HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: mhchem
  • failed: stix

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2212.13532v4 [physics.optics] 23 Jan 2024
thanks: These authors contributed equally to this work.thanks: These authors contributed equally to this work.

Direct laser-written optomechanical membranes in fiber Fabry-Perot cavities

Lukas Tenbrake ID Institute of Applied Physics, University of Bonn, Germany    Alexander Faßbender Institute of Physics, University of Bonn, Germany    Sebastian Hofferberth ID Institute of Applied Physics, University of Bonn, Germany    Stefan Linden ID Institute of Physics, University of Bonn, Germany    Hannes Pfeifer ID [email protected],
Current address: Department of Microtechnology and Nanoscience, Chalmers University of Technology, Gothenburg, Sweden
Institute of Applied Physics, University of Bonn, Germany
(January 23, 2024)
Abstract

Integrated micro- and nanophotonic optomechanical experiments enable the manipulation of mechanical resonators on the single phonon level. Interfacing these structures requires elaborate techniques limited in tunability, flexibility, and scaling towards multi-mode systems. Here, we demonstrate a cavity optomechanical experiment using 3D-laser-written polymer membranes inside fiber Fabry-Perot cavities. Vacuum coupling rates of g0/2π30 kHzsubscript𝑔02𝜋times30kilohertzg_{0}/2\pi\approx$30\text{\,}\mathrm{kHz}$italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2 italic_π ≈ start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG to the fundamental megahertz mechanical mode are reached. We observe optomechanical spring tuning of the mechanical resonator frequency by tens of kilohertz exceeding its linewidth at cryogenic temperatures. The direct fiber coupling, its scaling capabilities to coupled resonator systems, and the potential implementation of dissipation dilution structures and integration of electrodes make it a promising platform for fiber-tip integrated accelerometers, optomechanically tunable multi-mode mechanical systems, and directly fiber-coupled systems for microwave to optics conversion.

I Introduction

Cavity optomechanical experiments have been implemented on a multitude of different platforms [1] ranging from the canonical movable end mirror of a Fabry-Perot cavity [2], over membranes in cavities [3, 4], toroidal[5] and optomechanical crystal resonators [6], down to the vibrational modes of molecules in a plasmonic picocavity [7]. This platform diversification and constant improvement advanced the field during the past years leading to key achievements like ground-state cooling of mechanical resonators [8, 9], optomechanical state teleportation experiments [10], efficient microwave to optical conversion [11, 12], or sensing of the mechanical resonator below the standard quantum limit [13, 14, 15]. Among the current challenges in the field are the realization and addressing of multi-mode optomechanical systems and the integration of optomechanical elements for different sensing applications.

Optomechanical devices with high optical field concentration and correspondingly large optomechanical coupling are usually realized with on-chip platforms [6, 5, 16]. A different approach is taken by miniaturized Fabry-Perot cavities with concave mirror structures fabricated on optical fiber tips [17, 18]. These fiber Fabry-Perot cavities (FFPCs) have been established as a platform for light-matter interaction during the past years, including experiments on atoms inside FFPCs for photonic qubits or quantum networks [19, 20], and realizations of optomechanical experiments [21, 22, 23, 24, 25, 26, 27]. They feature a direct fiber-coupled optical access, small cavity lengths, high optical finesse, and an open resonator volume. This allows to introduce both conventional membrane-type resonators as well as more unconventional resonators like standing waves in liquid Helium.

With its first demonstration in the 1990s [28], 3D direct laser writing (DLW) enabled the fabrication of free-formed three- dimensional polymeric structures with sub-micrometer resolution. It led to the miniaturization of on-chip optical components [29], waveguides [30], and mechanical structures [31] including 3D acoustic metamaterials [32], and has been used for mask applications [18]. Apart from large planar substrates, writing on fiber-ends has been tackled for applications in endoscopy [33] or sensing [34, 35].

In this article, we demonstrate the integration of mechanical polymer membrane resonators into highly miniaturized FFPCs. We realize a miniaturized, fiber-coupled membrane-in-the-middle (MIM) experiment [3, 4] with superior scaling capabilities due to the 3D DLW fabrication process both integrated into highly stable monolithic FFPCs [36] and on flat distributed Bragg reflector (DBR) substrates. We analyze the achievable optomechanical coupling strength, and demonstrate a dispersive optomechanical spring effect tuning of the mechanical resonace in the presence of a thermal optical nonlinearity exceeding the mechanical linewidth at cryogenic temperatures. To characterize the basic internal material properties no elaborate dissipation dilution structures were used in this first proof-of-principle study.

Our results pave the way for using highly flexible DLW structure fabrication for optomechanical resonators. DLW enables new realizations of high-sensitivity fiber-cavity integrated accelerometers, and mechanical multi-mode structures that are interfaced using FFPCs, including multiple membranes inside a miniaturized Fabry-Perot cavity [37] (see methods). Furthermore, extended mechanical metamaterials of membrane resonators on DBR substrates for controlling vibrations in a thin film can be realized and combination with other light-matter interfaces as 2D materials and quantum emitters can be envisaged.

II Results

Polymer membrane in a fiber cavity

Refer to caption
Figure 1: Overview of the system. In a a schematic overview of the experimental structure is shown. The polymer drum resonator is fabricated on a highly reflective DBR substrate. The fiber mirror is positioned above the polymer drum and realizes a fiber Fabry-Perot cavity with the DBR beneath enclosing the drum membrane. The inset shows a microscope picture of an approaching fiber mirror (diameter: 125 µmtimes125micrometer125\text{\,}\mathrm{\SIUnitSymbolMicro m}start_ARG 125 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG) to a polymer drum array. In b the magnitude of the displacement field 𝐮𝐮\mathbf{u}bold_u of the fundamental drum mode is shown as retrieved from finite element simulations. The height hdrumsubscriptdrumh_{\text{drum}}italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT of the polymer drum supports and the drum membrane thickness tdrumsubscript𝑡drumt_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT determine the positions of the optical cavity mode intensity maxima and minima with respect to the membrane interfaces. In c an SEM micrograph of a fabricated rectangular polymer drum structure is shown.

The mechanical resonator in our experiments is a drum-like polymer membrane of 12 µm1times2micrometer1-$2\text{\,}\mathrm{\SIUnitSymbolMicro m}$1 - start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG thickness. It is fabricated using 3D direct laser writing (see methods) on top of a highly reflective DBR mirror (10 ppmtimes10ppm10\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG transmission111High reflectivity range: 750 nmtimes750nanometer750\text{\,}\mathrm{nm}start_ARG 750 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG to 805 nmtimes805nanometer805\text{\,}\mathrm{nm}start_ARG 805 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG) located either on an extended substrate or on an optical fiber-tip. A typical rectangular membrane used here spans 45 µm×60 µmsimilar-toabsenttimes45micrometertimes60micrometer\sim$45\text{\,}\mathrm{\SIUnitSymbolMicro m}$\times$60\text{\,}\mathrm{% \SIUnitSymbolMicro m}$∼ start_ARG 45 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG × start_ARG 60 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG and is placed on support bars that suspend the membrane 510 µm5times10micrometer5-$10\text{\,}\mathrm{\SIUnitSymbolMicro m}$5 - start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG above the mirror surface (see Fig. 1 a and c). The supports reduce the free membrane surface by 5 µmtimes5micrometer5\text{\,}\mathrm{\SIUnitSymbolMicro m}start_ARG 5 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG on each side. More elaborate geometries like multi-membrane structures or suspensions like soft-clam** [38] can be directly realized in the 3D DLW fabrication. Finite-element simulations of the fundamental mechanical mode, see Fig. 1 b and methods section, yield an expected mechanical resonance frequency of 2.1 MHztimes2.1megahertz2.1\text{\,}\mathrm{MHz}start_ARG 2.1 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG and an internal quality factor of 20similar-toabsent20\sim 20∼ 20 due to the comparably large elastic loss tangent of the 3D DLW polymer resist material at ambient conditions. As no specific dissipation dilution is implemented in this simple design, the mechanical quality factor reflects the intrinsic Q𝑄Qitalic_Q value (Qintsubscript𝑄intQ_{\text{int}}italic_Q start_POSTSUBSCRIPT int end_POSTSUBSCRIPT) of the material [39].

The membrane is integrated into an optical FFPC by approaching a fiber mirror – a fiber-tip with a concave-shaped facet and high-reflection coated surface [17, 18]. The cavity length is adjusted to 30 µmsimilar-toabsenttimes30micrometer\sim$30\text{\,}\mathrm{\SIUnitSymbolMicro m}$∼ start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG, but can be scanned using a piezo-electric element attached to the fiber mirror. The optical fiber leading to the approached fiber mirror serves as both input and output of the cavity. The transmission1 of 2000 ppmtimes2000ppm2000\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}start_ARG 2000 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG is chosen to retrieve a single-sided cavity geometry, with an approximate balance of the cavity-to-input-fiber coupling rate and the internal cavity losses. The internal cavity losses are dominated by scattering from the polymer surface that exhibits a roughness222RMS surface profile variation in AFM measurement of <5 nmabsenttimes5nanometer<$5\text{\,}\mathrm{nm}$< start_ARG 5 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG after post-fabrication polishing using an oxygen plasma ashing process. As the air-polymer interfaces of the membrane have a large transmittance, the field of the optical cavity extends through the membrane and over both open cavity domains. The optical cavity mode spectrum is characterized by scanning the cavity resonance over a probe laser tone with modulated sidebands (see methods, and [40, 36]). We measure optical linewidths of κ/2π𝜅2𝜋\kappa/2\piitalic_κ / 2 italic_π of few gigahertz corresponding to finesse values of =1400±300 timesuncertain1400300absent\mathcal{F}=$1400\pm 300\text{\,}$caligraphic_F = start_ARG start_ARG 1400 end_ARG ± start_ARG 300 end_ARG end_ARG start_ARG times end_ARG start_ARG end_ARG for the case of an intensity maximum on one of the polymer-air interfaces and a minimum on the other. As scattering from the membrane surface is the dominant optical loss mechanism, the finesse reaches the empty cavity value for the case of intensity minima on both interfaces.

In order to probe the mechanical mode, we lock the cavity to the probe laser using a feedback loop on a Pound-Drever-Hall (PDH) error signal [41, 42]. The thermal excitation of the membrane is transduced to cavity frequency noise visible in the noise spectrum of the calibrated PDH error signal, which is recorded using an electrical spectrum analyzer. The measured mechanical frequencies and linewidths are in good agreement with the finite element simulation with mechanical resonance frequencies between 1 and 4 MHztimes4megahertz4\text{\,}\mathrm{MHz}start_ARG 4 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG depending on the particular membrane geometry. The calibration of the PDH error signal slope at the lock point together with the temperature of the environment furthermore allows a quantification of the vacuum optomechanical coupling rate of the membrane modes (see methods, and [43, 36]).

