Warwick P. Bowen’s research while affiliated with ARC Centre of Excellence for Engineered Quantum Systems and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (345)


Fig. 2. (a) SEM of PhC region of photonic hydrophone. (b) Electric field inside and around PhC region of device and coating. (c) Plot of the change in eigenfrequency of the optical mode of our hydrophone í µí¼” against an applied pressure í µí¼Ž, for the photoelastic contributions from both the silicon device (black) and the PDMS coating (blue). These results are obtained from the FEM simulation. The inset figure displays the contribution from the silicon device to the eigenfrequency shift on GPa-scale.
Fig. 3. Diagram of experiment set-up used for characterisation of photonic hydrophone. The blue-shaded region represents the water-filled experiment chamber that our hydrophone, the commercial hydrophone and PZT are submerged in. The network response from 0-250 kHz and a single tone response at 30 kHz from our hydrophone (see inset figure on left) and the optical reflection spectrum of our hydrophone while submerged underwater (see inset figure on right) are also displayed.
Fig. 5. (a) Deployment of photonic hydrophone in a wave flume. (i) The photonic hydrophone and PZT submerged underwater in the wave flume. (ii) The photonic hydrophone mounted on a metal block. (iii) Microscope image of a prepared photonic hydrophone. (b) Network response of the photonic hydrophone. The blue line represents the network response, with the change to a light blue line indicating when the response sinks into the noise floor. The red line (dashed and solid) is the noise floor. (c) SNR of a 30 kHz single tone response with varying separation distance between the PZT and photonic hydrophone. The blue dots are the data points and the red dashed line corresponds to the linear trend-line.
Fibre-coupled photonic crystal hydrophone
  • Preprint
  • File available

January 2025

·

38 Reads

Lauren R. McQueen

·

Nathaniel Bawden

·

Benjamin J. Carey

·

[...]

·

Many applications, including medical diagnostics, sonar and navigation rely on the detection of acoustic waves. Photonic hydrophones demonstrate comparable sensitivity to piezoelectric-based hydrophones, but with significantly reduced size, weight and power requirements. In this paper we demonstrate a micron-sized free-standing silicon photonic hydrophone. We demonstrate sensitivity on the order of \simmPa/Hz\sqrt{\textrm{Hz}} from 10-200 kHz, with a minimum detectable pressure of 145 μ\muPa/Hz\sqrt{\textrm{Hz}} at 22 kHz. We also deployed our hydrophone in a wave flume to evaluate its suitability for underwater measurement and communication. Our hydrophone matches the sensitivity of commercial hydrophones, but is many orders of magnitude smaller in volume, which could enable high spatial resolution imaging of micron-sized acoustic features (i.e., living cell vibrations). Our hydrophone could also be used in underwater communication and imaging applications.

Download

Silicon double-disk optomechanical resonators from wafer-scale double-layered silicon-on-insulator

October 2024

·

17 Reads

·

1 Citation

Whispering Gallery Mode (WGM) optomechanical resonators are a promising technology for the simultaneous control and measurement of optical and mechanical degrees of freedom at the nanoscale. They offer potential for use across a wide range of applications such as sensors and quantum transducers. Double-disk WGM resonators, which host strongly interacting mechanical and optical modes co-localized around their circumference, are particularly attractive due to their high optomechanical coupling. Large-scale integrated fabrication of silicon double-disk WGM resonators has not previously been demonstrated. In this work, we present a process for the fabrication of double-layer silicon-on-insulator wafers, which we then use to fabricate functional optomechanical double silicon disk resonators with on-chip optical coupling. The integrated devices present experimentally observed optical quality factors of the order of 10⁵ and a single-photon optomechanical coupling of approximately 15 kHz.


