Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Towards cavity-free ground-state cooling of an acoustic-frequency silicon nitride membrane

Open Access Open Access

Abstract

We demonstrate feedback cooling of a millimeter-scale, 40 kHz SiN membrane from room temperature to 5 mK (3000 phonons) using a Michelson interferometer, and discuss the challenges to ground-state cooling without an optical cavity. This advance appears within reach of current membrane technology, positioning it as a compelling alternative to levitated systems for quantum sensing and fundamental weak force measurements.

© 2020 Optical Society of America

Strained thin-film resonators (strings and membranes) with millimeter dimensions can support acoustic frequency modes with extremely high quality factors, leveraging the effect of dissipation dilution [14]. It has been speculated that they may enable room-temperature quantum optomechanics [3], ultraprecise force and acceleration sensing [5,6], quantum memories, and detection of fundamental weak signals such as spontaneous waveform collapse [7] and ultralight dark matter [8].

Here we discuss an additional potential of acoustic frequency thin-film resonators, which is to remove the need for a cavity in quantum optomechanics experiments. This possibility is due to the large zero-point fluctuations of an acoustic frequency nanomechanical resonator, as exploited in experiments with levitated nanoparticles [9]. In contrast to levitated nanoparticles, thin-film resonators can be read out with high efficiency by direct reflection or near-field sensing [10]. The main challenge to reaching the quantum regime is technical noise, such as laser relaxation oscillations, which can be far in excess of shot noise at acoustic frequencies. Optical absorption can 2also lead to large bolometric effects in tethered nanostructures, while in principle they can be decoupled from motion of a levitated particle [11].

To illustrate the potential for “cavity-free” quantum optomechanics, we describe an experiment in which the fundamental mode of a 2.5 mm, high-stress silicon nitride (${{\rm Si}_3}{{\rm N}_4}$) trampoline resonator [3,4] is subject to radiation pressure feedback cooling using a Michelson interferometer. The conditions for ground-state cooling are twofold [12]: (1) the oscillator’s thermal decoherence rate ${\Gamma _{{\rm th}}}$ must not exceed its frequency ${\Omega _0}$,

$${\Gamma _{{\rm th}}} = \frac{{{k_{\rm B}}{T_0}}}{{\hbar {Q_0}}} \lt {\Omega _0},$$
and (2) the measurement imprecision $S_\textit{xx}^{{\rm imp}}$ (expressed as a single-sided power spectral density) must be low enough to resolve zero-point motion ${x_{{\rm zp}}}$ in the thermal decoherence time
$$S_\textit{xx}^{{\rm imp,gs}} = \frac{{4x_{{\rm zp}}^2}}{{{\Gamma _{{\rm th}}}}} = \frac{{2{\hbar ^2}}}{{{k_{\rm B}}{T_0}}}\frac{{{Q_0}}}{{m{\Omega _0}}},$$
where ${T_0}$ is the intrinsic device temperature.

In our experiment, operated at room temperature, an optimized trampoline design [3,4] yields a fundamental frequency of ${\Omega _0} = 2\pi \cdot 40\;{\rm kHz} $, a quality factor ${Q_0} = 3 \times {10^7}$, and an effective mass of $m = 12 \;{\rm ng}$, corresponding to a thermal decoherence rate of ${\Gamma _{{\rm th}}} = 5{\Omega _0}$, a zero-point displacement of ${x_{{\rm zp}}} = 4\;{\rm fm}$, and a ground-state cooling requirement of [Eq. (2)] ${(S_\textit{xx}^{{\rm imp,gs}})^{1/2}} \approx {10^{- 17}}\;{\rm m}/\sqrt {{\rm Hz}}$. The latter is 3 orders of magnitude below the sensitivity of our interferometer; nevertheless, using an auxiliary laser field as a radiation pressure actuator, we realize feedback cooling to an effective temperature of ${T_{{\rm eff}}} = 5\;{\rm mK}$, corresponding to a mean phonon number of

$$\langle n\rangle = \frac{{{k_B}{T_{{\rm eff}}}}}{{\hbar {\Omega _0}}} \approx \sqrt {\frac{{S_\textit{xx}^{{\rm imp}}}}{{S_\textit{xx}^{{\rm imp,gs}}}}} \approx 3 \times {10^3}.$$
Below we discuss the design and limitations of this experiment and speculate that $\langle n\rangle \sim 1$ should be possible with simple modifications, including pre-cooling in a helium cryostat and using a common path interferometer topology.