Characterization of the optomechanical coupling

A linear, dispersive optomechanical coupling manifests as a shift of the optical cavity frequency ΔωcavΔsubscript𝜔cav\Delta\omega_{\text{cav}}roman_Δ italic_ω start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT upon a displacement ΔzΔ𝑧\Delta zroman_Δ italic_z of the membrane [3, 4]. In turn, the cavity field photons act on the pliable membrane through their radiation pressure, closing the interaction loop. The magnitude of this effect strongly depends on the geometry of the MIM system, in particular on the optical intensity on both sides of the membrane. A difference in intensity will cause a non-vanishing net radiation pressure on the membrane resulting in an optomechanical interaction. The interaction strength is given by the vacuum optomechanical coupling rate g0=ωcavzzzpfG(1)zzpfsubscript𝑔0subscript𝜔cav𝑧subscript𝑧zpfsuperscript𝐺1subscript𝑧zpfg_{0}=\frac{\partial\omega_{\text{cav}}}{\partial z}z_{\text{zpf}}\coloneqq-G^% {(1)}z_{\text{zpf}}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG ∂ italic_ω start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_z end_ARG italic_z start_POSTSUBSCRIPT zpf end_POSTSUBSCRIPT ≔ - italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT italic_z start_POSTSUBSCRIPT zpf end_POSTSUBSCRIPT with zzpfsubscript𝑧zpfz_{\text{zpf}}italic_z start_POSTSUBSCRIPT zpf end_POSTSUBSCRIPT the mechanical resonator’s zero-point motion amplitude. As we fix our total cavity length to 30 µmsimilar-toabsenttimes30micrometer\sim$30\text{\,}\mathrm{\SIUnitSymbolMicro m}$∼ start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG, the relevant parameters determining the achievable coupling are the polymer drum support height hdrumsubscriptdrumh_{\text{drum}}italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT and the membrane thickness tdrumsubscript𝑡drumt_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT of the drum (see Fig. 1 b). Since the latter is comparable to the probe laser wavelength λ𝜆\lambdaitalic_λ, it has a strong effect on the local cavity intensity at the membrane surfaces.

We compute the expected frequency-pull factor G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) in our experiment using two different methods. After finding the resonance condition of the optical cavity field (see methods), the first method considers a small shift ΔzΔ𝑧\Delta zroman_Δ italic_z of the membrane position. Its effect on the cavity resonance frequency as expected from Maxwell’s equations is numerically extracted to determine the frequency-pull. For the second method, we employ perturbation theory for Maxwell’s equations with moving material boundaries [44] (for details, see methods). For this, we use the cavity resonance condition to find the explicit form of the cavity electrical field. The resulting pull-factor (see Eq. 7) scales linearly with the intensity difference on the two membrane-air interfaces and the polymer refractive index npolysubscript𝑛polyn_{\text{poly}}italic_n start_POSTSUBSCRIPT poly end_POSTSUBSCRIPT. Both methods yield identical results. The resulting G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) is shown in Fig. 2 a with some highlighted field configurations in b. Maximum frequency-pull factors of up to 11 GHz nm1times11timesgigahertznanometer111\text{\,}\mathrm{GHz}\text{\,}{\mathrm{nm}}^{-1}start_ARG 11 end_ARG start_ARG times end_ARG start_ARG start_ARG roman_GHz end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_nm end_ARG start_ARG - 1 end_ARG end_ARG end_ARG are predicted. The map features a periodic pattern of well-defined minima and maxima with a periodicity of half the (material) wavelength. The asymmetry of positive vs. negative coupling emerges from the two air domains of the cavity not having equal lengths.

Refer to caption
Figure 2: In a the calculated G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) are shown as a color-map, with {tdrum,hdrumsubscript𝑡drumsubscriptdrumt_{\text{drum}},h_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT} normalized to the respective material wavelength. Measured g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for two exemplary polymer drums are highlighted in green (drum 1111) and purple (drum 2222) and plotted along the third plot-axis. By scanning the probe-wavelength, the measurements of the two drums trace out cuts in the map. Four cavity geometries corresponding to special coupling scenarios are highlighted: and correspond to G(1)=0superscript𝐺10G^{(1)}=0italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = 0,4 and maximize |G(1)|superscript𝐺1|G^{(1)}|| italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT |. In b (bottom) a sketch of the cavity intensity Icavsubscript𝐼cavI_{\text{cav}}italic_I start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT of an exemplary cavity geometry is shown. The cavity intensity at the two polymer-air/vacuum interfaces is shown for the four scenarios from a ( , , , ). In c the measured g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT from image a are plotted against the scanned probe laser wavelength λ𝜆\lambdaitalic_λ for both drum 1 (green) and drum 2 (purple).

We test our model by measuring g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for different probe wavelengths on two exemplary polymer geometries: {tdrum 11.2 µm,hdrum 18.9 µm}formulae-sequencesimilar-tosubscript𝑡drum 1times1.2micrometersimilar-tosubscriptdrum 1times8.9micrometer\{t_{\text{drum 1}}\sim$1.2\text{\,}\mathrm{\SIUnitSymbolMicro m}$,\,h_{\text{% drum 1}}\sim$8.9\text{\,}\mathrm{\SIUnitSymbolMicro m}$\}{ italic_t start_POSTSUBSCRIPT drum 1 end_POSTSUBSCRIPT ∼ start_ARG 1.2 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG , italic_h start_POSTSUBSCRIPT drum 1 end_POSTSUBSCRIPT ∼ start_ARG 8.9 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG } and {tdrum 21.15 µm,hdrum 29.2 µm}formulae-sequencesimilar-tosubscript𝑡drum 2times1.15micrometersimilar-tosubscriptdrum 2times9.2micrometer\{t_{\text{drum 2}}\sim$1.15\text{\,}\mathrm{\SIUnitSymbolMicro m}$,\,h_{\text% {drum 2}}\sim$9.2\text{\,}\mathrm{\SIUnitSymbolMicro m}$\}{ italic_t start_POSTSUBSCRIPT drum 2 end_POSTSUBSCRIPT ∼ start_ARG 1.15 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG , italic_h start_POSTSUBSCRIPT drum 2 end_POSTSUBSCRIPT ∼ start_ARG 9.2 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG }. The calibrated optical cavity frequency noise spectrum 𝒮νν(f)subscript𝒮𝜈𝜈𝑓\mathcal{S}_{\nu\nu}(f)caligraphic_S start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT ( italic_f ) is measured (see methods) and the fundamental mechanical resonance is used to extract g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT at ambient conditions from its expected effect on the cavity frequency noise via [43]:

𝒮νν(f)=2g024π22Ωm2ΓkBT(Ω2Ωm2)2+Γ2Ω2.subscript𝒮𝜈𝜈𝑓2superscriptsubscript𝑔024superscript𝜋22subscriptΩ𝑚Planck-constant-over-2-pi2Γsubscript𝑘𝐵𝑇superscriptsuperscriptΩ2superscriptsubscriptΩ𝑚22superscriptΓ2superscriptΩ2\mathcal{S}_{\nu\nu}(f)=\frac{2g_{0}^{2}}{4\pi^{2}}\cdot\frac{2\Omega_{m}}{% \hbar}\cdot\frac{2\Gamma k_{B}T}{\left(\Omega^{2}-\Omega_{m}^{2}\right)^{2}+% \Gamma^{2}\Omega^{2}}.caligraphic_S start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT ( italic_f ) = divide start_ARG 2 italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ⋅ divide start_ARG 2 roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG roman_ℏ end_ARG ⋅ divide start_ARG 2 roman_Γ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG ( roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG .

Here, f=Ω/2π𝑓Ω2𝜋f=\Omega/2\piitalic_f = roman_Ω / 2 italic_π denotes the noise frequency, ΩmsubscriptΩ𝑚\Omega_{m}roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT the mechanical resonance frequency, Γ/2πΓ2𝜋\Gamma/2\piroman_Γ / 2 italic_π the mechanical linewidth, kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT Boltzmann’s constant and T𝑇Titalic_T the temperature of the polymer membrane. The results are shown in Fig. 2 a and c. As the probe laser wavelength is scanned, the coupling rate changes through the shift of the cavity intensity on the polymer drum surfaces. This corresponds to the diagonal cuts in Fig. 2 a across the coupling landscape for the two representative drum geometries (green – drum 1111, purple – drum 2222). Here, tuning the probe laser wavelength from 760 nmtimes760nanometer760\text{\,}\mathrm{nm}start_ARG 760 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG to 785 nmtimes785nanometer785\text{\,}\mathrm{nm}start_ARG 785 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG is equivalent to a few 100 nmtimes100nanometer100\text{\,}\mathrm{nm}start_ARG 100 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG variation of the geometry333hdrumsubscriptdrumh_{\text{drum}}italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT is chosen to be 8 µmgreater-than-or-equivalent-toabsenttimes8micrometer\gtrsim$8\text{\,}\mathrm{\SIUnitSymbolMicro m}$≳ start_ARG 8 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG for these geometries, such that tuning over the complete probe laser wavelength-range changes the cavity field node/anti-node positions by at least half of the (material) wavelength.. We find a maximum coupling rate of |g0|/2π=33±7 kHzsubscript𝑔02𝜋timesuncertain337kilohertz|g_{0}|/2\pi=$33\pm 7\text{\,}\mathrm{kHz}$| italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | / 2 italic_π = start_ARG start_ARG 33 end_ARG ± start_ARG 7 end_ARG end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG at λ=772 nm𝜆times772nanometer\lambda=$772\text{\,}\mathrm{nm}$italic_λ = start_ARG 772 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG (see Fig. 2 c for drum 2). The sign of g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in Fig. 2 is inferred from the fabricated geometry parameters and the expected asymmetry of positive and negative coupling regions.

Four special cases of the intensity distribution are highlighted in Fig. 2 a. Their corresponding cavity intensity distribution at the polymer membrane surfaces is shown in Fig. 2 b. Vanishing optomechanical coupling is observed for equal intensities on both sides of the membrane, in the simplest case by tdrumsubscript𝑡drumt_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT being a multiple of half the material wavelength. Optimal coupling is achieved, if a cavity field node (anti-node) is located at one side of the polymer membrane in combination with a corresponding anti-node (node) at the other444Note that the quadratic frequency-pull factor is maximized for equal intensities on either side along the connection line of the linear maxima of same tdrumsubscript𝑡drumt_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT. This corresponds to in Fig. 2 b. It also vanishes at tdrumsubscript𝑡drumt_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT being equal to a multiple of half the material wavelength ( ). The maximum expected quadratic coupling is G(2)200 GHz nm2greater-than-or-equivalent-tosuperscript𝐺2times200timesgigahertznanometer2G^{(2)}\gtrsim$200\text{\,}\mathrm{GHz}\text{\,}{\mathrm{nm}}^{-2}$italic_G start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ≳ start_ARG 200 end_ARG start_ARG times end_ARG start_ARG start_ARG roman_GHz end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_nm end_ARG start_ARG - 2 end_ARG end_ARG end_ARG..

Optomechanical spring effect

We now consider dynamical effects of the intracavity photon number on the mechanical modes of the polymer membrane. As we are working in the fast-cavity/Doppler-regime, where the cavity decay rate κΩmmuch-greater-than𝜅subscriptΩ𝑚\kappa\gg\Omega_{m}italic_κ ≫ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, the optomechanical spring effect shifts the fundamental flexural mode frequency ΩmsubscriptΩ𝑚\Omega_{m}roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT as

ΔΩm(Δ)=g02ncav2ΔΔ2+κ2/4ΔsubscriptΩ𝑚Δsuperscriptsubscript𝑔02subscript𝑛cav2ΔsuperscriptΔ2superscript𝜅24\Delta\Omega_{m}(\Delta)={g_{0}}^{2}n_{\text{cav}}\cdot\frac{2\Delta}{\Delta^{% 2}+\kappa^{2}/4}roman_Δ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( roman_Δ ) = italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT ⋅ divide start_ARG 2 roman_Δ end_ARG start_ARG roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 end_ARG (1)

with cavity detuning ΔΔ\Deltaroman_Δ, cavity photon number ncavsubscript𝑛cavn_{\text{cav}}italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT and vacuum optomechanical coupling rate g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [1]. Effects on the mechanical linewidth are neglected.

