)

2 0 where the substitution xzp = ~/2meffωm has been made. The modified mechanical susceptibility χ (ω) includes radiation pressure and photothermal optomechanical coupling and can be written in the form 0 −1 2 2 χ (ω) = meff(ωm + 2ωδωm − ω + iω[Γm + δΓm]). Isolating real and imaginary terms gives the modification to the mechanical resonance frequency δωm and decay rate δΓm as:

(" ! #) A(ω) βA 2κω2τ βA δω = (κ2 + ∆2 − ω2) 1 + − t (6.27) m 2 2 2 2 2ωm 1 + ω τt 1 + ω τt

(" !#) τ βA βA δΓ = −A(ω) (κ2 + ∆2 − ω2) t + 2κ 1 + (6.28) m 2 2 2 2 1 + ω τt 1 + ω τt

2| |2 4g0 α ωm∆ where A(ω) = |D(ω)D∗(−ω)|2 is proportional to the magnitude of the cavity response. Despite extensive experimental work on linear dispersive optomechanics over the last decade the photothermal effect has not received the same level of interest, potentially due to the perceived limi- tations of its inherently dissipative origin. To illustrate the nature of photothermal coupling, consider the limit that it dominates over radiation pressure, namely βA >> 1. In this situation two distinct regimes arise that can be realized depending on the thermalization rate τt. As previously mentioned the thermal response can be represented by a low pass filter with corner frequency 1/τt. If the corner frequency is larger than ωm then the dominant effect is a shift in the mechanical resonance frequency.

63 Conversely if the corner frequency is less than ωm then the resulting delay in the photothermal force leads primarily to a modified mechanical decay rate. These two scenarios are analogous to the pic- ture of “resolved” and “unresolved” dynamics in dispersive optomechanics, with similar requirements existing for sideband cooling to the ground state.

6.4.3 Observing and Characterizing Dynamical Backaction

Before characterizing dispersive dynamical backaction effects the anomalous on-resonance behaviour must first be calibrated. To this end, the measurements reported in Sec. 6.4.1 are used to add a phe- nomenological term to equation (6.27) and (6.28) that account for the power dependent modification of the resonance frequency and mechanical decay rate. Experimentally probing the optomechanical response from a detuned optical field, this section will focus on two mechanical modes which exhibit particularly high quality factors and signal-to-noise, as in Sec. 6.4.1. As previously mentioned, it was found that dynamical backaction on superfluid modes includes both a radiation pressure and photothermal component, with the latter dominating and thought to be due to the well-known superfluid “fountain effect”. The combination of radiation pressure and photothermal forces provides a mechanism to both cool and heat the superfluid excita- tions as shown experimentally in Fig. 6.4a, d. The strength of the photothermal response is determined by the spatial overlap between optical and mechanical modes. Subsequently, mechanical modes of similar frequency may experience vastly different photothermal forces. This is clearly evident in Fig. 6.4a where the shaded mode at 552.5 kHz is shown to have a significantly stronger response over the adjacent mechanical mode. The associated change in damping rate and resonance frequency is shown in Fig. 6.4b-c with good agreement to theory. The photothermal response seen in Fig. 6.4a-c has the same qualitative characteristics as ra- diation pressure, namely broadening of the mechanical mode accompanied by spring softening when red-detuned from the optical cavity. In contrast, the red-detuned response of another mechanical mode at 482 kHz shows linewidth narrowing and spring stiffening, as seen in Fig. 6.4d. This corresponds to a negative photothermal coefficient, with the change in damping rate and resonance frequency also agreeing with photothermal theory, as seen in Fig. 6.4e-f. The photo-induced rigidity ∇FOM due to the combined radiation pressure and photothermal forces may also be estimated, given that the effective mass is approximately equal to the mass of the film. ∇FOM optimally damps the superfluid motion 2 ∇FOM −8 when it equals the intrinsic rigidity, meffωm (108). In our experiments | 2 | ≈ 1 × 10 per photon; meffωm ∇FOM −6 this is compared to | 2 | = 7 × 10 per photon reported for other photothermal optomechanical meffωm systems (160). In typical photothermal optomechanical systems the slow response and independent control of photothermal forces compared to radiation pressure allows the magnitude and time constants of the two effects to be independently determined(107, 160, 169). Here, the photothermal effect is fast due to the exceptionally high thermal conductivity of superfluid helium(159). The characteristic time con- stant τt can be deduced from the functional form of the photothermal response seen in Fig. 6.4b-c and Fig. 6.4e-f and is found to be approximately 600 ns. However, it was not possible to independently isolate the relative strengths of the radiation pressure and photothermal forces, as required to experi-

64

π H

Δ

Δ HH

H ΔΓ

ΔΩ

Δ H Δκ

π H Δ

Δ

HH H ΔΓ

ΔΩ Δ H Δκ

Figure 6.4: Optomechanical heating and cooling of two third sound modes. Detuning the optical field results in cooling (a) or heating (d) of the superfluid excitations. The laser is red-detuned with respect to the cavity for both the 552 kHz mode and the 482 kHz mode. Whether the mode experiences heating or cooling depends on the relative magnitude of the radiation pressure forces and the sign of the photothermal constant (β), which itself depends on the optical mode. (a) Spectra of two closely spaced mechanical modes around 552 kHz with varying optical detuning. The shaded mode experiences photothermal induced broadening (β > 0), with a mechanical linewidth of Γm/2π = 77 Hz (∆/κ = 0), Γm/2π = 334 Hz (∆/κ = −0.25) and Γm/2π = 457 Hz (∆/κ = −0.5). (b-c) Change in mechanical damping rate and mechanical resonance frequency as a function of detuning for the 552 kHz mode. Solid line: theoretical fit to photothermal broadening. (d) Spectra of two closely spaced mechanical modes around 482 kHz with varying optical detuning. The shaded mode experiences photothermal induced narrowing (β < 0), with a mechanical linewidth of Γm/2π = 115 Hz (∆/κ = 0), Γm/2π = 63 Hz (∆/κ = −0.25) and Γm/2π = 36 Hz (∆/κ = −0.5). (e-f) Change in mechanical damping rate and mechanical resonance frequency as a function of detuning for the 482 kHz mode. Solid line: theoretical fit to photothermal narrowing. Traces in a and d are offset for clarity. All detuning measurements were taken with 200 nW of launched optical power.

65 mentally determine g0 and β. In the context of our experiments, this can be intuitively understood by considering the optomechanical response to an arbitrarily large radiation pressure force accompanied by a similarly large photothermal force with opposite sign (as in Fig. 6.4d-f). In such a situation the magnitude of the response only allows the cumulative effect to be determined, rather than the relative strengths of the two forces. However, one may place a bound on the relative strengths based on the likelihood of having a strong photothermal force delicately balanced by a strong radiation pressure force. Taking this argument, and assuming the photothermal force is at least 50% larger, allows an upper bound of g0/2π . 200 Hz to be established.

6.5 Duffing Nonlinearity

Nonlinearities are predicted to allow the generation of highly non-Gaussian states in optomechanical systems (98, 131). In quantum optomechanics the intrinsic nonlinear interaction is typically very weak, with state-of-the-art devices operating with coherent optical fields in the linearized regime, limiting the ability to generate highly nonclassical states. In contrast, superfluid films feature strong phonon-phonon interactions and a nonlinear van der Waals restoring force which scales as film thick- ness to the inverse fourth power. Further, they can have sub-monolayer thickness, comparable with the amplitude of motion (152), such that nonlinearities dominate the mechanical dynamics (70).