1. MEASUREMENT-BASED FEEDBACK COOLING

In feedback cooling protocols, a continuous position measurement is used to suppress the thermal motion of a mechanical oscillator by derivative feedback (velocity damping). The technique dates back to regulation of electrometers [13] and is commonly used in atomic force microscopes to increase their dynamic range [14]. More recently, feedback cooling has received attention in the cavity optomechanics community as a means to prepare a nanomechanical oscillator in its ground state [15,16]. Cooling to $\langle n\rangle \sim 4$ has been achieved with a free space optically levitated nanoparticle [17], while cavity-enhanced measurements have been used to cool a ${{\rm Si}_3}{{\rm N}_4}$ nanostring to $\langle n\rangle \sim 4$ [12] and, more recently, a ${{\rm Si}_3}{{\rm N}_4}$ membrane to $\langle n\rangle \sim 0.3$ [18].

An important feature of feedback cooling is that, unlike optomechanical sideband cooling [19,20], it does not require a “good” (sideband-resolved) cavity to reach the ground state [21,22]. It is only necessary to achieve sufficient measurement efficiency, which in fact requires a “bad” cavity and in principle requires no cavity at all (enabling use of nonresonant sensors such as single-electron transistors and superconducting quantum interference devices) [17,2326]. To this end, agnostic to the measurement scheme, consider a real-time estimate $y$ of the oscillator displacement $x$ obscured by imprecision noise ${x_{{\rm imp}}}$:

$$y = x + {x_{{\rm imp}}}.$$
Feedback cooling can be understood by including a velocity-proportional feedback force in the Langevin equation [24]
$$m\ddot x + m{\Gamma _{0}}\dot x + m\Omega _{0}^2x = \sqrt {2{k_B}{T_0}m{\Gamma _0}} \xi (t) - gm{\Gamma _0}\dot y,$$
where $\xi (t)$ is a normalized Gaussian white noise process and ${\Gamma _0} = {\Omega _0}/{Q_0}$ is the intrinsic mechanical damping rate. Applying the Wiener–Khinchin theorem, the spectral density of physical $(x)$ and apparent $(y)$ displacement can be expressed as [12,18]
$$\frac{{{S_\textit{xx}}[\Omega]}}{{2S_\textit{xx}^{{\rm zp}}}} = |{\chi _g}[\Omega {]|^2}\left({{n_{{\rm th}}} + {g^2}{n_{{\rm imp}}}} \right),$$
$$\frac{{{S_\textit{yy}}[\Omega]}}{{2S_\textit{xx}^{{\rm zp}}}} = |{\chi _g}[\Omega {]|^2}\left({{n_{{\rm th}}} + {{(1 + g)}^2}|{\chi _0}[\Omega {{]|}^{- 2}}{n_{{\rm imp}}}} \right),$$
where $S_\textit{xx}^{{\rm zp}} = 4x_{{\rm zp}}^2/{\Gamma _0}$ is the zero-point displacement spectral density, ${n_{{\rm th}}} = {k_{\rm B}}{T_0}/m{\Omega _0}$ is the thermal bath occupation, and $\chi _g^{- 1} \approx (1 + g) + 2i(\Omega - {\Omega _0})/{\Gamma _0}$ is the closed-loop mechanical susceptibility. Evidently feedback damping can be “cold” in the sense that ${n_{{\rm imp}}} = S_\textit{xx}^{{\rm imp}}/2S_\textit{xx}^{{\rm zp}} \lt {n_{{\rm th}}}$ when the measurement resolves the thermal motion. Increasing the feedback gain $g$ thus reduces the average displacement of the oscillator $\langle {x^2}\rangle = \int\! {S_\textit{xx}}[\Omega]/2\pi$, resulting in a mean phonon number of
$$\langle n\rangle + \frac{1}{2} = \frac{{\langle {x^2}\rangle}}{{2x_{{\rm zp}}^2}} = \frac{{{n_{{\rm th}}} + {g^2}{n_{{\rm imp}}}}}{{1 + g}} \ge 2\sqrt {{n_{{\rm th}}}{n_{{\rm imp}}}}.$$
Ground-state cooling requires accounting for measurement back-action ${n_{{\rm th}}} \to {n_{{\rm th}}} + {n_{{\rm ba}}} = {n_{{\rm th}}} + \eta /(16{n_{{\rm imp}}})$, where $\eta \in [0,1]$ is the measurement efficiency [12,18]. Equation (7) thus yields Eq. (2) for $\langle n\rangle \lt 1$ and Eq. (3) for $1 \ll \langle n\rangle \ll {n_{{\rm th}}}$ (noting that ${\Gamma _{{\rm th}}} = {\Gamma _0}{n_{{\rm th}}}$). In addition to high efficiency, we emphasize that reaching low occupancy is facilitated by having a high $Q/m{\Omega _0}$ factor, which is equivalent to a high force sensitivity $S_\textit{FF}^{{\rm th}} = 4{k_B}Tm{\Omega _0}/Q$.