To measure ΔΩmΔsubscriptΩ𝑚\Delta\Omega_{m}roman_Δ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, we make use of a two-tone measurement scheme. The cavity is locked in a PDH feedback loop to a probe laser beam fixed at 780 nmtimes780nanometer780\text{\,}\mathrm{nm}start_ARG 780 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG with low optical power (<50 µWabsenttimes50microwatt<$50\text{\,}\mathrm{\SIUnitSymbolMicro W}$< start_ARG 50 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_W end_ARG) (for details, see methods). ΩmsubscriptΩ𝑚\Omega_{m}roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is measured from 𝒮νν(f)subscript𝒮𝜈𝜈𝑓\mathcal{S}_{\nu\nu}(f)caligraphic_S start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT ( italic_f ) as extracted from the calibrated error signal of the feedback loop. To impose a frequency shift of the mechanical resonator, a tunable pump-laser (760 nm785 nmtimes760nanometertimes785nanometer$760\text{\,}\mathrm{nm}$-$785\text{\,}\mathrm{nm}$start_ARG 760 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG - start_ARG 785 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG) is additionally coupled into the input-port fiber addressing another optical cavity resonance separated by one free spectral range (5 THzsimilar-toabsenttimes5terahertz\sim$5\text{\,}\mathrm{THz}$∼ start_ARG 5 end_ARG start_ARG times end_ARG start_ARG roman_THz end_ARG) from the probe. The probe path is decoupled from the pump-laser radiation with a narrow pass-band interference filter. The pump-laser is scanned over the optical resonance modulating the detuning and the intracavity photon number ncavsubscript𝑛cavn_{\text{cav}}italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT. During this process, 𝒮νν(f)subscript𝒮𝜈𝜈𝑓\mathcal{S}_{\nu\nu}(f)caligraphic_S start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT ( italic_f ) is recorded to extract ΔΩmΔsubscriptΩ𝑚\Delta\Omega_{m}roman_Δ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT. The influence of the low-power probe beam on the measurements is negligible. The measurements are repeated for a sweep of different pump input powers Pinsubscript𝑃inP_{\text{in}}italic_P start_POSTSUBSCRIPT in end_POSTSUBSCRIPT.

Refer to caption
Figure 3: a shows the normalized frequency noise spectra Sνν/Sννmaxsubscript𝑆𝜈𝜈subscript𝑆𝜈𝜈S_{\nu\nu}/S_{\nu\nu\max}italic_S start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT / italic_S start_POSTSUBSCRIPT italic_ν italic_ν roman_max end_POSTSUBSCRIPT against cavity detuning ΔΔ\Deltaroman_Δ for Pin=6.8 mWsubscript𝑃intimes6.8milliwattP_{\text{in}}=$6.8\text{\,}\mathrm{mW}$italic_P start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = start_ARG 6.8 end_ARG start_ARG times end_ARG start_ARG roman_mW end_ARG. The dashed black lines highlight the effect of the optomechanical spring on the mechanical resonance frequency ΩmsubscriptΩ𝑚\Omega_{m}roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT. The legend in the upper right indicates the corresponding pump powers and maximum (on-resonance) cavity photon numbers for b-d. In b the pump-reflection signal PPDpumpsubscriptsubscript𝑃PDpump{P_{\text{PD}}}_{\text{pump}}italic_P start_POSTSUBSCRIPT PD end_POSTSUBSCRIPT start_POSTSUBSCRIPT pump end_POSTSUBSCRIPT is plotted against the cavity detuning ΔΔ\Deltaroman_Δ for different input powers Pinsubscript𝑃inP_{\text{in}}italic_P start_POSTSUBSCRIPT in end_POSTSUBSCRIPT. In c the corresponding mechanical resonance frequency shift ΔΩmΔsubscriptΩ𝑚\Delta\Omega_{m}roman_Δ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT induced by the optomechanical spring is shown. d shows the normalized pump-reflection signal PPDpumpsubscriptsubscript𝑃PDpump{P_{\text{PD}}}_{\text{pump}}italic_P start_POSTSUBSCRIPT PD end_POSTSUBSCRIPT start_POSTSUBSCRIPT pump end_POSTSUBSCRIPT and mechanical resonance frequency shift ΔΩmΔsubscriptΩ𝑚\Delta\Omega_{m}roman_Δ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT induced by the optomechanical spring against cavity detuning ΔΔ\Deltaroman_Δ for input power Pin=6.8 mWsubscript𝑃intimes6.8milliwattP_{\text{in}}=$6.8\text{\,}\mathrm{mW}$italic_P start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = start_ARG 6.8 end_ARG start_ARG times end_ARG start_ARG roman_mW end_ARG. The dashed lines correspond to the unstable or not scanned solutions of Eq. 3.

Fig. 3 a shows a typical measurement of 𝒮νν(f)subscript𝒮𝜈𝜈𝑓\mathcal{S}_{\nu\nu}(f)caligraphic_S start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT ( italic_f ) of the mechanical polymer drum resonance against the cavity detuning ΔΔ\Deltaroman_Δ for a comparably large pump input power of Pin=6.8 mWsubscript𝑃intimes6.8milliwattP_{\text{in}}=$6.8\text{\,}\mathrm{mW}$italic_P start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = start_ARG 6.8 end_ARG start_ARG times end_ARG start_ARG roman_mW end_ARG. At such a power level an additional photothermal nonlinearity of the optical cavity resonance shifts the resonance position during the detuning scan [45] leading to a discontinuous behavior (black dashed cross) of the signal. The photothermal nonlinearity of the optical mode is also observed in the reflection signal of the pump-laser shown in Fig. 3 b. An optomechanical bistability as the origin of this behavior can be excluded as our coupling rate g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and ncavsubscript𝑛cavn_{\text{cav}}italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT are too small for this effect [1]. The nonlinearity is likely related to absorption on the polymer surfaces as stronger nonlinearites are observed in geometries with intensity maxima at the polymer-air interface.

As the photothermal absorption is slow compared to the optomechanically induced dynamics, we include the shift and bistability of the optical resonance in our analysis by treating it as an additional static detuning of the cavity. The cavity detuning ΔΔ\Deltaroman_Δ is therefore modified to Δ=ΔαncavsuperscriptΔΔ𝛼subscript𝑛cav\Delta^{\prime}=\Delta-\alpha n_{\text{cav}}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = roman_Δ - italic_α italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT with α𝛼\alphaitalic_α being the photothermal frequency-pull factor. To find ncavsubscript𝑛cavn_{\text{cav}}italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT for a given detuning, we insert ΔsuperscriptΔ\Delta^{\prime}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in the steady-state solution of the equation-of-motion of the complex light amplitude a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG as obtained from input-output formalism [46]

κexa^in=(iΔκ/2)a^,subscript𝜅exsubscript^𝑎in𝑖superscriptΔ𝜅2^𝑎-\sqrt{\kappa_{\text{ex}}}\hat{a}_{\text{in}}=\left(i\Delta^{\prime}-\kappa/2% \right)\hat{a},- square-root start_ARG italic_κ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = ( italic_i roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_κ / 2 ) over^ start_ARG italic_a end_ARG , (2)

with cavity input coupling κexsubscript𝜅ex\kappa_{\text{ex}}italic_κ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT and complex input field amplitude a^insubscript^𝑎in\hat{a}_{\text{in}}over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT in end_POSTSUBSCRIPT. Squaring and averaging Eq. 2 on both sides leads to a third-order polynomial equation in the cavity photon number ncavsubscript𝑛cavn_{\text{cav}}italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT:

0=α2ncav 32Δαncav+2(Δ2+κ2/4)ncav κexnin 0=\alpha^{2}n_{\text{cav }}^{3}-2\Delta\cdot\alpha n_{\operatorname{cav}}{}^{2% }+\left(\Delta^{2}+\kappa^{2}/4\right)n_{\text{cav }}-\kappa_{\text{ex}}n_{% \text{in }}0 = italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 2 roman_Δ ⋅ italic_α italic_n start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT + ( roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 ) italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT in end_POSTSUBSCRIPT (3)

with cavity photon number ncav=a^a^subscript𝑛cavdelimited-⟨⟩superscript^𝑎^𝑎n_{\text{cav}}=\langle\hat{a}^{\dagger}\hat{a}\rangleitalic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT = ⟨ over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_a end_ARG ⟩ and input photon number rate nin=a^ina^insubscript𝑛indelimited-⟨⟩superscriptsubscript^𝑎insubscript^𝑎inn_{\text{in}}=\langle\hat{a}_{\text{in}}^{\dagger}\hat{a}_{\text{in}}\rangleitalic_n start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = ⟨ over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT in end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT in end_POSTSUBSCRIPT ⟩. The two stable solutions of ncavsubscript𝑛cavn_{\text{cav}}italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT correspond to the photon number expected from the two directions of the detuning scan, whilst the third, unstable solution is not reached [45]. As shown in Fig. 3 b & d (red curve), we start the scan from Δ<0Δ0\Delta<0roman_Δ < 0. As more and more photons are coupled into the cavity, the resonance shifts away from the approaching pump-laser following one of the stable branches for the cavity reflection signal. When the pump-scan catches up with the drifting resonance at the maximum amount of intracavity photons, any further detuning abruptly reduces the thermal frequency-drift of the resonance resulting in a sudden jump to the reflection signal baseline. Using the solutions of Eq. 3, the extracted cavity photon number ncav(Δ)subscript𝑛cavΔn_{\text{cav}}(\Delta)italic_n start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT ( roman_Δ ) can be inserted in Eq. 1 with ΔΔ\Deltaroman_Δ being replaced by ΔsuperscriptΔ\Delta^{\prime}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. This model is then used to fit the measured data displayed in Fig.  3 c & d (blue curve). The unstable solution and the not-scanned branch of the reflection signal are included as dashed lines tracing a loop shape. Towards low pump-powers, the transition to a normal dispersive lineshape of the optomechanical spring effect can be observed. The maximum optomechanical frequency shift measured here was ΔΩm/2π=31 kHzΔsubscriptΩ𝑚2𝜋times31kilohertz\Delta\Omega_{m}/2\pi=$31\text{\,}\mathrm{kHz}$roman_Δ roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / 2 italic_π = start_ARG 31 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG for the maximum pump-power that we reached of Pin=6.8 mWsubscript𝑃intimes6.8milliwattP_{\text{in}}=$6.8\text{\,}\mathrm{mW}$italic_P start_POSTSUBSCRIPT in end_POSTSUBSCRIPT = start_ARG 6.8 end_ARG start_ARG times end_ARG start_ARG roman_mW end_ARG. The cavity finesse in this case was =1400±300plus-or-minus1400300\mathcal{F}=1400\pm 300caligraphic_F = 1400 ± 300 and the photothermal frequency-pull factor α/2π=8±2 kHz𝛼2𝜋timesuncertain82kilohertz\alpha/2\pi=$8\pm 2\text{\,}\mathrm{kHz}$italic_α / 2 italic_π = start_ARG start_ARG 8 end_ARG ± start_ARG 2 end_ARG end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG. Improved cavity finesse by stronger surface polishing, and higher pump-powers will allow us to increase this shift. Currently it already surpasses the mechanical linewidth at cryogenic temperatures. As the shift is on the order of the current frequency disorder of printed polymer drums it could be used to dynamically tune single drums in a multi-mode system into and out of collective resonances.