6.5.1 Theoretical Treatment

The classical theory for nonlinear oscillators, particularly the Duffing oscillator, has been well studied for many decades (80) with a strong emphasis on numerical and analytic techniques probing the emergence and dynamics of chaos. Here I will consider the simple situation of weak nonlinearity and weak drive in the limit of high mechanical quality factor; both approximations relevant to our experimental system. The general classical equation of motion is given by

3 meff x¨ + meffΓm x˙ + kx + Ux = Fd(t) (6.29)

where Fd is a coherent drive tone, meff is the mass, U quantifies the strength of the nonlinearity, k and

Γm are the restoring force and mechanical dissipation rate respectively. Before applying perturbation techniques to find approximate analytic solutions it is instructive to first consider the fixed points permitted by the equation of motion and their respective stability. The fixed points are found by taking the steady state limit of the undriven system, namelyx ˙ = 0 and Fd = 0, giving the cubic equation

  x k + Ux2 = 0 (6.30)

Clearly x = 0 is a trivial solution that always exists. However, when kU < 0 two additional solu- q k tions exist at displacements x = ± − U . The type and stability of each steady state solution can be

66 determined from the eigenvalues of the Jacobian matrix.      ∂x˙ ∂x˙   0 1  J =  ∂x ∂v  =   (6.31)  ∂v˙ ∂v˙   2  ∂x ∂x −k − 3Ux −Γm √ 1 − ± 2 2 − The trivial solution at x = 0 gives the eigenvalues λ = 2 ( Γm Γ m 4k). For k > 0 the eigenval- ues are complex with a negative real component corresponding to a stable focus i.e. oscillatory decay. For a negative spring constant, when k < 0, the trivial solution has positive real eigenvalues indicating q an unstable node with no periodicity. At the steady state solutions x = ± − k the eigenvalues are √ U 1 − ± 2 2 λ = 2 ( Γm Γ m + 8k) corresponding to stable foci when k < 0 and U > 0 and unstable nodes when k > 0 and U < 0. The various steady state solutions and their respective stability can be easily seen from the potential energy diagrams shown in Fig. 6.5

k>0 U>0 k<0 U>0

15 10

10 0

5 −10

0 0

Potential(a.u.) 10 −5

0 −10

−10 −15

k>0 U<0 k<0 U<0 −2 −1 0 1 2 −2 −1 0 1 2 Displacement (a.u.)

Figure 6.5: Variations of the potential energy landscape of a Duffing oscillator showing stable (o) and unstable (+) fixed points from the combination of quadratic (dotted line) and quartic (dashed line) terms.

Of the 3 distinct types of nonlinear oscillations shown in Fig. 6.5 here I will only consider the case of positive spring constant, k > 0, given the restoring force provided by the van der Waals interaction. Despite the unknown physical origin of the cubic nonlinearity the experimentally observed “spring softening” effect corresponds to a negative nonlinear term, U < 0, which results in a metastable oscillation about x = 0. For the experimentally relevant situation of weak nonlinearity, weak drive and high mechanical quality factor the unstable region may be ignored, considering only the dynamics about the stable fixed point. Standard perturbation theory is typically used to approximate the solution of nonlinear equations of motion. However, for the Duffing oscillator, this technique gives rise to

67 “secular” terms that results in unbounded growth and unphysically divergent solutions (91, 114). Instead the Poincare-Lindstedt method, also know as secular perturbation theory, is used which takes the regular perturbation series and adds Fourier components to cancel, or destructively interfere, the problematic terms. This technique is covered in numerous textbooks, and in the case of the Duffing oscillator, can be shown to result in a modified mechanical resonance frequency that depends on the 0 3U 2 square of the amplitude of motion ω = ω0 + |x(ω)| . The modified resonance frequency can easily 0 8ω0 be substituted to the standard linear mechanical susceptibility to give ( ) 02 Fd(t) F {x¨ + Γm x˙ + ω x} = F (6.32) 0 m ! 3U 2 2 2 D x(ω) (ω0 + |x(ω)| ) − ω + iΓmω = δ (ω − ωd) (6.33) 8ω0 m where the D is the amplitude of the coherent drive at frequency ωd. In the limit of high mechanical quality factor and weak nonlinearity the square of the drive tone and modified resonance frequency 2 ≈ 2 − − 02 ≈ 2 3U 2 can be approximated as ωd ω0 2ω0∆, where ∆ = ω0 ωd, and ω0 ω0 + 4 x(ω) respectively. Applying the Dirac-Delta function and substituting the approximations into the absolute value squared of Eq. (6.33) gives the expression

 !2   2 2  3U 2 2 2 D x(ω)  2ω0∆ + x(ω) + Γ ω  = (6.34)  4 m 0 m

Expanding and rewriting the above equation in terms of the position squared, namely A(ω) = |x(ω)|2, gives the cubic polynomial

!2 3U   D2 A3(ω) + 3ω ∆UA2(ω) + 4ω2∆2 + Γ2 ω2 A(ω) − = 0 (6.35) 4 0 0 m 0 m which is easily numerically solved to obtain real roots that correspond to the squared amplitude of the forced oscillation. Under certain situations, and only in a frequency band near ω0, a bistability occurs resulting in two allowed oscillation amplitudes for a given drive frequency. This gives the well-known discontinuous spectral response of a Duffing oscillator and is experimentally shown in the next.

6.5.2 Experimental Observation

To investigate the nonlinear nature of third sound modes a coherent drive force is applied by amplitude modulating the injected optical field. As seen in Fig. 6.6, the response exhibits a classic Duffing non- linearity due to phonon-phonon interactions (80, 131). In the previous section the Duffing oscillator was characterized by the addition of a cubic term, namely Ux(t)3 where U is the nonlinear coefficient, to the equation for a simple harmonic oscillator (See Eq. 6.29). For sufficient drive strength the po- sition dependent frequency shift gives rise to unique behaviour such as multistability and hysteresis. Using this classical model our experimental observations are accurately described. Bistability is ev- ident at injected power levels as low as 10 nW, four orders-of-magnitude smaller than reported with other nonlinear optomechanical systems (89), and with the drive scattering as few as 37 photons into

68 3D

3 3376

3376

Figure 6.6: Duffing non-linearity. Network analyser traces as the laser is amplitude modulated at a frequency that is swept over the mechanical resonance. Solid lines: fit to data using the standard model of a Duffing oscillator. the intracavity modulation sidebands. This provides a tool to engineer the quantum behaviour of third sound modes, and may enable detection of weak optical fields in an approach similar to transition edge detectors used for photon counting (113). Furthermore, it may be possibly to greatly enhance the nonlinearity by tuning the film thickness since the minimum force required for mechanical bista- q 3 2 3 −1 p 3 −5/2 −1 bility scales as F ∝ ΓmmeffωmU = Γmd U , where d is the film thickness. However, the exact dependence of the dissipation rate and nonlinear coefficient with film thickness and temperature is still largely unknown in the superfluid community.

6.6 Conclusion

In this chapter a superfluid film cavity optomechanical system was proposed and experimentally demonstrated, allowing precise optical readout and control of superfluid helium excitations. This enables real-time observation of the thermal equilibrium excitations of third sound modes, as well as laser heating and cooling. Superfluid thin films provide unique prospects for quantum optomechanics due to the strong optomechanical coupling and low mechanical dissipation; and are well-suited for practical applications such as high-precision force and inertial sensing. Self-assembly of the film via wan der Waals forces offers a unique opportunity to explore novel physical systems such as superfluid phononic circuits, on-chip superfluid interferometers, optomechanics with quantized vortices, and quantum computing with floating electrons. Superfluid film optomechanics may also provide insight into the microscopic behaviour and interactions of superfluids, and improve our understanding of the dynamics of two-dimensional condensates.

69 Chapter 7

Squeezed Light For Enhanced Transduction Sensitivity

This chapter reports quantum enhanced mechanical transduction sensitivity via the application of squeezed light in a micron scale optomechanical system. The phase modulation signal, induced by mechanical fluctuations of a microtoroidal cavity, are measured with a sensitivity −0.72(±0.01) dB below the shot noise level. This is achieved by preparing the optical probe field in a coherent state with phase squeezed vacuum states at sideband frequencies covering a set of mechanical modes around

5 MHz. Since ωm/κ < 1, the experiments presented here are performed in the undercoupled regime, hence minimizing the degradation to squeezing from vacuum fluctuations entering via the intrinsic cavity loss. The figures presented in this chapter are derived from the following reference,

Ulrich B. Hoff, Glen I. Harris, Lars S. Madsen, Hugo Kerdoncuff, Mikael Lassen, Bo M. Nielsen, Warwick P. Bowen, and Ulrik L. Andersen, “Quantum-enhanced micromechanical displacement sen- sitivity,” Optics Letters 38, 1413-1415, (2013)

The detailed theoretical derivation of enhanced transduction sensitivity via squeezed light is not presented here but can be found in the following paper.