2. TRAMPOLINE RESONATOR

Our mechanical resonator is a modified version of the ${{\rm Si}_3}{{\rm N}_4}$ trampoline introduced by Reinhardt [4] and Norte et al. [3]. Trampoline resonators, like strings [2729], exhibit quality factors scaling as $Q \propto {Q_{\text{mat}}}(h)\sqrt \sigma L/h$, where $Q_{{\rm mat}}^{- 1}$ is the material loss tangent, $L$ is the tether length, $h$ is the film thickness, and $\sigma$ is the tensile stress in the film. Since ${\Omega _0} \propto \sqrt \sigma /L$ and $m \propto hL$, the implication is that $Q/m{\Omega _0} \propto {Q_{{\rm mat}}}(h)L/{h^2}$. Counterintuitively, larger devices can have larger zero-point fluctuations.

 figure: Fig. 1.

Fig. 1. ${{\rm Si}_3}{{\rm N}_4}$ trampoline resonator. (Top left) Camera image of a typical device. (Top right) Microscope image of the trampoline used in the experiment. (Bottom right) Finite element simulation of the fundamental 40 kHz vibrational mode. (Bottom left) Energy ringdown of the fundamental mode before (red) and after (blue) deposition of a dust particle onto a tether.

Download Full Size | PDF

The trampoline used in our experiment is shown in Fig. 1. The device is suspended from a $h \approx 90\,{\rm nm}$ thick ${{\rm Si}_3}{{\rm N}_4}$ film on a $200\,\unicode{x00B5}{\rm m}$ thick Si wafer (WaferPro) using a standard two-sided photolithography and wet etching technique [3,4,29]. A $200\,\unicode{x00B5}{\rm m}$ pad and $L \approx 1.7$ mm long, $w = 4.2\,\,\unicode{x00B5}{\rm m}$ wide tethers were chosen, as well as an “optimal” [30] $50\,\,\unicode{x00B5}{\rm m}$ radius fillet for this window size (2.5 mm) and tether width. Mechanical ringdown measurements performed in high vacuum [${ \lt\! 10^{- 7}}\; {\rm mbar}$, Fig. 1(c)] reveal a fundamental frequency of ${\Omega _0} = 2\pi \times 39.9\;{\rm kHz} $ and a quality factor as high as $Q = 4.4 \times {10^7}$, which agrees with finite element simulations (COMSOL) assuming a film stress of $\sigma = 0.9 \; {\rm GPa}$ and internal quality factor of ${Q_{{\rm mat}}} = 6 \times {10^3}$. (The latter is consistent with the ${Q_{\text{mat}}} \propto h$ surface loss model of Villanueva et al. [28], suggesting that our device is not limited by clamping loss.) For the experiments described below, dust deposited on the tether resulted in a reduced quality factor of $Q = 2.6 \times {10^7}$. Together with a simulated effective mass of $m = 12 \; {\rm ng}$, this implies a force sensitivity of $S_\textit{FF}^{{\rm th}} = (43\;{\rm aN}/\sqrt {{\rm Hz}} {)^2}$, a zero-point displacement of $S_\textit{xx}^{{\rm zp}} = (86 \; {\rm fm}/\sqrt {{\rm Hz}} {)^2}$, and a ground-state cooling requirement $S_\textit{xx}^{{\rm imp,gs}} = S_\textit{xx}^{{\rm zp}}/{n_{{\rm th}}} \approx {(0.68 \times {10^{- 17}} \;{\rm m}/\sqrt {{\rm Hz}})^2}$.