Mechanical resonance linewidth

Under ambient atmospheric conditions, the mechanical quality factor of the polymer membrane oscillators is limited to 20similar-toabsent20\sim 20∼ 20 by about equal parts through dam** of the membrane motion by surrounding gas and internal losses of the polymer. In comparison, radiation of mechanical energy into the substrate is negligible due to the impedance mismatch of sound in the polymer and the glass material below. Internal losses of the polymer are characterized by the loss tangent – the tangent of the phase between imaginary and real part of the dynamic modulus – of the material and vary strongly between different DLW resists [47]. In addition, the mechanical properties of polymers can exhibit a rich temperature dependence [48] with possible secondary glass transitions due to conformation changes of polymer chains.

Refer to caption
Figure 4: Temperature dependence of the mechanical resonator properties. The optomechanical cavity geometry for the experimental setup inside a liquid Helium flow cryostat uses an FFPC with two fiber mirrors (FM 11\,11 and FM 22\,22) inside a slotted glass ferrule that is tunable by an attached piezo [36]. The polymer drum is directly fabricated on one of the fiber mirrors as shown in a. Upon cool-down the polymer stiffens causing the resonance frequency to increase as shown in b. In c, the mechanical linewidth at different sensor temperatures is shown. For some exemplary measurements the quality factors and measured spectra are included.

To investigate the temperature dependence of the mechanical properties, we place a ferrule-based, passively stable FFPC configuration (for details, see [36]) in a liquid Helium continuous flow cryostat. The polymer membrane structure is fabricated on a highly reflective fiber mirror (see Fig. 4 a). The sample holder, on which the FFPC is mounted, sits in an evacuated chamber and can be cooled from room temperature down to 4 Ksimilar-toabsenttimes4kelvin\sim$4\text{\,}\mathrm{K}$∼ start_ARG 4 end_ARG start_ARG times end_ARG start_ARG roman_K end_ARG. The sensor temperature is, however, only a lower bound to the local temperature of the membrane. When cooling down the membrane resonator, the polymer material stiffness and pre-strain increase, which leads to an increase of the mechanical resonance frequency from about 3 MHztimes3megahertz3\text{\,}\mathrm{MHz}start_ARG 3 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG to 3.5 MHzsimilar-toabsenttimes3.5megahertz\sim$3.5\text{\,}\mathrm{MHz}$∼ start_ARG 3.5 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG (see Fig. 4 b). At the same time the mechanical linewidth drops from about 150 kHztimes150kilohertz150\text{\,}\mathrm{kHz}start_ARG 150 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG down to 6 kHzsimilar-toabsenttimes6kilohertz\sim$6\text{\,}\mathrm{kHz}$∼ start_ARG 6 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG corresponding to a mechanical quality factor of 600similar-toabsent600\sim 600∼ 600 as shown in Fig. 4 c. Kinks in the temperature dependence of the mechanical linewidth may indicate possible secondary glass transitions [48].

The remaining loss, decoherence or linewidth broading of the polymer membrane resonators at cryogenic temperature can be caused by various mechanisms. A possible candidate is a mechanical coupling to the conformation changes of polymer molecule chains. Also scattering of high energy phonons can appear, which would be suppressed at even lower temperatures. Due to the comparably large thickness of the membrane resonators, thermo-elastic dissipation [49, 50] is expected to play another major role. Engineering the mechanical mode to reduce strain gradients and further polishing using oxygen plasma ashing to thin down the membrane will help to reduce these loss channels and can be combined with isolation and soft-clam** techniques to further increase the mechanical quality factor [38, 39].

Pre-straining, together with a thinned membrane, will further help to dilute dissipation and improve the mechanical quality factor. An advantageous resource for pre-straining the material will be the usual shrinkage of the polymeric resist during polymerization. The comparably large tolerance of polymers to straining that is on par or even exceeding the limit strain of conventional materials used for micro-mechanical oscillators can enable comparably large dilution factors partly compensating the lower intrinsic mechanical quality factor. According to the manufacturer, the employed IP-S resist can show shrinkage between 2-12 %times12percent12\text{\,}\mathrm{\char 37}start_ARG 12 end_ARG start_ARG times end_ARG start_ARG % end_ARG under tuned fabrication conditions that were not explored here (other sintered glass-composite resists even up to 26.7 %times26.7percent26.7\text{\,}\mathrm{\char 37}start_ARG 26.7 end_ARG start_ARG times end_ARG start_ARG % end_ARG). Due to the anchoring of the structure the shrinkage in the development directly translates into strain. Assuming 10 %times10percent10\text{\,}\mathrm{\char 37}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG % end_ARG strain, a dilution factor of 280similar-toabsent280\sim 280∼ 280 would be reached for a 100 µmtimes100micrometer100\text{\,}\mathrm{\SIUnitSymbolMicro m}start_ARG 100 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG long, λ/4𝜆4\lambda/4italic_λ / 4–thick (195 nmtimes195nanometer195\text{\,}\mathrm{nm}start_ARG 195 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG) structure [39, 51, 52, 53]. For the cryogenic intrinsic quality factor this would result in a Qmech>1.5×105 subscript𝑄mechtimes1.5E5absentQ_{\text{mech}}>$1.5\text{\times}{10}^{5}\text{\,}$italic_Q start_POSTSUBSCRIPT mech end_POSTSUBSCRIPT > start_ARG start_ARG 1.5 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG 5 end_ARG end_ARG end_ARG start_ARG times end_ARG start_ARG end_ARG even without further engineering of the resonators. Other resists with higher intrinsic quality and even stronger shrinkage can potentially boost this prospect by another order of magnitude.

The exploration of the material properties of other 3D DLW resists under cryogenic conditions will also help to further identify low-loss materials. Moreover, 3D DLW structures can be used as blanks for subsequent material deposition, where the resist blank is later removed by oxygen plasma ashing or suit as stamp frames (e.g. from PDMS) in combination with other membranes or 2D materials like transition metal dichalcogenides.

III Discussion

We have demonstrated an FFPC-integrated optomechanical membrane-in-the-middle experiment with 3D direct laser-written membrane structures. Despite the relatively low membrane reflectivity, large optomechanical couplings of g030 kHzsimilar-tosubscript𝑔0times30kilohertzg_{0}\sim$30\text{\,}\mathrm{kHz}$italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG are realized since the comparatively large membrane thickness [3, 21, 22, 26, 27] of several λ/4𝜆4\lambda/4italic_λ / 4 allows us to maximize the intensity differences on the two membrane-air interfaces. Reducing the effective membrane mass by thinning the membrane thickness down to the order of a single λ/4𝜆4\lambda/4italic_λ / 4 thickness will further enhance the optomechanical coupling.

The unconventional polymer material stemming from the 3D DLW fabrication process results in moderate intrinsic mechanical quality factors at ambient conditions. However, a significant gain in the mechanical Q-factor is realized at cryogenic temperatures. Furthermore, the large range of specialized 3D DLW materials [47], the possible implementation of dissipation dilution techniques [39], the use of 3D DLW structures as blanks for material growth, and their capabilities as stam** frames will allow a fast advancement of the presented platform. In turn, optomechanical low-temperature experiments enable highly sensitive material characterization of DLW materials, which will contribute to the current rapid development of specialized resists.

Refer to caption
Figure 5: a shows the optical noise spectrum of a cavity with two mechanical membrane modes and their corresponding simulated displacement fields. The modes in this example structure are deliberately detuned by tilting the supports and thereby the size of the corresponding membrane. b Showcase example of multi-mode mechanical structures realized using direct laser-writing showing stacks of multiple, freely suspended drums (here 2 and 4) that are directly fabricated on the tip of an optical fiber.

The large flexibility of the 3D DLW fabrication allows the combination of this highly integrated platform with additional structures like electrodes [18] for electromechanical coupling, or emitters, and makes it scalable towards multi-membrane experiments. These can be both realized as membrane-stacks in a single FFPC that are considered to lead to improved values of the optomechanical coupling [54, 37], or as a planar 2D mechanical metamaterial on a DBR substrate. First experiments, where we introduce stacks of two mechanical membranes in a single optical cavity, see Fig. 5, show the feasibility of this approach. In contrast to other platforms [55, 56] the number of layers in such stacks can easily be extended beyond two and the geometry of each membrane separately adjusted to bring multiple membranes in tune. Using DLW, mechanical device layers can also be added and combined with a plethora of other platforms for light-matter interaction. Acoustic metamaterials in such device layers would benefit from the large optomechanical spring effects allowing for tunable mechanics and optical reconfiguration of mechanical multi-mode circuits with much less elaborate tuning techniques than required in other optomechanical platforms [57]. This will enable vibration-routing in 2D metamaterials, distributed sensing in multi-mode mechanical structures, and fiber-tip-integrated sensing of motion and force.