Hugo Kerdoncuff, Ulrich B. Hoff, Glen I.Harris, Warwick P. Bowen, and Ulrik L. Andersen, “Squeezing- enhanced measurement sensitivity in a cavity optomechanical system,” To be published in Annalen der Physik (2014)

7.1 Quantum Enhanced Sensing

First considered in the context of gravitational wave detection, the quantum limits to interferomet- ric measurements of mechanical displacements remains an active field of fundamental research; for example, enabling the quantum effects of mesoscopic mechanical systems to be revealed(140, 157). Furthermore, advances in fabrication and measurement techniques have also greatly enhanced appli- cations in cantilever-based sensing of single spins(137), mass(96), and magnetic and electric fields(35, 53). In the absence of classical noise it is well known that any optical readout of displacement has

70 two distinct contributions to the measurement uncertainty(29). The imprecision noise in the form of shot noise of the detected probe light and the quantum back-action noise due to the radiation pressure generated by the stochastic arrival of photons on the mechanical resonator. In principle, coherent states of light allow arbitrarily high displacement sensitivities to be achieved by simply increasing √ the optical power. This is due to the favorable scaling of imprecision noise as ∝ 1/ N, where N is the number of probe photons. Conversely, as the number of probe photons increases the resulting radiation pressure driving, also known as quantum back-action (QBA), eventually becomes the dom- √ inant source of white noise on the mechanical motion. With QBA scaling as ∝ N, the so-called standard quantum limit (SQL) is reached when the two noise sources contribute equally to the total measurement imprecision. The first proposal to surpass the shot-noise limit involves the injection of squeezed light into the dark port of an interferometer(30). To this end, there have been numerous experimental verifica- tions of improved interferometric sensitivity via squeezed light in both Mach-Zehnder (171), Sagnac (47), and gravitational-wave (38) interferometers. Furthermore, squeezed-light-enhanced micropar- ticle tracking has also recently been demonstrated (154). In this chapter, squeezed-light-enhanced mechanical transduction sensitivity is experimentally demonstrated using a tapered fiber coupled mi- crotoroidal resonator at room temperature; hence extending the applicability of quantum-enhanced interferometry to the regime of micromechanical oscillators. This has many practical applications such as improving the minimum achievable sensitivity at a given probe power, thus avoiding the instabilities and damage thresholds that typically limit further increases to the optical power that would otherwise reduce imprecision noise. Furthermore, squeezed light allows modifications to the SQL, typically in the form of shifting the characteristic power level by applying phase or amplitude squeezed light, but also enabling sensitivities beyond the SQL by exploiting quadrature anticorrela- tions when the squeezing phase is set to 45 deg(117).

7.1.1 Squeezed Light in Cavity Optomechanics

Some of the most impressive systems for high-sensitivity displacement detection have been mi- cron scale cavity optomechanical systems such as microtoroidal resonators(142) and photonic crystal cavities(32). In such systems, radiation pressure couples the high quality mechanical and optical degrees of freedom, resulting in the well known Hamiltonian

† † † † Hom = ~ωca a + ~ωmb b − ~g0a a(b + b) (7.1)

† † where a and a (b and b ) are the optical (mechanical) ladder operators and ωc (ωm) is the resonance frequency of the optical (mechanical) cavity. The strength of the interaction is quantified by the vacuum optomechanical coupling rate g = gx = ∂ω /∂xx where x = b† + b is the position and √ 0 zpm 0 zpm xzpm = ~/2meffωm the zero-point fluctuations. This coupling rate represents the optical frequency shift produced by a mechanical displacement equal to the zero point motion.

For the experimentally relevant situation of weak coupling i.e. g0 << {κ, Γm}, driving the mechani- cal degree of freedom leads to a harmonic modulation of the resonance frequency of the optical cavity.

71 This corresponds to a pure phase modulation of the transmitted light field at the mechanical oscillation frequency Ωm, when probing the optical cavity on resonance. To quantify the strength of the signal, a dimensionless parameter can be calculated called the modulation index, which is given by the peak frequency deviation to the modulation frequency ξ = gδx/Ωm (144). With this dimensionless param- eter, the mechanically phase modulation can be represented as a quantum mechanical displacement h √ i operation Dω0±Ωm iξα/ 2 of the ω0 ± Ωm sidebands. This process is graphically shown in Fig. 7.1 which depicts the excitation of weak coherent sideband states mediated by the optomechanical in- teraction. Subsequent phase sensitive homodyne detection of the sideband states reveals information about the mechanical resonance frequency, decay rate and oscillation amplitude. The imprecision noise, given by the spectral variance of the measured phase quadrature, is given by (61)

2 + − + −  V[X2(ω)] ∝ β V[X2 + X2 ] + V[X1 − X1 ] (7.2)

+ − + − where β is the local oscillator (LO) amplitude and X2 and X2 (X1 and X1 ) represent the phase (ampli- tude) quadrature around the ±ωm sideband frequency. Such a measurement process can be formally described as a joint measurement of the upper and lower sideband states at each detection frequency. In the common case of quantum noise limited coherent probing the sidebands are automatically initialized into a vacuum state, resulting in a subsequent measurement imprecision at the “shot noise” level. Alternatively, prior to interaction with the optomechanical system, the sideband states can be initialized into a phase squeezed vacuum state characterized by quantum correlations that are sym- metric around the carrier. Upon the joint measurement described above, these quantum correlations act to reduce the measurement imprecision below the shot noise level, yielding a quantum-enhanced displacement sensitivity.

7.2 Experimental Results

To experimentally realize quantum-enhanced interferometric displacement sensitivity, light from an Nd:YAG laser at λ = 1064 nm is split on a polarizing beam splitter to form the probe and LO fields. To “clean” the spectral and spatial properties of the laser light prior to entering the experimental apparatus, a FabryPerot ring cavity is used as a filtering mode cleaner which is monitored by a tap- off and actively stabilized using the Pound-Drever-Hall locking technique. As can be seen in the experimental schematic in Fig. 7.2 the probe field can follow two alternative beam paths for either coherent state probing (path 1.) or quantum-enhanced phase squeezed probing (path 2.). The squeezed vacuum sideband states are generated in a bow-tie configuration PPKTP-based optical parametric amplifier (OPA)(82). After selecting either path using flip-mount mirrors, the probe field is coupled into a cleaved optical fiber using an antireflection-coated aspheric singlet lens with f = 8.07 mm. A tapered optical fiber, produced by a brushed hydrogen flame technique, is used to couple light into the microtoroid resonator. The transmitted probe field is then coupled to free-space and interfered with the local oscillator followed by balanced homodyne detection. The differential DC component from the homodyne detection is used to actively stabilizing the LO phase via a piezo-actuated mirror and proportional-integral (PI) controller. The differential AC component contains the signal of interest,

72 Figure 7.1: Combined sideband and phase space representation of the optomechanical transduction mech- anism. The sidebands are initially prepared in a composite two mode squeezed state with the individual sidebands being in a thermal state (61). The optomechanical phase modulation scatters photons from the coherent carrier field (green) into the sidebands, displacing the initial quantum states δa± (red). Figure derived from (69) the mechanical motion, and is spectrally analyzed using an electronic spectrum analyzer (ESA). To characterize the squeezed state transmitted through the tapered fiber, the microtoroid is de- coupled and the phase of the local oscillator is varied by applying a saw-tooth signal to the PZT, whilst performing homodyne measurements of the field quadratures. As the LO phase is varied the interrogated optical quadrature cycles between phase and amplitude, revealing typical squeezed and anti-squeezed characteristics, as seen in Fig. 7.3A. By actively stabilizing the LO, linear fits to vac- uum and squeezed noise levels yields −1.20(±0.03) dB of squeezing, corresponding to a squeezing factor r = 0.138(±0.003). This modest squeezing level is limited by large coupling and propagation losses in the tapered fiber amounting to ≥ 30%, arising from the adhesion of dust (54) and/or water to the fiber, causing increased scattering from the tapered region. The homodyne visibility was mea- sured to be 98(±1)% and the quantum efficiency of the photodiodes (Epitaxx ETX-500) is 87(±2)% at 1060 nm. Optical coupling is achieved by piezo positioning of the microtoroid into the evanescent field of the tapered optical fiber. Due to the limited frequency scan range of the Nd:YAG laser, in addition to difficulties maintaining lock of the mode cleaning cavity and the OPA whilst scanning, the microtoroid optical resonance was thermally tuned onto the laser frequency. This was achieved by placing the mi- crotoroid, which ex, on an Peltier element with temperature stabilized feedback control. Based on a simple power transmission measurement the total optical loss was estimated to be κ/2π = 180 MHz

(FWHM) together with a coupling parameter of ηc = 0.025, corresponding to operation in the highly undercoupled regime. Figure 7.3B (green) shows the recorded power spectral densities of the bal- anced homodyne signal, normalized to shot noise, with the LO phase locked for phase quadrature detection whilst using phase squeezed probe states. The spectrum contains transduced signals from

73 Figure 7.2: Schematic diagram of the experimental setup. MC: mode cleaning cavity. PZT: piezoelectric transducer. PBS: polarizing beam splitter. LO: local oscillator field. AL: aspherical lens. OPA: optical parametric amplifier. Inset: SEM micrograph of the microtoroid with major diameter 60 µm and minor diameter 6 µm. Figure derived from (69)

74

--

--/3 π π π -

--/3 - - -/3 Figure 7.3: A Characterization of the transmitted vacuum squeezed state by modulating the local oscillator phase. All traces were recorded in zero span mode at a detection frequency of 4.9 MHz, RBW 100 kHz, and VBW 100 Hz. The detector dark noise was measured to be 25 dB below shot noise and has not been subtracted. B Power spectral density showing mechanical modes of a microtoroidal resonator measured with squeezed light (green), yielding a transduction sensitivity reduced below the SNL (black). All traces are averaged over 10 samples, recorded with RBW 10 kHz and VBW 100 Hz. Figure derived from (69)

three mechanical modes of the microtoroid at frequencies in the range 4 − 5.8 MHz. Critically, the background noise floor is reduced by −0.72(±0.01) dB below the shot noise level; the direct result of using phase squeezed probe states. The improved transduction sensitivity was evaluated by linear fits to the recorded noise levels (dashed green and black lines). All traces were recorded with LO and probe carrier powers of 1.2 mW and 20 µW, respectively. Extending beyond the frequency range shown in Fig. 7.3B, noise squeezing was observed from 3.5 MHz to the full 20 MHz bandwidth of the bow-tie cavity comprising the OPA.