Fast biological imaging with quantum-enhanced Raman microscopy

September 2024

·

78 Reads

·

2 Citations

Stimulated Raman scattering (SRS) microscopy is a powerful label-free imaging technique that probes the vibrational response of chemicals with high specificity and sensitivity. High-power, quantum-enhanced SRS microscopes have been recently demonstrated and applied to polymers and biological samples. Quantum correlations, in the form of squeezed light, enable the microscopes to operate below the shot noise limit, enhancing their performance without increasing the illumination intensity. This addresses the signal-to-noise ratio (SNR) and speed constraints introduced by photodamage in shot noise-limited microscopes. Previous microscopes have either used single-beam squeezing, but with insufficient brightness to reach the optimal ratio of pump-to-Stokes intensity for maximum SNR, or have used twin-beam squeezing and suffered a 3 dB noise penalty. Here we report a quantum-enhanced Raman microscope that uses a bright squeezed single-beam, enabling operation at the optimal efficiency of the SRS process. The increase in brightness leads to multimode effects that degrade the squeezing level, which we partially overcome using spatial filtering. We apply our quantum-enhanced SRS microscope to biological samples and demonstrate quantum-enhanced multispectral imaging of living cells. The imaging speed of 100×100 pixels in 18 seconds allows the dynamics of cell organelles to be resolved. The SNR achieved is compatible with video-rate imaging, with the quantum correlations yielding a 20% improvement in imaging speed compared to shot noise-limited operation.


(a) Schematic of a single Duffing oscillator and its applications in various computing platforms. Images of a nanomechanical circuit, superconducting circuit, and optical circuit reproduced from Refs.8,36,37, respectively. (b) Frequency/drive response of Duffing oscillator. Solid (dashed) lines represent stable (unstable) solutions of the Duffing equation. Jumps between the stable solutions are labelled with up and down arrows. (c) Time dynamics of a driven, damped Duffing oscillator subject to an SEU. The parameters of the oscillator are given by: m=10-12\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$m = 10^{-12}$$\end{document} kg, Γ=105\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\Gamma = 10^{5}$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document}, ω0=106\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega _0 = 10^{6}$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document}, α=3×1022\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\alpha = 3 \times 10^{22}$$\end{document}m-2s-2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {m}^{-2} s^{-2}$$\end{document}, ω=1.152×106\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega = 1.152 \times 10^{6}$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document}, F=5×10-7\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$F = 5 \times 10^{-7}$$\end{document} N, Δp=6×10-12\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\Delta p = 6 \times 10^{-12}$$\end{document}kg.m.s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {kg.m.s}^{-1}$$\end{document}. The oscillator is first initialised in the ‘0’ state, but then subject to an SEU which leads to a bit flip. The displacement is expressed in arbitrary units, such that the magnitude of a ‘1’ signal is 1.
(a) Schematic of mechanical error correction system, composed of three coupled Duffing oscillators. This functions as a majority voting system and can correct for single SEUs. (b) Time dynamics of error correction device. The coupled oscillators are initialised in their ‘0’ states and after 4.7 periods of oscillation, an impulse is applied to the third oscillator. The third oscillator temporarily transitions into its ‘1’ state, but quickly equilibrates back. The parameters of the oscillator are given by: m=10-12\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$m = 10^{-12}$$\end{document} kg, Γ=105\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\Gamma = 10^{5}$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document}, ω0=106\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega _0 = 10^{6}$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document} , α=3×1022\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\alpha = 3 \times 10^{22}$$\end{document}m-2.s-2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {m}^{-2}.s^{-2}$$\end{document}, ω=1.152×106\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega = 1.152 \times 10^{6}$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document}, F=1.048×10-6\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$F=1.048 \times 10^{-6}$$\end{document} N, Δp=6×10-12\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\Delta p = 6 \times 10^{-12}$$\end{document}kg.m.s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {kg.m.s}^{-1}$$\end{document}, β=2×1011\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\beta = 2 \times 10^{11}$$\end{document}s-2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-2}$$\end{document}. (c) Phase map of error correction device. The map is divided into four main regions, ‘1’ bias, ‘0’ bias, initialise and error correction. This map is produced using the same parameters as (b). The y-axis is normalised by Fcrit\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$F_{crit}$$\end{document}, which we define as the minimum drive force required to sustain a ‘111’ state. Here, Fcrit=1.68×10-7\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$F_{crit} = 1.68 \times 10^{-7} $$\end{document} N at ω=8.3×105\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega = 8.3 \times 10^5$$\end{document}s-1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\hbox {s}^{-1}$$\end{document}.
Single event upset on a 3-bit majority system. (a) Instantaneous energy of coupled system. Oscillators 1 and 2 are represented by the yellow trajectory and oscillator 3 is represented by the purple trajectory. The phase of the impulse relative to motion of the oscillator is indicated by the diagram on the top right corner of the figure. Since the amplitude of oscillator 3 is momentarily increased, its resonance frequency is up-shifted and it decouples from the other oscillators. The energy required for this to occur is indicated by a horizontal dashed line, and the region of decoupling is represented by grey shading. (b) Probability of error-correction failure (i.e. error occurs from impulse) with increasing amplitude of impulse. The horizontal axis is normalised to the maximum momentum of the oscillator when in the ‘1’ state. Inset: probability of error-correction failure for a larger range of impulse amplitudes. Grey region is the range of the main figure. Note that at high impulse amplitudes (i.e. Δp/pmax,`1'>8.5\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\Delta p/p_{\text {max,`1'}} > 8.5 $$\end{document}) the error-correction is always successful.
Simulated probability of system failure per event as a function of event probability per time interval. Red and blue dots represent the simulation results of single and coupled systems, respectively. Red and blue dashed lines represent the predicted model for single oscillator and majority vote of three independent oscillators, respectively. Red and blue solid lines represent the corrected model for single oscillator and coupled oscillators, respectively. Simulation parameters are the same as in Fig. 2.
Engineering error correcting dynamics in nanomechanical systems