Methods

Sample fabrication

The polymer drum structures were fabricated using a commercial 3D lithography system (Nanoscribe Photonic Professional GT+, Nanoscribe GmbH & Co. KG, Germany) that uses two-photon-polymerization. A 63×63\times63 × objective was used in combination with the photoresist IP-S in dip-in configuration to print on the end-facet of the single-mode fiber or DBR substrates using the system’s piezo-mode. The system features a femtosecond laser centered at 780 nm that patterns and polymerizes the resist at the objective’s focus. The hatching distance for the feet and frame were chosen comparably coarse, while the membrane was patterned fine (65 nmtimes65nanometer65\text{\,}\mathrm{nm}start_ARG 65 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG) using a laserpower of 22%percent2222\,\%22 % (corresponding to 11 mWtimes11milliwatt11\text{\,}\mathrm{mW}start_ARG 11 end_ARG start_ARG times end_ARG start_ARG roman_mW end_ARG) and a scanspeed of 50 µm/stimes50µms50\text{\,}\mathrm{\SIUnitSymbolMicro}\mathrm{m}\mathrm{/}\mathrm{s}start_ARG 50 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m / roman_s end_ARG. The unpolymerized resist part was removed via immersion in propylene glycol methyl ether acetate (PGMEA) (30 mintimes30min30\text{\,}\mathrm{m}\mathrm{i}\mathrm{n}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_min end_ARG) and a subsequent bath in isopropyl alcohol (30 mintimes30min30\text{\,}\mathrm{m}\mathrm{i}\mathrm{n}start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_min end_ARG). In a post-development step, the structure was flood exposed with a 6 Wtimes6watt6\text{\,}\mathrm{W}start_ARG 6 end_ARG start_ARG times end_ARG start_ARG roman_W end_ARG UV-lamp while sitting in isopropyl alcohol (10 mintimes10min10\text{\,}\mathrm{m}\mathrm{i}\mathrm{n}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_min end_ARG). Whilst this procedure can minimize structure deformation during the development, it also only results in minimal shrinkage of the resist. Under certain fabrication conditions IP-S is specified to exhibit 2-12 %times12percent12\text{\,}\mathrm{\char 37}start_ARG 12 end_ARG start_ARG times end_ARG start_ARG % end_ARG shrinkage. For comparison, the silica-compound based GP-Silica resist can even show up to 26.7 %times26.7percent26.7\text{\,}\mathrm{\char 37}start_ARG 26.7 end_ARG start_ARG times end_ARG start_ARG % end_ARG shrinkage in the sintering step alone. Further post-processing of the membrane was performed using oxygen plasma-ashing for surface polishing to reduce scattering losses from the interfaces. The polymer structures were fabricated on the DBR substrate or fiber mirror that constitute the higher reflective cavity mirror (10 ppmtimes10ppm10\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG transmission). To reduce the complexity of the analysis and to focus on the fundamental material properties of the resist, the comparably simple rectangular membrane design is presented in this article. However, using this process, several different geometries were realized and much more complex structures including for example membrane stacks as shown in Fig. 5 can be fabricated. The thinnest membrane size in such structures can be reduced below the voxel height hvsubscriptvh_{\text{v}}italic_h start_POSTSUBSCRIPT v end_POSTSUBSCRIPT (smallest voxel: (\diameterv,hv)(0.2,0.7) µmsimilar-tosubscript\diametervsubscriptv0.20.7timesabsentmicrometer(\diameter_{\text{v}},h_{\text{v}})\sim(0.2,0.7)\,$\text{\,}\mathrm{% \SIUnitSymbolMicro m}$( start_POSTSUBSCRIPT v end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT v end_POSTSUBSCRIPT ) ∼ ( 0.2 , 0.7 ) start_ARG end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG) by oxygen plasma-ashing. The accuracy of the placement of the membranes is given by the piezo stage positioning accuracy (10 nmtimes10nanometer10\text{\,}\mathrm{nm}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG in the full 300 µm×300 µmtimes300micrometertimes300micrometer$300\text{\,}\mathrm{\SIUnitSymbolMicro m}$\times$300\text{\,}\mathrm{% \SIUnitSymbolMicro m}$start_ARG 300 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG × start_ARG 300 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG writing field), which is considerably smaller than λ/4𝜆4\lambda/4italic_λ / 4 as required for the membrane positioning. Possible tilts αtiltsubscript𝛼tilt\alpha_{\text{tilt}}italic_α start_POSTSUBSCRIPT tilt end_POSTSUBSCRIPT of the structure are corrected by the automatic interface finder, but would also only play a role for tilts approaching αtilt(λ/4)/Lmemsimilar-tosubscript𝛼tilt𝜆4subscript𝐿mem\alpha_{\text{tilt}}\sim(\lambda/4)/L_{\text{mem}}italic_α start_POSTSUBSCRIPT tilt end_POSTSUBSCRIPT ∼ ( italic_λ / 4 ) / italic_L start_POSTSUBSCRIPT mem end_POSTSUBSCRIPT with Lmemsubscript𝐿memL_{\text{mem}}italic_L start_POSTSUBSCRIPT mem end_POSTSUBSCRIPT the length of the membranes. Detailed characterizations of such multi-membrane devices will be presented in future work.

Experimental setup

Refer to caption
Figure 6: Overview of the experiment setup. The green highlighted pump laser is only used within the optomechanical spring measurements. The optical cavity with the mechanical membrane resonator (gray box) is either located in ambient conditions (single fiber mirror and DBR substrate) or situated on the sample holder of a liquid Helium flow cryostat (two fiber mirrors in a glass ferrule). The optical cavity length can either be scanned or locked to the probe laser (highlighted in red) using either a Pound-Drever-Hall-type (highlighted in blue) or a side-of-fringe lock. For this purpose and as a frequency-meter in case of a scanned cavity, an electro-optic modulator is used to create sidebands to the main probe laser tone at adjustable RF-frequency up to 12 GHztimes12gigahertz12\text{\,}\mathrm{GHz}start_ARG 12 end_ARG start_ARG times end_ARG start_ARG roman_GHz end_ARG. The photodiode reflection signals and the electrical noise spectrum of the lock can be measured using conventional oscilloscopes and an electric spectrum analyzer.

The full experiment setup is sketched in Fig. 6. The optical mode is characterized by measuring the reflected optical power of a fixed wavelength probe laser (probe setup red highlighted in Fig. 6, see also [40, 36]) on a photodiode (PDprobeprobe{}_{\text{probe}}start_FLOATSUBSCRIPT probe end_FLOATSUBSCRIPT), while the cavity length is scanned using a piezo-electric element that is glued to the optical fiber. The high-voltage drive of the piezo is generated by a 100×100\times100 × voltage amplifier and a subsequent low pass filter (LPF) to reject high-frequency electric noise. The alignment of the fiber during the scan process is secured by a glass ferrule (131 µmtimes131micrometer131\text{\,}\mathrm{\SIUnitSymbolMicro m}start_ARG 131 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG bore diameter), in which the optical fiber can slide. The incoming and reflection signal are split on a polarizing beam splitter (PBS). The subsequent waveplates (λ/2𝜆2\lambda/2italic_λ / 2 and λ/4𝜆4\lambda/4italic_λ / 4) are used to adjust the polarization of the reflected light to be perpendicular to the polarization of the incoming laser beam causing the PBS to direct most of the reflection onto PDprobeprobe{}_{\text{probe}}start_FLOATSUBSCRIPT probe end_FLOATSUBSCRIPT. An electro-optic modulator (EOM) adds sidebands to the main probe laser tone. The reflection signal measured with the oscilloscope therefore features three reflection dips with a spacing of the RF-drive of the EOM that is thereby used as a frequency-meter.

To characterize the optomechanical coupling strength the optical cavity is locked to the probe laser tone using a Pound-Drever-Hall (PDH, blue highlighted) or side-of-fringe (SoF) lock. The slope of the lock signal is calibrated using the frequency-meter enabled by the EOM-sidebands. Using this, the measured voltage noise of PDprobeprobe{}_{\text{probe}}start_FLOATSUBSCRIPT probe end_FLOATSUBSCRIPT can be converted to the optical cavity frequency noise. As the bandwidth of the feedback loop is small (1 kHzsimilar-toabsenttimes1kilohertz\sim$1\text{\,}\mathrm{kHz}$∼ start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG) compared to the mechanical resonance frequency (1 MHzgreater-than-or-equivalent-toabsenttimes1megahertz\gtrsim$1\text{\,}\mathrm{MHz}$≳ start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG) a bias tee is used to split the low from the high-frequency component. The DC-like part is used in the feedback loop, whilst the high-frequency part is analyzed using an electric spectrum analyzer (ESA). Mechanical mode frequency and linewidth can directly be extracted from there. The vacuum optomechanical coupling strength is then inferred by comparing the optical cavity frequency noise with the expected thermal noise of a mechanical oscillator of the measured frequency and linewidth at ambient temperature conditions [43].

In order to measure the dynamic optomechanical spring effect a second pump laser is used (highlighted in green), while the optical cavity remains tightly locked to the probe laser. The pump laser addresses a second optical cavity mode separated by one free spectral range towards higher frequencies from the probe-cavity-resonance. For variable pump laser powers, the pump frequency is scanned. The reflection signal of the pump is recorded on a second photodiode (PDpumppump{}_{\text{pump}}start_FLOATSUBSCRIPT pump end_FLOATSUBSCRIPT), while the locked probe laser error signal is used for the characterization of the mechanical mode frequency.

Optical properties of the membrane-cavity system

The FFPC designs used in this work are based on either a hemi-cavity design with a fiber mirror that can scan structures on a flat, highly reflective substrate inspired by [58], or a passively stable, monolithic FFPC realization with two fiber mirrors inside a glass ferrule as demonstrated in [36].

We use a standard \ceCO2\ce𝐶𝑂2\ce{CO2}italic_C italic_O 2 laser ablation system to machine spherical-like depressions onto the center of single-mode optical fibers [17]. The resulting concave indentation on our fiber end-facets features typical radii of curvature of 200 µmsimilar-toabsenttimes200micrometer\sim$200\text{\,}\mathrm{\SIUnitSymbolMicro m}$∼ start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG with usable spherical diameters of >40 µmabsenttimes40micrometer>$40\text{\,}\mathrm{\SIUnitSymbolMicro m}$> start_ARG 40 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG. The prepared fiber-tips are coated with alternating layers of \ceTa2O5\ce𝑇𝑎2𝑂5\ce{Ta2O5}italic_T italic_a 2 italic_O 5 and \ceSiO2\ce𝑆𝑖𝑂2\ce{SiO2}italic_S italic_i italic_O 2, resulting in highly reflective fiber end-facets for wavelengths ranging from 750800 nm750times800nanometer750-$800\text{\,}\mathrm{nm}$750 - start_ARG 800 end_ARG start_ARG times end_ARG start_ARG roman_nm end_ARG at AOI=0 °AOItimes0degree\text{AOI}=$0\text{\,}\mathrm{\SIUnitSymbolDegree}$AOI = start_ARG 0 end_ARG start_ARG times end_ARG start_ARG ° end_ARG. These fiber mirrors make up the cavity geometry and are coated for high transmission (2000 ppmtimes2000ppm2000\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}start_ARG 2000 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG) to be used as in-coupling fiber mirrors to the cavity and low transmission (10 ppmtimes10ppm10\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG) blank mirrors for interfacing the direct laser-written membranes. The empty cavity finesse of 2800similar-toabsent2800\sim 2800∼ 2800 (no polymer membrane inside cavity volume) is primarily determined by the transmission losses of the cavity mirrors.

Refer to caption
Figure 7: Effect of oxygen plasma polishing on the amount of surface scattering from a DLW-fabricated thin film. The cavity losses are determined by measuring the minimal optical finesse of a polymer block structure with a height variation placed directly on a mirror substrate. Evaluating at the thickness of minimal finesse thereby ensures an intensity maximum of the intra-cavity field at the scattering polymer interface. The reduction of losses can therefore be attributed to the surface losses only, whilst their absolute value contains additional losses from the imperfections in the large bulk block below. In each step the same polymer structure underwent subsequent plasma steps of duration tpolishsubscript𝑡polisht_{\text{polish}}italic_t start_POSTSUBSCRIPT polish end_POSTSUBSCRIPT (minutes) at plasma power Pplasmasubscript𝑃plasmaP_{\text{plasma}}italic_P start_POSTSUBSCRIPT plasma end_POSTSUBSCRIPT (%) yielding the horizontal axis sets (tpolish,Pplasma)subscript𝑡polishsubscript𝑃plasma(t_{\text{polish}},P_{\text{plasma}})( italic_t start_POSTSUBSCRIPT polish end_POSTSUBSCRIPT , italic_P start_POSTSUBSCRIPT plasma end_POSTSUBSCRIPT ).

The optical losses induced by the polymer membrane at lower input laser powers (<500 µWabsenttimes500microwatt<$500\text{\,}\mathrm{\SIUnitSymbolMicro W}$< start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_W end_ARG) are dominated by surface scattering effects at the polymer-air interfaces. Depending on the cavity field distribution, optical scattering losses induced by the polymer membrane can range from 5000 ppmsimilar-toabsenttimes5000ppm\sim$5000\text{\,}\mathrm{p}\mathrm{p}\mathrm{m}$∼ start_ARG 5000 end_ARG start_ARG times end_ARG start_ARG roman_ppm end_ARG with cavity intensity maxima on both polymer-air interfaces to almost fully recovering the empty cavity optical quality in the case of intensity minima on both interfaces. The scattering from the polymer surfaces can be strongly reduced by polishing through oxygen plasma555Plasma cleaner system: Zepto-BR-200-PCCE, which smoothens the surface as shown in Fig. 7. As the plasma removes material from the surface, this technique can also be used to reduce the thickness of the membrane. Another possibility to in the future enhance the optical properties is to combine the fabrication with an atomic layer deposition based uniform growth that can flatten the surface even further.