75 7.3 Conclusion

In conclusion, using phase squeezed states of light, a quantum enhanced measurement of the mechan- ical modes of a fiber-coupled microtoroidal resonator is demonstrated. The transduction sensitivity was shown to be reduced by −0.72 dB below the shot-noise-limit. This value is limited primarily by losses in the tapered fiber, permitting significant improvement through further optimization of the tapered fiber production, shielding and nitrogen purging of the taper-toroid region. With these mod- ifications, in addition to increased homodyne detection efficiency by use of high-quantum efficiency diodes, it is experimentally feasible to attain over 3 dB of improvement to the mechanical transduc- tion sensitivity. At this level of squeezing, one immediate application would be to improve the per- formance of feedback cooling schemes, where the ultimate cooling limit is set by the signal-to-noise ratio of the transduction signal (85). It is important to note that the quantum-enhancement technique reported in this chapter not only applies to cavity optomechanics in the undercoupled regime. While in general the best transduction sensitivity is achieved at critical coupling, since our system is “unre- solved” i.e. Ωm/κ < 1, working in the undercoupled regime minimizes the degradation of squeezing at relevant frequencies. However, a “resolved” system with Ωm/κ > 1, such as that in Ref. (142), can be operated in the critically coupling regime whilst maintaining squeezing, yielding an optimal sensitivity enhancement.

76 Chapter 8

Conclusion

In summary, the bulk of this thesis reports the development of experimental techniques in micron scale WGM optomechanical systems, such as opto-electromechanical feedback cooling, feedback stabilization of parametric instability, and quantum enhanced transduction of mechanical motion. In addition to experimental developments, real-time estimation techniques were considered with the aim of enhanced optomechanical sensing. In this context, it was shown that under a broad set of circum- stances, feedback onto a linear optomechanical system provides no advantage to sensing applications over that possible with real-time estimation. Finally, a promising and novel optomechanical system is demonstrated in the form of an evanescently coupled thin film of superfluid helium-4. This sys- tem was shown shown to exhibit a strong Duffing nonlinearity in addition to photothermal, radiation pressure and dissipative dynamical backaction effects. More specifically, Chap. 3 provides a detailed mathematical treatment of a dual probe position measurement of a mechanical oscillator experiencing feedback cooling from one of the probes. This was then experimentally realized in a microtoroidal resonator using opto-electromechanical feed- back, with the predicted squashing effect shown to limit the ability to accurately characterize the oscillator whilst feedback cooling, hence highlighting the importance of the second probe. Chap- ter 4 characterizes the degradation in measurement sensitivity from parametric instability induced by radiation pressure dynamical backaction. To suppress the instability a generalized control tech- nique was designed, and experimentally realized, demonstrating enhanced measurement sensitivity in the region of instability. In the following chapter (Chap. 5), the role of feedback and estimation (i.e post-processing) in optomechanical sensors is evaluated. It was theoretically shown that, even in the presence of non-Gaussian or correlated noise and non-stationary processes, a real-time estima- tion strategy always exists that can mimic any linear feedback protocol; hence invalidating the use of linear feedback onto a linear system to enhance sensing applications. Chapter 7 reports the use of phase squeezed light to enhance the transduction sensitivity of a microtoroidal resonator, extend- ing the applicability of non-classical light to the regime of micromechanical oscillators for the first time. The final chapter (Chap. 6) focuses on surfaces waves of superfluid helium-4, demonstrating precise optical readout of the Brownian motion of a specific mechanical mode, in addition to laser based heating and cooling via dynamical backaction. Furthermore, weak driving of the mechanical mode results in a strong nonlinear response which is shown to be described by the Duffing oscillator.

77 Given the predicted properties of mechanical oscillators based on superfluid films, this system is a very promising candidate to study macroscopic non-classical mechanical states.

8.0.1 Other Work

This thesis reports the most influential results attained during my PhD, however there were other projects that are neither published in a journal nor discussed in any of the main chapters.

• Coherent control: I developed a simple theoretical model that considered “recirculating” the outcoupled light from a microtoroidal resonator for all optical coherent control of the me- chanical motion. The model gave promising preliminary results, motivating the design and construction of an experimental setup that utilizes a fiber ring cavity to recycle the outcoupled optical field. The signal phase delay, or control filter function, was defined by the total length of the ring cavity, hence allowing heating or cooling of the mechanical mode by appropriate choice of delay. While preliminary experimental data suggested weak modifications to the me- chanical susceptibility, the experiments presented in Chap. 5 where deemed higher impact and subsequently took precedence.

• Optimal control: A theoretical protocol for optimal estimation and optimal control was derived in the context of feedback cooling of a micromechanical system. For a purely classical system this is well known within the engineering community as a Kalman filter, however this project had also included quantum effects, such as radiation pressure shot noise, into the model. While an optimal filter was successfully derived, it was decided that the result did not significantly contribute to the well established field of feedback control.

• Fast Feedback Cooling: Based on our understanding of feedback cooling from Chap. 3, a fast, variable gain feedback loop was built to enhance cooling performance. It was experimentally demonstrated that by adapting the gain of a feedback loop during the cool down period, the minimum temperature could be reached twice as fast than that achievable with a stationary gain value.

8.1 Future work

Future work will likely be focused towards developing the superfluid optomechanical system. In this context, the first goal is to improve the experimental design by increasing the sample chamber size (a small chamber limits the taper transmission through bending loss) and integrating a closed cycle liquid helium “pot” onto the 3K stage. This additional pot of helium provides a cooling mechanism via boil-off while the PTC is turned off, maintaining low temperature operation while minimizing vibration. These improvements, alongside others, are expected to increase the hold-time from min- utes at 600mK to hours at 300mK, enabling accurate characterization of dissipation and nonlinearity mechanisms within the superfluid film. Also, efforts will be focused towards developing techniques to

78 independently calibrate the photothermal and radiation pressure effects in the experimentally relevant situation where both processes are faster than the mechanical resonance frequency. Once these improvements have been made, efforts will be focused towards cooling the mechanical motion close to the ground state, either by radiation pressure or photothermal backaction, potentially enabling the observation of quantum backaction and ponderomotive squeezing. Furthermore, utiliz- ing the strong mechanical nonlinearity may enable the generation of non-classical states of light or motion.

79 References

[1] A. Abramovici, W. E. Althouse, R. W. P. Gorodetsky, J. P. Kotthaus, and T. J. Drever, Y. Grsel, S. Kawamura, and Kippenberg. Measuring nanomechani- e. al. Ligo: The laser interfer- cal motion with an imprecision below the ometer gravitational-wave observatory. standard quantum limit. Phys. Rev. A, 256(5055):325–333, 1992. 82(6):061804, 2010.