September 2024

·

42 Reads

Nanomechanical oscillators are an alternative platform for computation in harsh environments. However, external perturbations arising from such environments may hinder information processing by introducing errors into the computing system. Here, we simulate the dynamics of three coupled Duffing oscillators whose multiple equilibrium states can be used for information processing and storage. Our analysis reveals that, within experimentally relevant parameters, error correcting dynamics can emerge, wherein the system’s state is robust against random external impulses. We find that oscillators in this configuration have several surprising and attractive features, including dynamic isolation of resonators exposed to extreme impulses and the ability to correct simultaneous errors.


FIG. 3. Leading-order contributions to the dark matter-phonon scattering rate. Phonons are represented by double lines.
FIG. 4. Projected 90% CL upper limits on the dark matternucleon cross section at ODIN assuming a run time of 100 days.
FIG. 5. Relationship between the optical wavelength λ, the corresponding Brillouin mode frequency Ω ¼ 2πc s =λ (left axis), and the phonon occupation of the Brillouin mode at 4 mK (right axis).
Optomechanical dark matter instrument for direct detection

August 2024

·

46 Reads

·

6 Citations

Physical Review D

We propose the Optomechanical Dark-matter Instrument (ODIN), based on a new method for the direct detection of low-mass dark matter. We consider dark matter interacting with superfluid helium in an optomechanical cavity. Using an effective field theory, we calculate the rate at which dark matter scatters off phonons in a highly populated, driven acoustic mode of the cavity. This scattering process deposits a phonon into a second acoustic mode in its ground state. The deposited phonon ( μ eV range) is then converted to a photon (eV range) via an optomechanical interaction with a pump laser. This photon can be efficiently detected, providing a means to sensitively probe keV scale dark matter. We provide realistic estimates of the backgrounds and discuss the technical challenges associated with such an experiment. We calculate projected limits on dark matter–nucleon interactions for dark matter masses ranging from 0.5 to 300 keV and estimate that a future device could probe cross sections as low as O ( 10 − 32 ) cm 2 . Published by the American Physical Society 2024