At higher input laser powers (>500 µWabsenttimes500microwatt>$500\text{\,}\mathrm{\SIUnitSymbolMicro W}$> start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_W end_ARG), additional photothermal nonlinearities of the optical cavity resonance become relevant. Since we observe stronger nonlinearities in cavity geometries with intensity maxima at the polymer-air interfaces, these effects can most likely be attributed to surface absorption as for example induced by not-passivated bonds of the polymer. The additional absorption will then lead to heating and mechanical expansion of the membrane, causing the additional static photothermal detuning as observed in the experiments.

Coupling calculations

To compute the expected optomechanical frequency-pull factor G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) for a particular drum thickness and support height, we use two methods as described in section II. For both methods, we first numerically find the set of {ldrum,tdrum,hdrum}subscript𝑙drumsubscript𝑡drumsubscriptdrum\{l_{\text{drum}},t_{\text{drum}},h_{\text{drum}}\}{ italic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT } that fulfils the resonance condition of the optical cavity with the probe light at λ=2π/k0𝜆2𝜋subscript𝑘0\lambda={2\pi}/{k_{0}}italic_λ = 2 italic_π / italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Here, hdrumsubscriptdrumh_{\text{drum}}italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT is the polymer drum support height, tdrumsubscript𝑡drumt_{\text{drum}}italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT the polymer membrane thickness, and ldrum=Lcavtdrumhdrumsubscript𝑙drumsubscript𝐿cavsubscript𝑡drumsubscriptdruml_{\text{drum}}=L_{\text{cav}}-t_{\text{drum}}-h_{\text{drum}}italic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT - italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT, where, Lcavsubscript𝐿cavL_{\text{cav}}italic_L start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT denotes the separation between the two cavity mirrors (for an overview see Fig. 1 b). For each pair {tdrum,hdrum}subscript𝑡drumsubscriptdrum\{t_{\text{drum}},h_{\text{drum}}\}{ italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT }, a ldrumsubscript𝑙druml_{\text{drum}}italic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT at Lcav30 µmsimilar-tosubscript𝐿cavtimes30micrometerL_{\text{cav}}\sim$30\text{\,}\mathrm{\SIUnitSymbolMicro m}$italic_L start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT ∼ start_ARG 30 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_m end_ARG can be chosen to match the resonance condition R(ldrum,tdrum,hdrum)=0𝑅subscript𝑙drumsubscript𝑡drumsubscriptdrum0R(l_{\text{drum}},t_{\text{drum}},h_{\text{drum}})=0italic_R ( italic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) = 0, with (suppressing subscripts):

R(l,h,t)=A+(l)einpolyk0tA(h)einpolyk0t𝑅𝑙𝑡subscript𝐴𝑙superscript𝑒𝑖subscript𝑛polysubscript𝑘0𝑡subscript𝐴superscript𝑒𝑖subscript𝑛polysubscript𝑘0𝑡R\left(l,h,t\right)=A_{+}\left(l\right)\cdot e^{-in_{\text{poly}}k_{0}t}-A_{-}% \left(h\right)\cdot e^{in_{\text{poly}}k_{0}t}italic_R ( italic_l , italic_h , italic_t ) = italic_A start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( italic_l ) ⋅ italic_e start_POSTSUPERSCRIPT - italic_i italic_n start_POSTSUBSCRIPT poly end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT - italic_A start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( italic_h ) ⋅ italic_e start_POSTSUPERSCRIPT italic_i italic_n start_POSTSUBSCRIPT poly end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT (4)

and with coefficients A±(z)subscript𝐴plus-or-minus𝑧A_{\pm}(z)italic_A start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_z ):

A±(z)=1inpolytan(k0z)1±inpolytan(k0z).subscript𝐴plus-or-minus𝑧minus-or-plus1𝑖subscript𝑛polysubscript𝑘0𝑧plus-or-minus1𝑖subscript𝑛polysubscript𝑘0𝑧A_{\pm}(z)=\frac{1\mp in_{\text{poly}}\cdot\tan\left(k_{0}z\right)}{1\pm in_{% \text{poly}}\cdot\tan\left(k_{0}z\right)}\;.italic_A start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_z ) = divide start_ARG 1 ∓ italic_i italic_n start_POSTSUBSCRIPT poly end_POSTSUBSCRIPT ⋅ roman_tan ( italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_z ) end_ARG start_ARG 1 ± italic_i italic_n start_POSTSUBSCRIPT poly end_POSTSUBSCRIPT ⋅ roman_tan ( italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_z ) end_ARG .

In the picture used here, the cavity consists of a loss-less dielectric membrane with refractive index npolysubscript𝑛polyn_{\text{poly}}italic_n start_POSTSUBSCRIPT poly end_POSTSUBSCRIPT. It is placed between two mirrors with perfectly conducting surfaces, enforcing field nodes at their positions (the penetration depth of the DBR will in reality lead to a slightly modified resonance cavity length). As the Rayleigh length zRLcavmuch-greater-thansubscript𝑧𝑅subscript𝐿cavz_{R}\gg L_{\text{cav}}italic_z start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ≫ italic_L start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT, the Gaussian-beam properties of the cavity field are neglected and a simple standing-wave ansatz is made. There, the tangential component of the electric field at both mirror surfaces needs to vanish and the tangential components of both the electric- and magnetic fields need to be continuous at the dielectric interfaces [59]. Applying these conditions to the cavity field, we numerically find the condition that fixes the {ldrum,tdrum,hdrum}subscript𝑙drumsubscript𝑡drumsubscriptdrum\{l_{\text{drum}},t_{\text{drum}},h_{\text{drum}}\}{ italic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT } - triplet. This is now used to extract the optomechanical frequency-pull factor G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ):

In Method 1, we use Eq. 4 by applying a small shift to the membrane, altering the cavity geometry to ldrumldrum+Δzsubscript𝑙drumsubscript𝑙drumΔ𝑧l_{\text{drum}}\rightarrow l_{\text{drum}}+\Delta zitalic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT → italic_l start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT + roman_Δ italic_z, hdrumhdrumΔzsubscriptdrumsubscriptdrumΔ𝑧h_{\text{drum}}\rightarrow h_{\text{drum}}-\Delta zitalic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT → italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT - roman_Δ italic_z and k0k0+Δksubscript𝑘0subscript𝑘0Δ𝑘k_{0}\rightarrow k_{0}+\Delta kitalic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT → italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_k. This simulates the effect of the linear displacement of the membrane center region on the cavity resonance condition666For this mathematical treatment, hdrumsubscriptdrumh_{\text{drum}}italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT does not directly correspond to the polymer drum support height, but more accurately to the distance between the displaced membrane at its center and the subjacent mirror. The mode width of the optical mode is sufficiently smaller than the effective membrane radius to ensure that the optical mode only overlaps with an almost constant displacement of the drum membrane .. Inserting the shifted geometry into Eq. 4 allows to numerically extract a ΔkΔ𝑘\Delta kroman_Δ italic_k that again fulfils Eq. 4 and thereby reflects the shift on the resonance frequency ΔωcavΔsubscript𝜔cav\Delta\omega_{\text{cav}}roman_Δ italic_ω start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT. The ratio between the frequency shift and the small displacement allows us to numerically determine the frequency-pull G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ). Aside from the linear frequency-pull factor, higher orders can also be numerically evaluated. E.g. for the quadratic pull-factor values of 200 GHz nm2greater-than-or-equivalent-toabsenttimes200timesgigahertznanometer2\gtrsim$200\text{\,}\mathrm{GHz}\text{\,}{\mathrm{nm}}^{-2}$≳ start_ARG 200 end_ARG start_ARG times end_ARG start_ARG start_ARG roman_GHz end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_nm end_ARG start_ARG - 2 end_ARG end_ARG end_ARG are expected.

For Method 2, we use the optical field that fulfills the resonant cavity condition and apply perturbation theory to find the shift of the optical resonance associated with a shift of the dielectric boundaries of our geometry as derived by Johnson et al. in [44]. Here, Maxwell’s equations are written as an eigenproblem of the electric cavity field777For convenience, the basis-independent representation of the electric field as ”Bra”- and ”Ket”-vectors is utilized, with inner product EE𝐄*𝐄dVinner-product𝐸superscript𝐸superscript𝐄superscript𝐄𝑑𝑉{\left\langle E\mid E^{\prime}\right\rangle\equiv\int\mathbf{E}^{*}\cdot% \mathbf{E}^{\prime}\mathop{dV}}⟨ italic_E ∣ italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⟩ ≡ ∫ bold_E start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ⋅ bold_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_BIGOP italic_d italic_V end_BIGOP. |Eket𝐸|E\rangle| italic_E ⟩ with eigenfrequency ωcavsubscript𝜔cav\omega_{\text{cav}}italic_ω start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT given by the well-known source-free wave equation with overall cavity permittivity ϵ(z)italic-ϵ𝑧\epsilon(z)italic_ϵ ( italic_z ):

2|E=(ωcav c)2ϵ(z)|E.superscript2ket𝐸superscriptsubscript𝜔cav 𝑐2italic-ϵ𝑧ket𝐸\nabla^{2}|E\rangle=\left(\frac{\omega_{\text{cav }}}{c}\right)^{2}\epsilon(z)% |E\rangle.∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_E ⟩ = ( divide start_ARG italic_ω start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT end_ARG start_ARG italic_c end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϵ ( italic_z ) | italic_E ⟩ . (5)

Due to the vibrational motion of the drum membrane, the effective dielectric permittivity of the cavity geometry experiences a local shift δϵ𝛿italic-ϵ\mathop{\delta\epsilon}italic_δ italic_ϵ due to a perturbative shift in position of the drum membrane δz𝛿𝑧\mathop{\delta z}italic_δ italic_z. We expand |Eket𝐸|E\rangle| italic_E ⟩, ωcavsubscript𝜔cav\omega_{\text{cav}}italic_ω start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT to first-order in dz𝑑𝑧\mathop{dz}italic_d italic_z. Plugging these expressions back into Eq. 5 and neglecting terms of 𝒪(dz2)𝒪superscript𝑑𝑧2\mathcal{O}\left(\mathop{dz}^{2}\right)caligraphic_O ( start_BIGOP italic_d italic_z end_BIGOP start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) leads to the first-order correction of the resonator frequency in differential form [44]:

G(1)=dωcav(1)dz=ωcav(0)2E(0)|dϵdz|E(0)E(0)|ϵ|E(0)superscript𝐺1𝑑superscriptsubscript𝜔cav1𝑑𝑧superscriptsubscript𝜔cav02quantum-operator-productsuperscript𝐸0𝑑italic-ϵ𝑑𝑧superscript𝐸0quantum-operator-productsuperscript𝐸0italic-ϵsuperscript𝐸0G^{(1)}=\frac{d\omega_{\mathrm{cav}}^{(1)}}{dz}=-\frac{\omega_{\mathrm{cav}}^{% (0)}}{2}\frac{\left\langle E^{(0)}\left|\frac{d\epsilon}{dz}\right|E^{(0)}% \right\rangle}{\left\langle E^{(0)}|\epsilon|E^{(0)}\right\rangle}italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = divide start_ARG italic_d italic_ω start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_z end_ARG = - divide start_ARG italic_ω start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG divide start_ARG ⟨ italic_E start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT | divide start_ARG italic_d italic_ϵ end_ARG start_ARG italic_d italic_z end_ARG | italic_E start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ⟩ end_ARG start_ARG ⟨ italic_E start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT | italic_ϵ | italic_E start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ⟩ end_ARG (6)