 figure: Fig. 2.

Fig. 2. Setup for probing the trampoline, consisting of a confocal microscope embedded in a balanced Michelson interferometer. Electronics for stabilizing the interferometer path length (PI = proportional integral controller, Newport LB1005) and for radiation pressure feedback cooling (see main text for details) are indicated in black. An image of the focused optical beam on the trampoline pad is shown at bottom left.

Download Full Size | PDF

3. INTERFEROMETRIC READOUT

Displacement of the trampoline was read out using a confocal microscope integrated into a Michelson interferometer [10]. Details are shown in Fig. 2. To minimize gas damping, the device chip is mounted in a high vacuum chamber operating at ${ \lt\! 10^{- 7}}\; {\rm mbar}$. This is enabled by a long-working-distance microscope objective (Mitutoyo M Plan APO 10X) with a spot diameter of ${\lt}5\,\,\unicode{x00B5}{\rm m}$. The light source used for the experiment was an 850 nm external cavity diode laser (Newport TLB-6716). To mitigate laser frequency and intensity noise, both the arm length and power of the interferometer were carefully balanced.

 figure: Fig. 3.

Fig. 3. Characterization of interferometer sensitivity. Upper plot: imprecision in noise quanta units versus power, compared to Eq. (8) (blue line). Lower plot: apparent displacement spectrum of the trampoline versus frequency for different optical powers. The dashed line is a model for ${S_\textit{xx}}$. The blue line is obtained by blocking the signal arm of the interferometer at highest power. (Inset: broadband spectrum for highest power).

Download Full Size | PDF

Ideally, the interferometer sensitivity is limited by shot noise

$$S_\textit{xx}^{{\rm imp,shot}} = \frac{{\hbar c\lambda}}{{16\pi\! \eta}}\frac{{{R_{\rm m}}}}{P},$$
where $P$ is the power incident on the membrane, $\lambda$ is the laser wavelength, ${R_{\rm m}}$ is the membrane reflectance, and $\eta \in [0,1]$ is the detection efficiency. We investigated this limit by recording apparent displacement spectra ${S_\textit{yy}}$ at different optical powers, where $y$ is proportional to the voltage signal produced by the balanced photoreceiver (Newport 1807). Results are shown in Fig. 3. To calibrate each measurement, a piezo underneath the device chip was used to drive the trampoline near its fundamental resonance with a coherent amplitude of ${x_{{\rm cal}}} = 0.8$ pm (inferred by bootstrapping to the area beneath the thermal noise peak, $\langle {x^2}\rangle \approx 2x_{{\rm zp}}^2{n_{{\rm th}}}$, after an averaging time ${\gt}{Q_0}/{\Omega _0}$). At low powers ($P \lt 100\; \unicode{x00B5}{\rm W}$), $S_\textit{xx}^{{\rm imp}}$ scales inversely with power, as expected for shot noise, with an apparent efficiency of $\eta \approx 10\%$ (using ${R_{\rm m}} = 0.3$ [31]). This value is consistent with the return loss of our microscope objective (which employs a free-space-to-fiber coupler), and in principle allows cooling to $\langle n\rangle \approx 1.1$.