[2] G. S. Agarwal and S. S. Jha. Theory of [9] G. Anetsberger, R. Riviere, A. Schliesser, optomechanical interactions in superfluid O. Arcizet, and T. J. Kippenberg. he. Phys. Rev. A, 90(2):023812, 2014. Ultralow-dissipation optomechanical resonators on a chip. Nat , [3] J. F. Allen and H. Jones. New phenom- 2(10):627–633, 2008. ena connected with heat flow in helium ii. Nature, 141:243–244, 1938. [10] G. Anetsberger, E. M. Weig, J. P. Kot- thaus, and T. J. Kippenberg. Cavity op- [4] V. R. Almeida, C. A. Barrios, R. R. tomechanics and cooling nanomechani- Panepucci, and M. Lipson. All-optical cal oscillators using microresonator en- control of light on a silicon chip. Nature, hanced evanescent near-field coupling. 431(7012):1081–1084, 2004. 12(910):800–816, 2011. [5] V. Ambegaokar, B. I. Halperin, D. R. Nel- son, and E. D. Siggia. Dynamics of su- [11] O. Arcizet, T. Briant, A. Heidmann, and perfluid films. Phys. Rev. B, 21(5):1806– M. Pinard. Beating quantum limits in an 1826, 1980. optomechanical sensor by cavity detun- ing. Phys Rev A, 73(3):033819, 2006. [6] H. S.-N. Amir and P. Oskar. Proposal for an optomechanical traveling wave [12] O. Arcizet, P. F. Cohadon, T. Briant, phononphoton translator. 13(1):013017, M. Pinard, A. Heidmann, and e. al. High- 2011. sensitivity optical monitoring of a mi- cromechanical resonator with a quantum- [7] G. Anetsberger, O. Arcizet, Q. P. Un- limited optomechanical sensor. Phys Rev terreithmeier, R. Riviere, A. Schliesser, Lett, 97(13):133601, 2006. E. M. Weig, J. P. Kotthaus, and T. J. Kip- penberg. Near-field cavity optomechan- [13] A. Ashkin. Applications of laser radiation ics with nanomechanical oscillators. Nat. pressure. 210(4474):1081–1088, 1980. Phys., 5(12):909–914, 2009. [14] M. Aspelmeyer, T. Kippenberg, and [8] G. Anetsberger, E. Gavartin, O. Arcizet, F. Marquardt. Cavity optomechanics. Q. P. Unterreithmeier, E. M. Weig, M. L. arXiv e-prints, page 1303.0733v1, 2013.

80 [15] K. R. Atkins. Third and fourth sound in bility in fabry-perot interferometer. Phys liquid helium ii. Phys. Rev., 113(4):962– Lett A, 287(5-6):331–338, 2001. 965, 1959. [25] N. Brahms, T. Botter, S. Schreppler, [16] H. Bachor and T. Ralph. A Guide to D. W. C. Brooks, and D. M. Stamper- Experiments in Quantum Optics. Wiley, Kurn. Optical detection of the quantiza- 2004. tion of collective atomic motion. Phys. Rev. Lett., 108(13):133601, 2012. [17] M. Bagheri, M. Poot, M. Li, W. P. H. Per- [26] S. T. Bramwell and B. Keimer. Neutron nice, and H. X. Tang. Dynamic manipula- scattering from quantum condensed mat- tion of nanomechanical resonators in the ter. Nat. Mat., 13(8):763–767, 2014. high-amplitude regime and non-volatile mechanical memory operation. Nat Nan- [27] D. W. C. Brooks, T. Botter, S. Schrep- otechnol, 6(11):726–732, 2011. pler, T. P. Purdy, N. Brahms, and D. M. Stamper-Kurn. Non-classical light gener- [18] P. F. Barker. Doppler cooling a micro- ated by quantum-noise-driven cavity op- sphere. 105(7):073002, 2010. tomechanics. Nature, 488(7412):476– [19] D. Bergman. Hydrodynamics and third 480, 2012. Phys. Rev. sound in thin he ii films. , [28] T. Carmon, L. Yang, and K. J. Vahala. 188(1):370, 1969. Dynamical thermal behavior and thermal Opt Ex- [20] G. Binnig, C. F. Quate, and C. Gerber. self-stability of microcavities. press Atomic force microscope. Phys Rev Lett, , 12(20):4742–4750, 2004. 56(9):930–933, 1986. [29] C. M. Caves. Quantum-mechanical radiation-pressure fluctuations in an inter- [21] D. G. Blair, E. N. Ivanov, M. E. To- ferometer. Phys. Rev. Lett., 45(2):75–79, bar, P. J. Turner, F. Vankann, and I. S. 1980. Heng. High-sensitivity gravitational- wave antenna with parametric transducer [30] C. M. Caves. Quantum-mechanical noise readout. Phys Rev Lett, 74(11):1908– in an interferometer. Phys Rev D, 1911, 1995. 23(8):1693–1708, 1981.

[22] S. Bose, K. Jacobs, and P. L. Knight. [31] C. M. Caves, K. S. Thorne, R. W. P. Dr- Scheme to probe the decoherence of a ever, V. D. Sandberg, and M. Zimmer- macroscopic object. 59(5):3204–3210, mann. On the measurement of a weak 1999. classical force coupled to a quantum- mechanical oscillator .1. issues of prin- [23] V. Braginski, F. Khalili, and K. Thorne. ciple. Rev. Mod. Phys., 52(2):341–392, Quantum Measurement. Cambridge Uni- 1980. versity Press, 1995. [32] J. Chan, T. P. M. Alegre, A. H. Safavi- [24] V. B. Braginsky, S. E. Strigin, and S. P. Naeini, J. T. Hill, A. Krause, S. Grob- Vyatchanin. Parametric oscillatory insta- lacher, M. Aspelmeyer, and O. Painter.

81 Laser cooling of a nanomechanical oscil- [40] T. Corbitt, C. Wipf, T. Bodiya, D. Ot- lator into its quantum ground state. Na- taway, D. Sigg, N. Smith, S. Whit- ture, 478(7367):89–92, 2011. comb, and N. Mavalvala. Optical di- lution and feedback cooling of a gram- [33] D. E. Chang, C. A. Regal, S. B. Papp, scale oscillator to 6.9 mk. Phys Rev Lett, D. J. Wilson, J. Ye, O. Painter, H. J. Kim- 99(16):160801, 2007. ble, and P. Zoller. Cavity opto-mechanics using an optically levitated nanosphere. [41] B. D. Cuthbertson, M. E. Tobar, E. N. 107(3):1005–1010, 2010. Ivanov, and D. G. Blair. Parametric back- action effects in a high-q cyrogenic sap- [34] J. H. Chow, M. A. Taylor, T. T. Y. phire transducer. 67(7):2435–2442, 1996. Lam, J. Knittel, J. D. Sawtell-Rickson, D. A. Shaddock, M. B. Gray, D. E. Mc- [42] P. Del’Haye, T. Herr, E. Gavartin, M. L. Clelland, and W. P. Bowen. Critical Gorodetsky, R. Holzwarth, and T. J. coupling control of a microresonator by Kippenberg. Octave spanning tunable laser amplitude modulation. Opt. Exp., frequency comb from a microresonator. 20(11):12622–12630, 2012. Phys. Rev. Lett., 107(6):063901, 2011.

[43] L. A. DeLorenzo and K. Schwab. Super- [35] A. N. Cleland and M. L. Roukes. A fluid optomechanics: Coupling of a su- nanometre-scale mechanical electrome- perfluid to a superconducting condensate. ter. Nature, 392(6672):160–162, 1998. arXiv e-prints, 2013. [36] P. F. Cohadon, A. Heidmann, and [44] A. Dorsel, J. D. Mccullen, P. Meystre, M. Pinard. Cooling of a mirror by E. Vignes, and H. Walther. Optical bista- radiation pressure. Phys. Rev. Lett., bility and mirror confinement induced 83(16):3174–3177, 1999. by radiation pressure. Phys Rev Lett, 51(17):1550–1553, 1983. [37] L. Collaboration. Observation of a kilogram-scale oscillator near its quantum [45] G. Drake. Springer Handbook of Atomic, ground state. New J. Phys., 11:073032, Molecular, and Optical Physics. Springer, 2009. 2006.