We then make use of the explicit parametrization of the cavity permittivity

ϵ(z)=ϵ2+Δϵ(Θ(zz0)Θ(z(z0+tdrum))),italic-ϵ𝑧subscriptitalic-ϵ2Δitalic-ϵΘ𝑧subscript𝑧0Θ𝑧subscript𝑧0subscript𝑡drum\epsilon(z)=\epsilon_{2}+\Delta\epsilon\big{(}\Theta(z-z_{0})-\Theta(z-(z_{0}+% t_{\text{drum}}))\big{)}\;,italic_ϵ ( italic_z ) = italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_Δ italic_ϵ ( roman_Θ ( italic_z - italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - roman_Θ ( italic_z - ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) ) ) ,

with Δϵ=ϵ1ϵ2Δitalic-ϵsubscriptitalic-ϵ1subscriptitalic-ϵ2\Delta\epsilon=\epsilon_{1}-\epsilon_{2}roman_Δ italic_ϵ = italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (ϵ1subscriptitalic-ϵ1\epsilon_{1}italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT: polymer, ϵ2subscriptitalic-ϵ2\epsilon_{2}italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT: air/vacuum) and the Heaviside step function Θ(z)Θ𝑧\Theta(z)roman_Θ ( italic_z ). The first drum surface SLsubscript𝑆LS_{\text{L}}italic_S start_POSTSUBSCRIPT L end_POSTSUBSCRIPT is located at z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and the second drum surface SRsubscript𝑆RS_{\text{R}}italic_S start_POSTSUBSCRIPT R end_POSTSUBSCRIPT at z0+tdrumsubscript𝑧0subscript𝑡drumz_{0}+t_{\text{drum}}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT following the coordinate conventions of Fig. 2 b.

Inserting this expression back into Eq. 6, we arrive at the explicit form of the optomechanical frequency-pull G(1)superscript𝐺1G^{(1)}italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT:

G(1)=dωcavdz=ωcav(0)2SRΔϵ|𝐄(0)|2dSSLΔϵ|𝐄(0)|2dSVϵ(z)|𝑬(0)|2dV,superscript𝐺1𝑑subscript𝜔cav𝑑𝑧superscriptsubscript𝜔cav02subscriptsubscript𝑆𝑅Δitalic-ϵsuperscriptsuperscriptsubscript𝐄02𝑑𝑆subscriptsubscript𝑆𝐿Δitalic-ϵsuperscriptsuperscriptsubscript𝐄02𝑑𝑆subscript𝑉italic-ϵ𝑧superscriptsuperscriptsubscript𝑬02𝑑𝑉G^{(1)}=\frac{\mathop{d\omega_{\mathrm{cav}}}}{\mathop{dz}}=\frac{\omega_{% \mathrm{cav}}^{(0)}}{2}\frac{\int_{S_{R}}\Delta\epsilon\left|\mathbf{E}_{\|}^{% (0)}\right|^{2}\mathop{dS}-\int_{S_{L}}\Delta\epsilon\left|\mathbf{E}_{\|}^{(0% )}\right|^{2}\mathop{dS}}{\int_{V}\epsilon(z)\left|\boldsymbol{E}_{\|}^{(0)}% \right|^{2}\mathop{dV}}\;,italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = divide start_ARG italic_d italic_ω start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_z end_ARG = divide start_ARG italic_ω start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG divide start_ARG ∫ start_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Δ italic_ϵ | bold_E start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_BIGOP italic_d italic_S end_BIGOP - ∫ start_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Δ italic_ϵ | bold_E start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_BIGOP italic_d italic_S end_BIGOP end_ARG start_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_ϵ ( italic_z ) | bold_italic_E start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_BIGOP italic_d italic_V end_BIGOP end_ARG , (7)

with the total resonator volume V𝑉Vitalic_V and parallel electric field component 𝐄(0)superscriptsubscript𝐄0\mathbf{E}_{\|}^{(0)}bold_E start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT (explicit form given by Eq. 4). As detailed in section II, Eq. 7 allows us to read-off the frequency-pull G(1)(tdrum,hdrum)superscript𝐺1subscript𝑡drumsubscriptdrumG^{(1)}(t_{\text{drum}},h_{\text{drum}})italic_G start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT drum end_POSTSUBSCRIPT ) for a specific cavity geometry by considering the cavity field distribution on both surfaces of the polymer drum. It also provides the physical interpretation of the coupling to be caused by differing field intensities on the membrane-air interfaces. The coupling landscape retrieved from method 1 and method 2 agree for all tested geometries. For our cavity geometry of interest, the results are shown in Fig. 2 a.

Finite element simulations

We use COMSOL Multiphysics® [60] to perform finite element simulations of the mechanical resonator structures. The simulated geometry consists of the polymer membrane (IP-S: Dynamic modulus E=\complexqty5.33+0.26i\giga𝐸\complexqty5.330.26𝑖\gigaE=\complexqty{5.33+0.26i}{\giga}italic_E = 5.33 + 0.26 italic_i, Poisson’s ratio ν=0.3𝜈0.3\nu=0.3italic_ν = 0.3, density ρ=1.15 kg dm3𝜌times1.15timeskilogramdecimeter3\rho=$1.15\text{\,}\mathrm{kg}\text{\,}{\mathrm{dm}}^{-3}$italic_ρ = start_ARG 1.15 end_ARG start_ARG times end_ARG start_ARG start_ARG roman_kg end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_dm end_ARG start_ARG - 3 end_ARG end_ARG end_ARG[47]) with a silica substrate below. An outer shell of the substrate is defined as a perfectly matched layer (PML) to implement radiation losses into the substrate (see Fig. 1 b). Whilst these radiation losses are not the dominating loss mechanism at room temperature, they are strongly dependent on the polymer geometry and can be reduced to limiting quality factors of 106much-greater-thanabsentsuperscript106\gg 10^{6}≫ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT for optimized support dimensions and with included isolation cuts. By reducing the internal polymer losses, e.g. through other resist materials, high quality factor mechanical resonators would be feasible. Aside from the quality factor and resonance frequency, we use the simulation to extract the effective mass of the fundamental flexural mode (on the order of 2.4 ngsimilar-toabsenttimes2.4nanogram\sim$2.4\text{\,}\mathrm{ng}$∼ start_ARG 2.4 end_ARG start_ARG times end_ARG start_ARG roman_ng end_ARG depending on the specific geometry) and the corresponding zero point motion (3.2 fmsimilar-toabsenttimes3.2femtometer\sim$3.2\text{\,}\mathrm{fm}$∼ start_ARG 3.2 end_ARG start_ARG times end_ARG start_ARG roman_fm end_ARG) [61].