In practice, the interferometer is limited by extraneous noise at sufficiently high power. This is seen in Fig. 3 for powers above 1 mW, where the noise floor saturates to $S_\textit{xx}^{{\rm imp,ext}} \approx 10 \;{\rm fm}/\sqrt {{\rm Hz}}$. Broadband measurements (Fig. 3, inset) suggest that $S_\textit{xx}^{{\rm imp,ext}}$ is related to differential polarization or path-length fluctuations, possibly exacerbated by peaking of the homodyne phase lock. (We note that extraneous laser frequency noise was ruled out by introducing a path length imbalance of several millimeters, to no apparent effect.) Although an impressive 2 orders of magnitude below the intrinsic zero-point motion $({n_{{\rm imp,ext}}} \approx 0.01)$, this extraneous noise practically limits feedback cooling of the fundamental trampoline mode to $\langle n\rangle \approx 2.5 \times {10^3}$ starting at room temperature (${n_{{\rm th}}} = 1.6 \times {10^8}$), according to Eq. (7).

Heating from optical absorption is another important consideration at high powers. To investigate this effect, we consider the off-resonant thermal noise in Fig. 3, which in the presence of heating should increase linearly with power (${S_\textit{xx}}[\Omega - {\Omega _0} \gg \;g{\Gamma _0}] \propto {n_{{\rm th}}}{\Gamma _0}/(\Omega - {\Omega _0}{)^2}$). The variation observed is within the ${\sim}10\%$ statistical error of the spectral density estimate, suggesting that heating is less than $10 \; {\rm K}/{\rm mW}$. This value is consistent with a simple heat conduction model $dT/dP \approx \alpha L/4wh\kappa$, assuming a thermal conductivity of $\kappa = 3 \; {\rm W}/{\rm m} {\rm K}$ and a conservative optical absorption coefficient of $\alpha = 10\;{\rm ppm} $ [31].

4. RADIATION PRESSURE FEEDBACK COOLING

Feedback cooling was carried out using radiation pressure actuation. The main advantage of this approach is its high bandwidth; however, we note that other methods such as piezoelectric [32] and dielectric [33] actuation are in principle equally viable and may be simpler to interface with feedback electronics.

To implement radiation pressure feedback, we introduce a second laser beam into the microscope, which is intensity modulated by an amplified copy of the photosignal. Specifically, we use a 670 nm laser diode (Hitachi HL6712) modulated by dithering its drive current about the threshold value. To approximate derivate feedback while suppressing feedback to higher-order modes, the photosignal is passed through a 10–50 kHz bandpass filter and a delay line, resulting in an approximately $\phi = 90^ \circ$ phase shift for frequencies near mechanical resonance. The feedback force can in this case be approximated as

$$\delta\! {F_{{\rm fb}}} \approx - g\!m{\Gamma _0}(\dot y + {\Omega _0}\cot (\phi)y),$$
corresponding to a normalized susceptibility ${\chi _g}{[\Omega]^{- 1}} \approx (1 + g) + 2i(\Omega - {\Omega _0})/{\Gamma _0} + ig\cot (\phi)$, where $ig\cot (\phi)$ is a residual feedback stiffening term that contributes negligibly to cooling.

The results of feedback cooling with a 3 mW read out beam and a 60 µW feedback beam are shown in Fig. 4. The feedback gain is tuned electronically using a voltage preamplifier (Stanford Research Systems SR560). To estimate $\langle n\rangle$, thermal noise spectra are fit to Eq. (6a) with $g$ as a free parameter, assuming ${n_{{\rm th}}} = 1.56 \times {10^8}$, ${n_{{\rm imp}}} = 0.013$, ${\Gamma _0} = 2\pi \cdot 1.5\; {\rm mHz}$, and $\phi = - 0.15$ [Eq. (9)]. To facilitate fitting, in Fig. 4, we focus on high gain settings for which the loaded damping rate $(1 + g){\Gamma _0} \gt 1\; {\rm Hz}$. The model accurately reproduces the noise spectra until the damped peak coincides with the noise floor, for which the inferred gain is $g = 1.4 \times {10^5}$, corresponding to $\langle n\rangle = 3.0 \times {10^3}$. At higher gain, the noise floor exhibits typical “squashing” behavior [12,34], and the inferred $\langle n\rangle$ begins to increase.

 figure: Fig. 4.

Fig. 4. Radiation pressure feedback cooling. Upper plot: feedback cooling curve for parameters described in the main text. Colored points correspond to models overlaying experimental data. Lower plot: experimental measurements (colored) overlaid with models (dashed curves) using Eq. (9). The solid black curve is a model for $g = 0$ (no feedback).