[38] T. L. S. Collaboration. A gravitational [46] A. Duh, A. Suhel, B. D. Hauer, R. Saeedi, wave observatory operating beyond the P. H. Kim, T. S. Biswas, and J. P. Davis. quantum shot-noise limit. Nat. Phys., Microfluidic and nanofluidic cavities for 7(12):962–965, 2011. quantum fluids experiments. J. Low Temp. Phys., 168(1-2):31–39, 2012. [39] T. Corbitt, D. Ottaway, E. Innerhofer, J. Pelc, and N. Mavalvala. Measure- [47] T. Eberle, S. Steinlechner, J. Bauchrowitz, ment of radiation-pressure-induced op- V. Hndchen, H. Vahlbruch, M. Mehmet, tomechanical dynamics in a suspended H. Mller-Ebhardt, and R. Schnabel. fabry-perot cavity. 74(2):021802, 2006. Quantum enhancement of the zero-area

82 sagnac interferometer topology for gravi- degradation in tapered fibers. Opt. tational wave detection. 104(25):251102, Express, 19(9):8596–8601, 2011. 2010. [55] C. Gardiner and P. Zoller. Quan- [48] M. Eichenfield, R. Camacho, J. Chan, tum Noise: A Handbook of Markovian K. J. Vahala, and O. Painter. A and Non-Markovian Quantum Stochastic picogram- and nanometre-scale photonic- Methods with Applications to Quantum crystal optomechanical cavity. Nature, Optics. Springer, 2004. 459(7246):550–U79, 2009. [56] E. Gavartin, P. Verlot, and T. J. Kippen- [49] A. Eichler, J. Moser, J. Chaste, M. Zdro- berg. A hybrid on-chip optomechanical jek, I. Wilson-Rae, and A. Bachtold. transducer for ultrasensitive force mea- Nonlinear damping in mechanical res- surements. Nat. Nano., 7(8):509–514, onators made from carbon nanotubes and 2012. graphene. Nat Nanotechnol, 6(6):339– 342, 2011. [57] E. Gavartin, P. Verlot, and T. J. Kippen- berg. Stabilization of a linear nanome- [50] F. M. Ellis and H. Luo. Observation of the chanical oscillator to its thermodynamic persistent-current splitting of a 3rd-sound limit. 4, 2013. resonator. Phys. Rev. B, 39(4):2703–2706, 1989. [58] S. Gigan, H. R. Bohm, M. Paternostro, F. Blaser, G. Langer, J. B. Hertzberg, [51] N. E. Flowers-Jacobs, A. D. Kashkanova, K. C. Schwab, D. Bauerle, M. As- A. B. Shkarin, S. W. Hoch, C. Deutsch, pelmeyer, and A. Zeilinger. Self-cooling D. Reichel, and J. G. E. Harris. Fiber- of a micromirror by radiation pressure. cavity optomechanics with superfluid he- Nature, 444(7115):67–70, 2006. lium. In APS March Meeting. APS.

[52] S. Forstner, S. Prams, J. Knittel, E. D. [59] V. Giovannetti, S. Lloyd, and L. Mac- van Ooijen, J. D. Swaim, G. I. Har- cone. Quantum-enhanced measurements: ris, A. Szorkovszky, W. P. Bowen, and Beating the standard quantum limit. H. Rubinsztein-Dunlop. Cavity optome- 306(5700):1330–1336, 2004. chanical magnetometer. Phys. Rev. Lett., [60] V. Giovannetti and D. Vitali. Phase-noise 108(12):120801, 2012. measurement in a cavity with a mov- [53] S. Forstner, E. Sheridan, J. Knit- able mirror undergoing quantum brown- tel, C. L. Humphreys, G. A. Brawley, ian motion. 63(2):023812, 2001. H. Rubinsztein-Dunlop, and W. P. Bowen. [61] O. Glockl. Quantenverschrnkung Ultrasensitive optomechanical magne- und Interferometrie fr die Quanten- tometry. Adv. Mat., 26(36):6348–6353, informationsverarbeitung. Lehrstuhl 2014. fr Mikrocharakterisierung, Friedrich- [54] M. Fujiwara, K. Toubaru, and Alexander-Univ. Erlangen-Nrnberg, S. Takeuchi. Optical transmittance 2005.

83 [62] L. F. Gomez, K. R. Ferguson, J. P. ical displacement sensitivity. Opt Lett, Cryan, C. Bacellar, R. M. P. Tanyag, 38(9):1413–1415, 2013. and e. al. Shapes and vorticities of su- [70] J. A. Hoffmann, K. Penanen, J. C. Davis, perfluid helium nanodroplets. Science, and R. E. Packard. Measurements of 345(6199):906–909, 2014. attenuation of third sound: Evidence of [63] S. Groblacher, K. Hammerer, M. R. Van- trapped vorticity in thick films of super- ner, and M. Aspelmeyer. Observation of fluid he-4. J. Low Temp. Phys., 135(3- strong coupling between a micromechan- 4):177–202, 2004. ical resonator and an optical cavity field. [71] M. Hossein-Zadeh and K. J. Vahala. Ob- Nature, 460(7256):724–727, 2009. servation of optical spring effect in a [64] I. S. Grudinin, H. Lee, O. Painter, and microtoroidal optomechanical resonator. K. J. Vahala. Phonon laser action in a tun- Opt Lett, 32(12):1611–1613, 2007. able two-level system. Phys. Rev. Lett., [72] O. L. R. Jacobs. Introduction to control 104(8):083901, 2010. theory. Clarendon Press, Oxford, 1974. [65] G. I. Harris, U. L. Andersen, J. Knittel, [73] A. Jerri. Introduction to Integral Equa- and W. P. Bowen. Feedback-enhanced tions with Applications. Wiley, 1999. sensitivity in optomechanics: Surpassing the parametric instability barrier. Phys. [74] L. Ju, D. G. Blair, C. Zhao, S. Gras, Rev. A, 85(6):061802, 2012. Z. Zhang, P. Barriga, H. Miao, Y. Fan, and L. Merrill. Strategies for the control of [66] G. I. Harris, D. L. McAuslan, T. M. Stace, parametric instability in advanced gravita- A. C. Doherty, and W. P. Bowen. Mini- tional wave detectors. Class. and Quant. mum requirements for feedback enhanced Grav., 26(1):015002, 2009. force sensing. 111(10):103603, 2013. [75] T. J. Kippenberg, H. Rokhsari, T. Car- [67] J. G. E. Harris, J. D. Thompson, B. M. mon, A. Scherer, and K. J. Vahala. Analy- Zwickl, A. M. Jayich, F. Marquardt, sis of radiation-pressure induced mechan- and S. M. Girvin. Strong dispersive ical oscillation of an optical microcavity. coupling of a high-finesse cavity to a Phys. Rev. Lett., 95(3):033901, 2005. micromechanical membrane. Nature, 452(7183):72–U5, 2008. [76] T. J. Kippenberg and K. J. Vahala. Cavity opto-mechanics. Opt. Exp., [68] A. Heidmann, Y. Hadjar, and M. Pinard. 15(25):17172–17205, 2007. Quantum nondemolition measurement by optomechanical coupling. 64(2):173– [77] D. Kleckner and D. Bouwmeester. Sub- 180, 1997. kelvin optical cooling of a micromechan- ical resonator. Nature, 444(7115):75–78, [69] U. B. Hoff, G. I. Harris, L. S. Mad- 2006. sen, H. Kerdoncuff, M. Lassen, B. M. Nielsen, W. P. Bowen, and U. L. Ander- [78] R. G. Knobel and A. N. Cleland. sen. Quantum-enhanced micromechan- Nanometre-scale displacement sensing

84 using a single electron transistor. Nature, [87] A. J. Leggett. Quantum Liquids: Bose 424(6946):291–293, 2003. condensation and Cooper pairing in condensed-matter systems. 2006. [79] A. G. Krause, M. Winger, T. D. Blasius, Q. Lin, and O. Painter. A high-resolution [88] M. J. Levin. Power spectrum parameter- microchip optomechanical accelerometer. estimation. IEEE Trans. Inf. Theory, Nat. Phot., 6(11):768–772, 2012. 11(1):100–107, 1965.

[80] L. D. Landau and E. M. Lifshitz. Mechan- [89] M. Li, W. H. P. Pernice, C. Xiong, ics. Pergamon Press, 1960. T. Baehr-Jones, M. Hochberg, and H. X. Tang. Harnessing optical forces in [81] L. D. Landau and E. M. Lifshitz. Fluid integrated photonic circuits. Nature, Mechanics. 1987. 456(7221):480, 2008.

[82] M. Lassen, M. Sabuncu, A. Huck, [90] M. Li, H. X. Tang, and M. L. Roukes. J. Niset, G. Leuchs, N. J. Cerf, and Ultra-sensitive nems-based cantilevers for U. L. Andersen. Quantum optical co- sensing, scanned probe and very high- herence can survive photon losses using frequency applications. Nat. Nanotech., a continuous-variable quantum erasure- 2(2):114–120, 2007. correcting code. Nat. Phot., 4(10):700– [91] R. Lifshitz and M. C. Cross. Nonlinear 705, 2010. Dynamics of Nanomechanical and Mi- [83] H. Lck and G. E. O. T. the. The geo600 cromechanical Resonators. Wiley-VCH project. 14(6):1471, 1997. Verlag GmbH and Co. KGaA, 2009.