Fig. 1. Axisymmetric eigenmode FEM simulation results of the double-disk resonators. (a) The optical field distribution of a hybridized (transverse electric) whispering gallery mode. (b) Mechanical displacement profile of the symmetric mechanical resonance (inset: asymmetric mode-profile). (c) Effective refractive index of the optical mode (cf. Eq. 1) with differing disk separation (insets: mode profiles with 0 nm & 300 nm spacings). (d) Optomechanical coupling rate of the disks for varying separations (the vertical line corresponds to the 60 nm used in our fabricated devices).
Fig. 2. Wafer scale production of the Si double-disks from two standard silicon-oninsulator (SOI) wafers (not to scale). (a) Initial SOI wafer stacks. (b) Oxidation of the top layers prior to wafer bonding, followed by backside etching from one side to remove one carrier and buried oxide layer (detail in Fig. S2). (c) The resultant double-layer SOI wafer. (d) The wafers are then diced into 15 mm × 15 mm chips and patterned with electron beam lithography, etched and undercut to create the devices (detail in Fig. S4). (e) A released double-disk resonator. Blue, Si; magenta, silica; green, electron-beam resist.
Fig. 3. (a) Optical micrograph of a fabricated device (a wide-view photograph is shown in Fig. S5). Scanning Electron Microscope (SEM) images of a (b) double-disk resonator; (c) suspended waveguide; and (d) edge of the released area.
Fig. 4. Optical transmission spectrum in ambient conditions. The sharp dips are optical resonances of the double-disks. The inset shows the transmission of a narrow span (0.2 THz).
Fig. 6. (a) Sideband-detuned PSD of the double-disk motion (blue) and fit according to Eq. S.1 (green), with vertical bars marking the center frequencies of each resonance. Data were acquired with a detuning of −0.744 GHz. (b) The PSDs as a function of optical detuning. Here the dotted white lines correspond to the centre of the fitted Lorentzian (i.e. Ω í µí±– ). (c) Optical spring effect (Eq. 6) frequency shifts extracted from (b). The color of each curve corresponds to those allocated in (a). Experimental points are shown as circles and fits are shown as dashed/dotted lines.
Silicon Double-Disk Optomechanical Resonators from Wafer-Scale Double-Layered Silicon-on-Insulator

July 2024

·

53 Reads

Whispering gallery mode (WGM) optomechanical resonators are a promising technology for the simultaneous control and measurement of optical and mechanical degrees of freedom at the nanoscale. They offer potential for use across a wide range of applications such as sensors and quantum transducers. Double-disk WGM resonators, which host strongly interacting mechanical and optical modes co-localized around their circumference, are particularly attractive due to their high optomechanical coupling. Large-scale integrated fabrication of silicon double-disk WGM resonators has not previously been demonstrated. In this work we present a process for the fabrication of double-layer silicon-on-insulator wafers, which we then use to fabricate functional optomechanical double silicon disk resonators with on-chip optical coupling. The integrated devices present an experimentally observed optical quality factors of the order of 10^5 and a single-photon optomechanical coupling of approximately 15 kHz.



Acoustically driven single-frequency mechanical logic

May 2024

·

58 Reads

·

1 Citation

Physical Review Applied

Nanomechanical computers promise robust, low-energy information processing; however, to date, electronics have generally been required to drive gates, and logical operations have generally involved bits with different oscillation frequencies. This limits the scalability of nanomechanical logic. Here we demonstrate an acoustically driven logic gate that has a single frequency of operation. Our gate uses the bistability of a nonlinear mechanical resonator to define logical states. These states are efficiently coupled into and out of the gate via nanomechanical waveguides, providing the mechanical equivalent of electrical wires and allowing purely mechanical information transfer. Since the inputs and output all share the same frequency, they are compatible with cascaded chains of gates. Our architecture is CMOS compatible, and with miniaturization could allow an energy cost that approaches the fundamental Landauer limit. Together this presents a pathway towards large-scale nanomechanical computers.


Dynamic interaction between chiral currents and surface waves in topological superfluids: a pathway to detect Majorana fermions?

April 2024

·

41 Reads

Despite extensive experimental efforts over the past two decades, the quest for Majorana fermions in superconductors remains inconclusive. We propose an experimental method that can conclusively confirm, or rule out, the existence of these quasiparticles: Firstly, we shift focus from superconductors, whose very topological nature is disputed, to the unambiguous topological superfluid helium-3. Secondly, we identify the interaction between surface waves and the chiral Majorana current in the bulk of a topological superfluid of varying density. The proposed experiment provides a path towards the detection of the Majorana fermion, an 80-year-old theoretical prediction. It is realistically achievable based on the advent of microscopic superfluid resonators coupled to optical cavities. The proposal may open the door to experiments ranging from simulations of exotic cosmological particles to topological acoustics and fault tolerant quantum computing.