Download Full Size | PDF

5. SUMMARY AND OUTLOOK

We have demonstrated measurement-based feedback cooling of a 40 kHz ${{\rm Si}_3}{{\rm N}_4}$ trampoline resonator from room temperature ($1.6 \times {10^8}$ phonons) to an effective temperature of 5 mK ($3 \times {10^3}$ phonons) using a simple two-path interferometer. The main limitation of our experiment is technical noise at the level of $10 \; {\rm fm}/\sqrt {{\rm Hz}}$. Absent this noise, the apparent 10% efficiency of our interferometer would in principle enable cooling to $\langle n\rangle \sim 100$ with a probe power of several megawatts. Operating at 4 K, assuming no increase in mechanical Q and no photothermal heating, would enable cooling to $\langle n\rangle \sim 10$, for which motional sideband asymmetry could be readily measured.

We speculate that a combination of monolithic interferometer design and moderate cryogenics could give access to $\langle n\rangle \sim 1$ without an optical cavity for state-of-the-art ${{\rm Si}_3}{{\rm N}_4}$ thin-film resonators. Particularly compelling are “soft-clamped” nanobeams [2], which have demonstrated megahertz modes with quality factors approaching ${10^9}$ and zero-point spectral densities exceeding $1 \;{\rm pm}/\sqrt {{\rm Hz}}$, and can be read out with high efficiency by evanescent coupling to optical waveguide. For soft-clamped resonators, an important challenge is the large density of states and low thermal conductance of the phononic crystal shield, which reduces power handling capacity and can introduce extraneous thermal noise. Clamp-optimized trampolines [3,4,30] might offer a simpler route, since the fundamental mode is well isolated and can also have ${Q_0}{ \gt 10^8}$ at millimeter dimensions [2]. In the future, resonators made of strained crystalline thin films promise ${Q_0}{ \gt 10^9}$ and increased thermal conductivity at 4 K [35]. A recent proposal for soft-clamping fundamental modes using a “fractal clamp” might push this performance even further [36].

Funding

National Science Foundation (ECCS-1725571).

Acknowledgment

C.M.P. gratefully acknowledges support from an Amherst College Fellowship. The reactive ion etcher used in this study was acquired by an NSF MRI grant ECCS-1725571.

Disclosures

The authors declare no conflicts of interest.

REFERENCES AND NOTE

1. Y. Tsaturyan, A. Barg, E. S. Polzik, and A. Schliesser, “Ultracoherent nanomechanical resonators via soft clamping and dissipation dilution,” Nat. Nanotechnol. 12, 776–783 (2017). [CrossRef]  

2. A. H. Ghadimi, S. A. Fedorov, N. J. Engelsen, M. J. Bereyhi, R. Schilling, D. J. Wilson, and T. J. Kippenberg, “Elastic strain engineering for ultralow mechanical dissipation,” Science 360, 764–768 (2018). [CrossRef]  

3. R. A. Norte, J. P. Moura, and S. Gröblacher, “Mechanical resonators for quantum optomechanics experiments at room temperature,” Phys. Rev. Lett. 116, 147202 (2016). [CrossRef]  

4. C. Reinhardt, T. Müller, A. Bourassa, and J. C. Sankey, “Ultralow-noise sin trampoline resonators for sensing and optomechanics,” Phys. Rev. X 6, 021001 (2016). [CrossRef]  

5. R. Fischer, D. P. McNally, C. Reetz, G. G. Assumpcao, T. Knief, Y. Lin, and C. A. Regal, “Spin detection with a micromechanical trampoline: towards magnetic resonance microscopy harnessing cavity optomechanics,” New J. Phys. 21, 043049 (2019). [CrossRef]  

6. A. G. Krause, M. Winger, T. D. Blasius, Q. Lin, and O. Painter, “A high-resolution microchip optomechanical accelerometer,” Nat. Photonics 6, 768–772 (2012). [CrossRef]  

7. S. Nimmrichter, K. Hornberger, and K. Hammerer, “Optomechanical sensing of spontaneous wave-function collapse,” Phys. Rev. Lett. 113, 020405 (2014). [CrossRef]  

8. D. Carney, A. Hook, Z. Liu, J. M. Taylor, and Y. Zhao, “Ultralight dark matter detection with mechanical quantum sensors,” arXiv preprint arXiv:1908.04797 (2019).