References

  • Aspelmeyer et al. [2014a] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity optomechanics, Reviews of Modern Physics 86, 1391 (2014a).
  • Gröblacher et al. [2009] S. Gröblacher, K. Hammerer, M. R. Vanner, and M. Aspelmeyer, Observation of strong coupling between a micromechanical resonator and an optical cavity field, Nature 460, 724 (2009).
  • Thompson et al. [2008] J. Thompson, B. Zwickl, A. Jayich, F. Marquardt, S. Girvin, and J. Harris, Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane, Nature 452, 72 (2008).
  • Jayich et al. [2008] A. Jayich, J. Sankey, B. Zwickl, C. Yang, J. Thompson, S. Girvin, A. Clerk, F. Marquardt, and J. Harris, Dispersive optomechanics: a membrane inside a cavity, New Journal of Physics 10, 095008 (2008).
  • Schliesser et al. [2008] A. Schliesser, R. Rivière, G. Anetsberger, O. Arcizet, and T. J. Kippenberg, Resolved-sideband cooling of a micromechanical oscillator, Nature Physics 4, 415 (2008).
  • Eichenfield et al. [2009] M. Eichenfield, J. Chan, R. M. Camacho, K. J. Vahala, and O. Painter, Optomechanical crystals, Nature 462, 78 (2009).
  • Benz et al. [2016] F. Benz, M. K. Schmidt, A. Dreismann, R. Chikkaraddy, Y. Zhang, A. Demetriadou, C. Carnegie, H. Ohadi, B. De Nijs, R. Esteban, et al., Single-molecule optomechanics in “picocavities”, Science 354, 726 (2016).
  • Chan et al. [2011] J. Chan, T. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer, and O. Painter, Laser cooling of a nanomechanical oscillator into its quantum ground state, Nature 478, 89 (2011).
  • Teufel et al. [2011] J. D. Teufel, T. Donner, D. Li, J. W. Harlow, M. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W. Simmonds, Sideband cooling of micromechanical motion to the quantum ground state, Nature 475, 359 (2011).
  • Fiaschi et al. [2021] N. Fiaschi, B. Hensen, A. Wallucks, R. Benevides, J. Li, T. P. M. Alegre, and S. Gröblacher, Optomechanical quantum teleportation, Nature Photonics 15, 817 (2021).
  • Andrews et al. [2014] R. W. Andrews, R. W. Peterson, T. P. Purdy, K. Cicak, R. W. Simmonds, C. A. Regal, and K. W. Lehnert, Bidirectional and efficient conversion between microwave and optical light, Nature Physics 10, 321 (2014).
  • Han et al. [2021] X. Han, W. Fu, C.-L. Zou, L. Jiang, and H. X. Tang, Microwave-optical quantum frequency conversion, Optica 8, 1050 (2021).
  • Teufel et al. [2009] J. D. Teufel, T. Donner, M. Castellanos-Beltran, J. W. Harlow, and K. W. Lehnert, Nanomechanical motion measured with an imprecision below that at the standard quantum limit, Nature nanotechnology 4, 820 (2009).
  • Anetsberger et al. [2010] G. Anetsberger, E. Gavartin, O. Arcizet, Q. P. Unterreithmeier, E. M. Weig, M. L. Gorodetsky, J. P. Kotthaus, and T. J. Kippenberg, Measuring nanomechanical motion with an imprecision below the standard quantum limit, Physical Review A 82, 061804 (2010).
  • Mason et al. [2019] D. Mason, J. Chen, M. Rossi, Y. Tsaturyan, and A. Schliesser, Continuous force and displacement measurement below the standard quantum limit, Nature Physics 15, 745 (2019).
  • Zhang et al. [2012] M. Zhang, G. S. Wiederhecker, S. Manipatruni, A. Barnard, P. McEuen, and M. Lipson, Synchronization of micromechanical oscillators using light, Physical Review Letters 109, 233906 (2012).
  • Hunger et al. [2010] D. Hunger, T. Steinmetz, Y. Colombe, C. Deutsch, T. W. Hänsch, and J. Reichel, A fiber fabry–perot cavity with high finesse, New Journal of Physics 12, 065038 (2010).
  • Pfeifer et al. [2022] H. Pfeifer, L. Ratschbacher, J. Gallego, C. Saavedra, A. Faßbender, A. von Haaren, W. Alt, S. Hofferberth, M. Köhl, S. Linden, et al., Achievements and perspectives of optical fiber fabry–perot cavities, Applied Physics B 128, 1 (2022).
  • Brekenfeld et al. [2020] M. Brekenfeld, D. Niemietz, J. D. Christesen, and G. Rempe, A quantum network node with crossed optical fibre cavities, Nature Physics 16, 647 (2020).
  • Niemietz et al. [2021] D. Niemietz, P. Farrera, S. Langenfeld, and G. Rempe, Nondestructive detection of photonic qubits, Nature 591, 570 (2021).
  • Flowers-Jacobs et al. [2012] N. Flowers-Jacobs, S. Hoch, J. Sankey, A. Kashkanova, A. Jayich, C. Deutsch, J. Reichel, and J. Harris, Fiber-cavity-based optomechanical device, Applied Physics Letters 101, 221109 (2012).
  • Shkarin et al. [2014] A. Shkarin, N. Flowers-Jacobs, S. Hoch, A. Kashkanova, C. Deutsch, J. Reichel, and J. Harris, Optically mediated hybridization between two mechanical modes, Physical Review Letters 112, 013602 (2014).
  • Kashkanova et al. [2017a] A. Kashkanova, A. Shkarin, C. Brown, N. Flowers-Jacobs, L. Childress, S. Hoch, L. Hohmann, K. Ott, J. Reichel, and J. Harris, Superfluid brillouin optomechanics, Nature Physics 13, 74 (2017a).
  • Kashkanova et al. [2017b] A. Kashkanova, A. Shkarin, C. Brown, N. Flowers-Jacobs, L. Childress, S. Hoch, L. Hohmann, K. Ott, J. Reichel, and J. Harris, Optomechanics in superfluid helium coupled to a fiber-based cavity, Journal of Optics 19, 034001 (2017b).
  • Shkarin et al. [2019] A. Shkarin, A. Kashkanova, C. Brown, S. Garcia, K. Ott, J. Reichel, and J. Harris, Quantum optomechanics in a liquid, Physical Review Letters 122, 153601 (2019).
  • Fogliano et al. [2021] F. Fogliano, B. Besga, A. Reigue, P. Heringlake, L. M. de Lépinay, C. Vaneph, J. Reichel, B. Pigeau, and O. Arcizet, Map** the cavity optomechanical interaction with subwavelength-sized ultrasensitive nanomechanical force sensors, Physical Review X 11, 021009 (2021).
  • Rochau et al. [2021] F. Rochau, I. S. Arribas, A. Brieussel, S. Stapfner, D. Hunger, and E. M. Weig, Dynamical backaction in an ultrahigh-finesse fiber-based microcavity, Physical Review Applied 16, 014013 (2021).
  • Maruo et al. [1997] S. Maruo, O. Nakamura, and S. Kawata, Three-dimensional microfabrication with two-photon-absorbed photopolymerization, Optics letters 22, 132 (1997).
  • Gissibl et al. [2016] T. Gissibl, S. Thiele, A. Herkommer, and H. Giessen, Two-photon direct laser writing of ultracompact multi-lens objectives, Nature photonics 10, 554 (2016).
  • Jörg et al. [2020] C. Jörg, G. Queraltó, M. Kremer, G. Pelegrí, J. Schulz, A. Szameit, G. von Freymann, J. Mompart, and V. Ahufinger, Artificial gauge field switching using orbital angular momentum modes in optical waveguides, Light: Science & Applications 9, 1 (2020).
  • Williams [2020] J. C. Williams, Dynamic Micromechanical Fabry-Perot Cavity Sensors Fabricated by Multiphoton Absorption Onto Optical Fiber Tips, Thesis and Disserations (2020), 3164.
  • Frenzel et al. [2017] T. Frenzel, M. Kadic, and M. Wegener, Three-dimensional mechanical metamaterials with a twist, Science 358, 1072 (2017).
  • Li et al. [2020] J. Li, S. Thiele, B. C. Quirk, R. W. Kirk, J. W. Verjans, E. Akers, C. A. Bursill, S. J. Nicholls, A. M. Herkommer, H. Giessen, et al., Ultrathin monolithic 3D printed optical coherence tomography endoscopy for preclinical and clinical use, Light: Science & Applications 9, 1 (2020).
  • Thompson et al. [2018] A. J. Thompson, M. Power, and G.-Z. Yang, Micro-scale fiber-optic force sensor fabricated using direct laser writing and calibrated using machine learning, Optics Express 26, 14186 (2018).
  • Zou et al. [2021] M. Zou, C. Liao, S. Liu, C. Xiong, C. Zhao, J. Zhao, Z. Gan, Y. Chen, K. Yang, D. Liu, et al., Fiber-tip polymer clamped-beam probe for high-sensitivity nanoforce measurements, Light: Science & Applications 10, 1 (2021).
  • Saavedra et al. [2021] C. Saavedra, D. Pandey, W. Alt, H. Pfeifer, and D. Meschede, Tunable fiber fabry-perot cavities with high passive stability, Optics Express 29, 974 (2021).
  • Xuereb et al. [2012] A. Xuereb, C. Genes, and A. Dantan, Strong coupling and long-range collective interactions in optomechanical arrays, Physical Review Letters 109, 223601 (2012).
  • Tsaturyan et al. [2017] Y. Tsaturyan, A. Barg, E. S. Polzik, and A. Schliesser, Ultracoherent nanomechanical resonators via soft clam** and dissipation dilution, Nature nanotechnology 12, 776 (2017).
  • Fedorov et al. [2019] S. A. Fedorov, N. J. Engelsen, A. H. Ghadimi, M. J. Bereyhi, R. Schilling, D. J. Wilson, and T. J. Kippenberg, Generalized dissipation dilution in strained mechanical resonators, Physical Review B 99, 054107 (2019).
  • Gallego et al. [2016] J. Gallego, S. Ghosh, S. K. Alavi, W. Alt, M. Martinez-Dorantes, D. Meschede, and L. Ratschbacher, High-finesse fiber fabry–perot cavities: stabilization and mode matching analysis, Applied Physics B 122, 1 (2016).
  • Drever et al. [1983] R. Drever, J. L. Hall, F. Kowalski, J. Hough, G. Ford, A. Munley, and H. Ward, Laser phase and frequency stabilization using an optical resonator, Applied Physics B 31, 97 (1983).
  • Black [2001] E. D. Black, An introduction to Pound–Drever–Hall laser frequency stabilization, American journal of physics 69, 79 (2001).
  • Gorodetksy et al. [2010] M. Gorodetksy, A. Schliesser, G. Anetsberger, S. Deleglise, and T. J. Kippenberg, Determination of the vacuum optomechanical coupling rate using frequency noise calibration, Optics Express 18, 23236 (2010).
  • Johnson et al. [2002] S. G. Johnson, M. Ibanescu, M. A. Skorobogatiy, O. Weisberg, J. D. Joannopoulos, and Y. Fink, Perturbation theory for maxwell’s equations with shifting material boundaries, Phys. Rev. E 65, 066611 (2002).
  • Carmon et al. [2004] T. Carmon, L. Yang, and K. J. Vahala, Dynamical thermal behavior and thermal self-stability of microcavities, Opt. Express 12, 4742 (2004).
  • Gardiner and Collett [1985] C. W. Gardiner and M. J. Collett, Input and output in damped quantum systems: Quantum stochastic differential equations and the master equation, Physical Review A 31, 3761 (1985).
  • Nanoscribe GmbH & Co. KG [2021] Nanoscribe GmbH & Co. KG, IP Photoresins, https://www.nanoscribe.com/de/produkte/ip-photoresins/ (2021).
  • Hartwig [1995] G. Hartwig, Polymer properties at room and cryogenic temperatures (Springer Science & Business Media, 1995).
  • Zener [1938] C. Zener, Internal friction in solids ii. general theory of thermoelastic internal friction, Physical review 53, 90 (1938).
  • Serra et al. [2012] E. Serra, F. Cataliotti, F. Marin, F. Marino, A. Pontin, G. Prodi, and M. Bonaldi, Inhomogeneous mechanical losses in micro-oscillators with high reflectivity coating, Journal of Applied Physics 111, 113109 (2012).
  • Villanueva and Schmid [2014] L. G. Villanueva and S. Schmid, Evidence of surface loss as ubiquitous limiting dam** mechanism in sin micro-and nanomechanical resonators, Physical review letters 113, 227201 (2014).
  • González and Saulson [1994] G. I. González and P. R. Saulson, Brownian motion of a mass suspended by an anelastic wire, The Journal of the Acoustical Society of America 96, 207 (1994).
  • Manjeshwar et al. [2023] S. K. Manjeshwar, A. Ciers, F. Hellman, J. Bläsing, A. Strittmatter, and W. Wieczorek, High-q trampoline resonators from strained crystalline ingap for integrated free-space optomechanics, Nano Letters  (2023).
  • Bhattacharya and Meystre [2008] M. Bhattacharya and P. Meystre, Multiple membrane cavity optomechanics, Physical Review A 78, 041801 (2008).
  • Gaärtner et al. [2018] C. Gaärtner, J. P. Moura, W. Haaxman, R. A. Norte, and S. Gröblacher, Integrated optomechanical arrays of two high reflectivity sin membranes, Nano letters 18, 7171 (2018).
  • Manjeshwar [2023] S. K. Manjeshwar, Free-Space Cavity Optomechanical Systems on a Chip with III-V HeterostructuresPh.D. thesis, Chalmers Tekniska Hogskola (Sweden) (2023).
  • Pfeifer et al. [2016] H. Pfeifer, T. Paraïso, L. Zang, and O. Painter, Design of tunable GHz-frequency optomechanical crystal resonators, Optics Express 24, 11407 (2016).
  • Mader et al. [2015] M. Mader, J. Reichel, T. W. Hänsch, and D. Hunger, A scanning cavity microscope, Nature Communications 6, 1 (2015).
  • Ujihara [2010] K. Ujihara, Output Coupling in Optical Cavities and Lasers: A Quantum Theoretical Approach (Wiley-VCH, 2010).
  • [60] COMSOL AB, Stockholm, Sweden, COMSOL Multiphysics®v. 6.1, www.comsol.com.
  • Aspelmeyer et al. [2014b] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity optomechanics: nano-and micromechanical resonators interacting with light (Springer, 2014).

Acknowledgements

The authors would like to thank Prof. Dieter Meschede for supporting the first experiments on this project. The authors acknowledge funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy – Cluster of Excellence Matter and Light for Quantum Computing (ML4Q) EXC 2004/1 – 390534769 as well as funding from the Bundesministerium für Bildung und Forschung (BMBF, Federal Ministry for Education and Research) - project FaResQ. S.H. furthermore acknowledges funding from the European Union’s Horizon 2020 program under the ERC consolidator grant RYD-QNLO (Grant No. 771417).

Author contributions statement

H.P., L.T., A.F., and S.L. came up with the concept and planned the experiments. A.F. and L.T. performed the fabrication and L.T. conducted the optical measurements. L.T., A.F., and H.P. set up the numerical simulations and analyzed the data. All authors contributed to the writing of the manuscript.

Data availability

The data of this study is available from the corresponding author (H.P.) upon reasonable request.

Competing interests

The authors declare no competing interests.