(a) Principle of operation for the waveguide-coupled acoustic directional emitter. Two electrodes separated by a distance L act as discrete wave sources. A controllable phase difference ϕ between the two electrode drives is chosen such that the waves they produce constructively interfere on the right hand side and destructively interfere on the left hand side, producing unidirectional rightward power emission. Another phase difference (not illustrated) would produce constructive interference on the left hand side and destructive interference on the right hand side. (b) Schematic illustration of the physical implementation in an on-chip acoustic waveguide.²⁶ (c) The electrodes introduce reflections, which we model through an amplitude reflection coefficient r and transmission coefficient t.
(a) Optical microscope image of the acoustic spiral waveguide with length 21 cm. Iridescence arises due to the sub-micron hole pattern (not visible at this scale) used to release the membranes.²⁶ (b) Close-up of the actuating electrodes. (c) 3D illustration of the experimental setup. The waveguide is a released membrane of SiN (blue). Two gold electrodes on the waveguide drive acoustic waves (not to scale), with the doped silicon substrate serving as the ground plane of the capacitive actuation. The electrodes are powered by alternating currents provided by the signal generator and tuned by variable attenuators and phase shifters. The probe laser is delivered through the lensed fiber above the waveguide.
(a) Typical photocurrent power spectral density observed on a side of the electrodes when the two wave trains are constructively interfering. f mech = 8.69   MHz is the mechanical drive frequency. Sidebands are the result of electronic noise. (b) Typical distribution when there is destructive interference. (c) Open circles: value of the photocurrent power spectral density at f = f mech + f L O, taken at fixed locations on either side of the electrodes. Because of the acoustic reflectivity discussed earlier, as the phase difference was incremented from − 188 ° to 210 °, the voltages to each electrode were also incremented between 200   mV and 1   V. Solid lines: fit using Eq. (1) with r = 0.9. Dashed lines: simulation data from the COMSOL model (for more details, see the supplementary material).
2D scans of the waveguide, while the electrode drives were set for emission toward electrode A (top) and electrode B (bottom). Each surface plot is produced by assembling a grid of 201 × 11 measurements, covering an area of 140 × 34  μ m 2. The output is smoothed in the x direction by a Gaussian filter of standard deviation of 8 pixels or ∼ 11   μ m.
Directional emission in an on-chip acoustic waveguide

January 2024

·

38 Reads

·

2 Citations

Integrated acoustic circuits leverage guided acoustic waves for applications ranging from radio frequency filters to quantum state transfer, biochemical sensing, and nanomechanical computing. In many applications, it is desirable to have a method for unidirectional acoustic wave emission. In this work, we demonstrate directional emission in an integrated single-mode, on-chip membrane waveguide, demonstrating over 99.9% directional suppression and reconfigurable directionality. This avoids both loss and unwanted crosstalk, allowing the creation of more complex and compact phononic circuits.


Citations (41)


... This experimentally observed sensitivity is consistent with that predicted by our modelling (see Eq. 2), SN min = 4.1 mPa/ √ Hz, for an optical resonance with = 0.5 nm and 0 = 3164.8. The and 0 values are extracted by fitting a Fano resonance [56,57] to the experimental data, as described in Supplement 1 (Sec. S2, Fig. S2). ...

Reference:

Fibre-coupled photonic crystal hydrophone
Silicon double-disk optomechanical resonators from wafer-scale double-layered silicon-on-insulator

... Only the lowest temperature reheating scenarios are plausibly within the reach of future experiments. While a number of experimental strategies have been suggested to search for dark matter in this low mass regime, mostly exploiting darkmatter-induced phonon excitations in polar materials or superfluid helium [27][28][29][30][31][32][33][34], optimistic estimates place their sensitivities at around σ χp ∼ 10 −44 -10 −45 cm 2 , several orders of magnitude larger than our predictions for scalar dark matter with even the lowest possible reheat temperatures. ...