[92] R. Lifshitz, I. Katz, A. Retzker, and [84] K. K. Leang and S. Devasia. Feedback- R. Straub. Classical to quantum transition linearized inverse feedforward for creep, of a driven nonlinear nanomechanical res- hysteresis, and vibration compensation in onator. New J. Phys., 10:125023, 2008. afm piezoactuators. IEEE Trans. Control Syst. Technol., 15(5):927–935, 2007. [93] F. Liu, S. Alaie, Z. C. Leseman, and M. Hossein-Zadeh. Sub-pg mass sensing [85] K. H. Lee, T. G. McRae, G. I. Harris, and measurement with an optomechani- J. Knittel, and W. P. Bowen. Cooling and cal oscillator. Opt. Exp., 21(17):19555– control of a cavity optoelectromechanical 19567, 2013. system. Phys. Rev. Lett., 104(12):123604, 2010. [94] Y. Liu, H. Miao, V. Aksyuk, and K. Srinivasan. Wide cantilever stiffness [86] S. I. Lee, S. W. Howell, A. Raman, and range cavity optomechanical sensors for R. Reifenberger. Nonlinear dynamics of atomic force microscopy. Opt. Exp., microcantilevers in tapping mode atomic 20(16):18268–18280, 2012. force microscopy: A comparison between theory and experiment. Phys. Rev. B, [95] M. Ludwig, B. Kubala, and F. Mar- 66(11):115409, 2002. quardt. The optomechanical instability

85 in the quantum regime. New J. Phys., bandgap shield. arXiv e-prints 1403.3703, 10:095013, 2008. 2014.

[96] H. J. Mamin and D. Rugar. Sub- [104] J. Mertz, O. Marti, and J. Mlynek. attonewton force detection at mil- Regulation of a microcantilever response likelvin temperatures. Appl. Phys. Lett., by force feedback. Appl. Phys. Lett., 79(20):3358–3360, 2001. 62(19):2344–2346, 1993.

[97] S. Mancini, D. Vitali, and P. Tombesi. [105] H. Metcalf and P. van der Straten. Laser Scheme for teleportation of quantum Cooling and Trapping. Springer New states onto a mechanical resonator. Phys. York, 1999. Rev. Lett., 90(13):137901, 2003. [106] M. Metcalfe. Applications of cav- [98] W. Marshall, C. Simon, R. Penrose, and ity optomechanics. App. Phys. Rev., D. Bouwmeester. Towards quantum su- 1(3):031105, 2014. perpositions of a mirror. Phys. Rev. Lett., 91(13):130401, 2003. [107] C. Metzger, I. Favero, A. Ortlieb, and K. Karrai. Optical self cooling of a de- [99] J. C. Maxwell. A treatise on electric- formable fabry-perot cavity in the clas- ity and magnetism, volume 1. Clarendon sical limit. Phys. Rev. B, 78(3):035309, press, 1881. 2008.

[100] T. G. McRae and W. P. Bowen. Near [108] C. H. Metzger and K. Karrai. Cav- threshold all-optical backaction amplifier. ity cooling of a microlever. Nature, Appl. Phys. Lett., 100(20):201101, 2012. 432(7020):1002–1005, 2004.

[101] T. G. McRae, K. H. Lee, G. I. Har- [109] P. Meystre. A short walk through quan- ris, J. Knittel, and W. P. Bowen. Cav- tum optomechanics. Annalen Der Physik, ity optoelectromechanical system com- 525(3):215–233, 2013. bining strong electrical actuation with ul- trasensitive transduction. Phys. Rev. A, [110] H. X. Miao, K. Srinivasan, and 82(2):023825, 2010. V. Aksyuk. A microelectromechani- cally controlled cavity optomechanical [102] T. G. McRae, K. H. Lee, M. McGovern, sensing system. New J. Phys., 14:075015, D. Gwyther, and W. P. Bowen. Thermo- 2012. optic locking of a semiconductor laser to a microcavity resonance. Opt. Exp., [111] G. Milburn and M. Woolley. An intro- 17(24):21977–21985, 2009. duction to quantum optomechanics. Acta. Phys. Slov., 61(5):483–601, 2011. [103] S. Meenehan, J. Cohen, S. Grblacher, J. Hill, A. Safavi-Naeini, M. Aspelmeyer, [112] M. Montinaro, A. Mehlin, H. S. and O. Painter. Thermalization proper- Solanki, P. Peddibhotla, S. Mack, ties at mk temperatures of a nanoscale D. D. Awschalom, and M. Poggio. Feed- optomechanical resonator with acoustic back cooling of cantilever motion using a

86 quantum point contact transducer. Appl. [122] K. Penanen and R. E. Packard. A model Phys. Lett., 101(13):133104, 2012. for third sound attenuation in thick he-4 films. J. Low Temp. Phys., 128(1-2):25– [113] C. M. Natarajan, M. G. Tanner, and R. H. 35, 2002. Hadfield. Superconducting nanowire single-photon detectors: physics and ap- [123] I. Pikovski, M. R. Vanner, M. As- plications. S. Sci. Tech., 25(6):063001, pelmeyer, M. S. Kim, and C. Brukner. 2012. Probing planck-scale physics with quan- tum optics. 8(5):393–397, 2012. [114] A. H. Nayfeh and D. T. Mook. Nonlinear Oscillations. 1979. [124] M. Pinard, P. F. Cohadon, T. Briant, and A. Heidmann. Full mechanical character- [115] E. F. Nichols and G. F. Hull. The pressure ization of a cold damped mirror. Phys. due to radiation. 17(1):26–50, 1903. Rev. A, 63(1):013808, 2000.

[116] A. D. O’Connell, M. Hofheinz, M. Ans- [125] P. M. Platzman and M. I. Dykman. Quan- mann, R. C. Bialczak, M. Lenander, and tum computing with electrons floating on e. al. Quantum ground state and single- liquid helium. Science, 284(5422):1967– phonon control of a mechanical resonator. 1969, 1999. Nature, 464(7289):697–703, 2010. [126] M. Poggio, C. L. Degen, H. J. Mamin, [117] A. F. Pace, M. J. Collett, and D. F. Walls. and D. Rugar. Feedback cooling of a can- Quantum limits in interferometric detec- tilever’s fundamental mode below 5 mk. tion of gravitational-radiation. Phys. Rev. Phys. Rev. Lett., 99(1):017201, 2007. A, 47(4):3173–3189, 1993. [127] A. Polianin and A. Manzhirov. Handbook [118] R. E. Packard. The role of the josephson- of Integral Equations. Taylor and Francis anderson equation in superfluid helium. Group, 1998. Rev. Mod. Phys., 70(2):641–651, 1998. [128] T. P. Purdy, D. W. C. Brooks, T. Bot- [119] T. A. Palomaki, J. W. Harlow, J. D. Teufel, ter, N. Brahms, Z. Y. Ma, and D. M. R. W. Simmonds, and K. W. Lehnert. Co- Stamper-Kurn. Tunable cavity optome- herent state transfer between itinerant mi- chanics with ultracold atoms. Phys Rev crowave fields and a mechanical oscilla- Lett, 105(13):133602, 2010. tor. Nature, 495(7440):210–214, 2013. [129] T. P. Purdy, R. W. Peterson, and C. A. [120] T. A. Palomaki, J. D. Teufel, R. W. Regal. Observation of radiation pressure Simmonds, and K. W. Lehnert. En- shot noise on a macroscopic object. Sci- tangling mechanical motion with mi- ence, 339(6121):801–804, 2013. crowave fields. Science, 342(6159):710– [130] J. Restrepo, J. Gabelli, C. Ciuti, and 713, 2013. I. Favero. Classical and quantum theory [121] J. W. Patchell and O. L. R. Jacobs. Sep- of photothermal cavity cooling of a me- arability, neutrality and certainty equiva- chanical oscillator. C. R. Phys., 12(9- lence. Int. J. Control, 13(2):337, 1971. 10):860–870, 2011.