Optomechanical dark matter instrument for direct detection

Physical Review D

... Similarly, the use of squeezed light sources, which can reduce photon number fluctuations at the optimal operating point of classical Raman experiments, points to exciting opportunities. Proof-of-principle experiments have already demonstrated that quantum-enhanced Raman measurements are within reach, whether with narrowband fields [27] or ultrafast pulses [28][29][30], and in stimulated Brillouin scattering [31,32]. These advances, along with parallel progress in stimulated emission imaging [33], highlight the role of squeezed light in near-future nonlinear measurement setups. ...

Fast biological imaging with quantum-enhanced Raman microscopy

... [30]. We show that acoustic input bits can be coupled into the gate with 99% efficiency and that they can drive transitions between two bistable states [31][32][33]. This enables an approach to nanomechanical computing [34] for which we demonstrate a universal set of purely mechanical logic operations. ...

Directional emission in an on-chip acoustic waveguide

... One particularly promising area of research, where further advances may be imagined, lies in the use of quantumenhanced measurement strategies that are designed within the framework of quantum metrology [13]. Uniting parameter estimation theory with quantum measurements, this framework establishes fundamental precision limits in diverse settings [14,15]. ...

Quantum light microscopy
  • Citing Article
  • December 2023

Contemporary Physics

... Notably, benefiting from the resonant enhancement of both optical and mechanical responses, many outstanding cavity optomechanical sensors have been achieved for various physical quantities . Recently, cavity optomechanical magnetometers have attracted extensive attentions for measurement of weak alternating current (AC) magnetic fields [35][36][37][38][39][40][41][42][43][44], owing to the unique advantages of room-temperature operation, magnetic shielding free, large frequency range (from kHz to GHz), and high sensitivity. However, they are unsuitable for measurement of direct current (DC) magnetic fields, which generally coexist with AC magnetic fields in most practical application scenarios [45,46]. ...

Waveguide-integrated chip-scale optomechanical magnetometer

... In addition, the capability to gain control over quantum systems at large mass scales is increasing rapidly, with the preparation of non-classical states of mechanical oscillators [6][7][8][9][10]. This progress has driven new ideas for signatures of the quantum nature of gravity [11], such as quantum features of gravity testing phenomenological modifications to known physics [12][13][14][15], including tests of collapse models [16][17][18][19][20][21][22][23][24][25][26], quantum features of gravitational source masses [27][28][29][30][31][32][33][34][35][36], and quantum features of gravitational radiation [37][38][39][40][41][42]. In light of these advancements it was recently proposed that single gravitons can be detected through projective measurements of the energy of a ground state cooled kg-scale resonant mass detector [43], which would provide a gravitational version of the photo-electric effect. ...

Testing spontaneous wavefunction collapse with quantum electromechanics

... To date, various QST machines have been developed to meet diverse functionalities required by the quantum information tasks. For examples, optical-to-mechanical QST machine performs the transfer of a quantum state between optical photon and microwave states [17,18]. Continuousvariable-to-discrete-variable QST machine is capable of transferring arbitrary continuous-variable quantum state into a few discrete qubits and back [19]. ...

Continuous Optical-to-Mechanical Quantum State Transfer in the Unresolved Sideband Regime
  • Citing Article
  • June 2023

Physical Review Letters

... Even after more than four decades since its initial applications, optical tweezers are being constantly improved leading to new discoveries [6,10]. At present there are several designs for single, dual and multiple optical traps that are in use for a variety of applications [6,13]. Of these, dual-trap optical tweezers (DTOT) are particularly interesting because of their applications in measurement of microscopic interaction potentials [14,15], optical binding studies [16], two point micro-rheology [17][18][19][20], single molecule force measurement [6,[21][22][23], etc. ...

Roadmap for Optical Tweezers

... Tools like atomic force microscopy (AFM) [50,51] and micropipette aspiration [52] give the mechanical properties of the cell's outer surface and the underlying cortex. Magnetic [53,54] or optical tweezers [55,56,57], as well as particle tracking microrheology (PTM) [58] enable researchers to explore the distinct mechanical properties and particle mobilities within the cell, which are in general drastically different to the cortex mechanics. ...

Roadmap for Optical Tweezers 2023