87 [131] S. Rips, M. Kiffner, I. Wilson-Rae, and [138] D. Rugar and P. Grutter. Mechanical para- M. J. Hartmann. Steady-state negative metric amplification and thermomechan- wigner functions of nonlinear nanome- ical noise squeezing. Phys. Rev. Lett., chanical oscillators. New J. Phys., 67(6):699–702, 1991. 14:023042, 2012. [139] A. H. Safavi-Naeini, T. P. M. Alegre, [132] R. Riviere, O. Arcizet, A. Schliesser, and J. Chan, M. Eichenfield, M. Winger, T. J. Kippenberg. Evanescent straight Q. Lin, J. T. Hill, D. E. Chang, and tapered-fiber coupling of ultra-high q op- O. Painter. Electromagnetically induced tomechanical micro-resonators in a low- transparency and slow light with optome- vibration helium-4 exchange-gas cryostat. chanics. Nature, 472(7341):69–73, 2011. Rev. Sci. Instrum., 84(4):043108, 2013. [140] A. H. Safavi-Naeini, J. Chan, J. T. Hill, [133] R. Riviere, S. Deleglise, S. Weis, T. P. M. Alegre, A. Krause, and O. Painter. E. Gavartin, O. Arcizet, A. Schliesser, and Observation of quantum motion of a T. J. Kippenberg. Optomechanical side- nanomechanical resonator. Phys. Rev. band cooling of a micromechanical oscil- Lett., 108(3):033602, 2012. lator close to the quantum ground state. Phys. Rev. A, 83(6):063835, 2011. [141] A. Schliesser, G. Anetsberger, R. Riviere, O. Arcizet, and T. J. Kippenberg. High- [134] T. Rocheleau, T. Ndukum, C. Macklin, sensitivity monitoring of micromechan- J. B. Hertzberg, A. A. Clerk, and K. C. ical vibration using optical whispering Schwab. Preparation and detection of gallery mode resonators. New J. Phys., a mechanical resonator near the ground 10:095015, 2008. state of motion. Nature, 463(7277):72– 75, 2010. [142] A. Schliesser, O. Arcizet, R. Riviere, [135] H. Rokhsari, T. J. Kippenberg, T. Carmon, G. Anetsberger, and T. J. Kippenberg. and K. J. Vahala. Radiation-pressure- Resolved-sideband cooling and position driven micro-mechanical oscillator. Opt. measurement of a micromechanical oscil- Exp., 13(14):5293–5301, 2005. lator close to the heisenberg uncertainty limit. Nat. Phys., 5(7):509–514, 2009. [136] H. Rokhsari, T. J. Kippenberg, T. Car- mon, and K. J. Vahala. Theoretical and [143] A. Schliesser and T. J. Kippenberg. experimental study of radiation pressure- Cavity optomechanics with whispering- induced mechanical oscillations (para- gallery mode optical micro-resonators. metric instability) in optical microcavi- Adv Atom Mol Opt Phy, 58:207–323, ties. IEEE J. Sel. Top. Quant., 12(1):96– 2010. 107, 2006. [144] A. Schliesser and T. J. Kippenberg. [137] D. Rugar, R. Budakian, H. J. Mamin, and Chapter 5 - Cavity Optomechanics with B. W. Chui. Single spin detection by mag- Whispering-Gallery Mode Optical Micro- netic resonance force microscopy. Nature, Resonators, volume Volume 58, pages 430(6997):329–332, 2004. 207–323. Academic Press, 2010.

88 [145] A. Schliesser, R. Riviere, G. Anets- with bose-einstein condensates. Nature, berger, O. Arcizet, and T. J. Kippenberg. 409(6816):63–66, 2001. Resolved-sideband cooling of a microme- chanical oscillator. Nat. Phys., 4(5):415– [154] M. A. Taylor, J. Janousek, V. Daria, 419, 2008. J. Knittel, B. Hage, A. BachorHans, and W. P. Bowen. Biological measurement be- [146] R. J. Schoelkopf and S. M. Girvin. yond the quantum limit. 7(3):229–233, Wiring up quantum systems. Nature, 2013. 451(7179):664–669, 2008. [155] M. A. Taylor, A. Szorkovszky, J. Knit- [147] J. H. Scholtz, E. O. McLean, and I. Rud- tel, K. H. Lee, T. G. McRae, and W. P. nick. Third sound and the healing length Bowen. Cavity optoelectromechanical of he ii in films as thin as 2.1 atomic regenerative amplification. Opt. Exp., layers. Phys. Rev. Lett., 32(4):147–151, 20(12):12742–12751, 2012. 1974. [156] J. D. Teufel, T. Donner, M. A. [148] S. Schreppler, N. Spethmann, N. Brahms, Castellanos-Beltran, J. W. Harlow, T. Botter, M. Barrios, and D. M. Stamper- and K. W. Lehnert. Nanomechanical Kurn. Optically measuring force near motion measured with an imprecision the standard quantum limit. Science, below that at the standard quantum limit. 344(6191):1486–1489, 2014. Nat Nanotechnol, 4(12):820–823, 2009. [149] K. C. Schwab, J. B. Hertzberg, T. Roche- leau, T. Ndukum, M. Savva, and A. A. [157] J. D. Teufel, T. Donner, D. L. Li, J. W. Clerk. Back-action-evading measure- Harlow, M. S. Allman, K. Cicak, A. J. ments of nanomechanical motion. Nat. Sirois, J. D. Whittaker, K. W. Lehnert, and Phys., 6(3):213–217, 2010. R. W. Simmonds. Sideband cooling of micromechanical motion to the quantum [150] K. C. Schwab and M. L. Roukes. ground state. Nature, 475(7356):359– Putting mechanics into quantum mechan- 363, 2011. ics. Phys. Today, 58(7):36–42, 2005. [158] J. D. Teufel, D. Li, M. S. Allman, [151] B. S. Sheard, M. B. Gray, C. M. Mow- K. Cicak, A. J. Sirois, J. D. Whittaker, Lowry, D. E. McClelland, and S. E. Whit- and R. W. Simmonds. Circuit cavity comb. Observation and characteriza- electromechanics in the strong-coupling tion of an optical spring. Phys. Rev. A, regime. 471(7337):204–208, 2011. 69(5):051801(R), 2004.

[152] P. J. Shirron and J. M. Mochel. Atomi- [159] D. Tilley and J. Tilley. Superfluidity and cally thin superfluid-helium films on solid Superconductivity. CRC Press, Bristol, hydrogen. Phys. Rev. Lett., 67(9):1118– 1990. 1121, 1991. [160] K. Usami, A. Naesby, T. Bagci, B. M. [153] A. Sorensen, L. M. Duan, J. I. Cirac, Nielsen, J. Liu, S. Stobbe, P. Lodahl, and P. Zoller. Many-particle entanglement and E. S. Polzik. Optical cavity cooling

89 of mechanical modes of a semiconduc- [169] D. Woolf, P. C. Hui, E. Iwase, M. Khan, tor nanomembrane. Nat. Phys., 8(2):168– A. W. Rodriguez, P. Deotare, I. Bulu, 172, 2012. S. G. Johnson, F. Capasso, and M. Lon- car. Optomechanical and photothermal in- [161] H. Van Trees. Detection Estimation and teractions in suspended photonic crystal Modulation Theory, Radar-Sonar Sig- membranes. Opt. Exp., 21(6):7258–7275, nal Processing and Gaussian Signals in 2013. Noise. John Wiley and Sons, 1971. [170] M. J. Woolley, G. J. Milburn, and C. M. [162] E. Verhagen, S. Deleglise, S. Weis, Caves. Nonlinear quantum metrology us- A. Schliesser, and T. J. Kippenberg. ing coupled nanomechanical resonators. Quantum-coherent coupling of a mechan- New J. Phys., 10:125018, 2008. ical oscillator to an optical cavity mode. Nature, 482(7383):63–67, 2012. [171] M. Xiao, L. A. Wu, and H. J. Kimble. Precision-measurement beyond the shot- [163] A. Vinante, M. Bonaldi, F. Marin, and J. P. noise limit. Phys. Rev. Lett., 59(3):278– Zendri. Dissipative feedback does not im- 281, 1987. prove the optimal resolution of incoher- ent force detection. Nat. Nano., 8(7):470, [172] R. X. Yan, D. Gargas, and P. D. 2013. Yang. Nanowire photonics. Nat. Phot., 3(10):569–576, 2009. [164] D. Vitali, S. Gigan, A. Ferreira, H. R. Bhm, P. Tombesi, A. Guerreiro, V. Ve- [173] C. Zhao, L. Ju, Y. Fan, S. Gras, B. J. J. dral, A. Zeilinger, and M. Aspelmeyer. Slagmolen, and e. al. Observation Optomechanical entanglement between a of three-mode parametric interactions in movable mirror and a cavity field. Phys. long optical cavities. Phys. Rev. A, Rev. Lett., 98(3):030405, 2007. 78(2):023807, 2008.

[165] D. Vitali, S. Mancini, and P. Tombesi. Optomechanical scheme for the detection of weak impulsive forces. Phys. Rev. A, 64(5):051401, 2001.

[166] A. Vorontsov and J. A. Sauls. Spectrum of third sound cavity modes on superfluid he-3 films. J. Low Temp. Phys., 134(3- 4):1001–1008, 2004.

[167] D. Walls and G. Milburn. Quantum Op- tics. Springer, 2008.

[168] L. Wang and B. Li. Thermal logic gates: Computation with phonons. Phys. Rev. Lett., 99(17):177208, 2007.

90