9. Z.-Q. Yin, A. A. Geraci, and T. Li, “Optomechanics of levitated dielectric particles,” Int. J. Mod. Phys. B 27, 1330018 (2013). [CrossRef]  

10. A. Barg, Y. Tsaturyan, E. Belhage, W. H. Nielsen, C. B. Møller, and A. Schliesser, “Measuring and imaging nanomechanical motion with laser light,” Appl. Phys. B 123, 8 (2017). [CrossRef]  

11. E. Hebestreit, R. Reimann, M. Frimmer, and L. Novotny, “Measuring the internal temperature of a levitated nanoparticle in high vacuum,” Phys. Rev. A 97, 043803 (2018). [CrossRef]  

12. D. Wilson, V. Sudhir, N. Piro, R. Schilling, A. Ghadimi, and T. Kippenberg, “Measurement-based control of a mechanical oscillator at its thermal decoherence rate,” Nature 524, 325–329 (2015). [CrossRef]  

13. J. Milatz, J. Van Zolingen, and B. Van Iperen, “The reduction in the Brownian motion of electrometers,” Physica 19, 195–202 (1953). [CrossRef]  

14. J. Mertz, O. Marti, and J. Mlynek, “Regulation of a microcantilever response by force feedback,” App. Phys. Lett. 62, 2344–2346 (1993). [CrossRef]  

15. J.-M. Courty, A. Heidmann, and M. Pinard, “Quantum limits of cold damping with optomechanical coupling,” Eur. Phys. J. D 17, 399–408 (2001). [CrossRef]  

16. D. Vitali, S. Mancini, L. Ribichini, and P. Tombesi, “Macroscopic mechanical oscillators at the quantum limit through optomechanical cooling,” J. Opt. Soc. Am. B 20, 1054–1065 (2003). [CrossRef]  

17. F. Tebbenjohanns, M. Frimmer, V. Jain, D. Windey, and L. Novotny, “Motional sideband asymmetry of a nanoparticle optically levitated in free space,” Phys. Rev. Lett. 124, 013603 (2020). [CrossRef]  

18. M. Rossi, D. Mason, J. Chen, Y. Tsaturyan, and A. Schliesser, “Measurement-based quantum control of mechanical motion,” Nature 563, 53–58 (2018). [CrossRef]  

19. I. Wilson-Rae, N. Nooshi, W. Zwerger, and T. J. Kippenberg, “Theory of ground state cooling of a mechanical oscillator using dynamical backaction,” Phys. Rev. Lett. 99, 093901 (2007). [CrossRef]  

20. F. Marquardt, J. P. Chen, A. A. Clerk, and S. Girvin, “Quantum theory of cavity-assisted sideband cooling of mechanical motion,” Phys. Rev. Lett. 99, 093902 (2007). [CrossRef]  

21. C. Genes, D. Vitali, P. Tombesi, S. Gigan, and M. Aspelmeyer, “Ground-state cooling of a micromechanical oscillator: Comparing cold damping and cavity-assisted cooling schemes,” Phys. Rev. A 77, 033804 (2008). [CrossRef]  

22. K. Jacobs, H. I. Nurdin, F. W. Strauch, and M. James, “Comparing resolved-sideband cooling and measurement-based feedback cooling on an equal footing: analytical results in the regime of ground-state cooling,” Phys. Rev. A 91, 043812 (2015). [CrossRef]  

23. P. Bushev, D. Rotter, A. Wilson, F. Dubin, C. Becher, J. Eschner, R. Blatt, V. Steixner, P. Rabl, and P. Zoller, “Feedback cooling of a single trapped ion,” Phys. Rev. Lett. 96, 043003 (2006). [CrossRef]  

24. M. Poggio, C. Degen, H. Mamin, and D. Rugar, “Feedback cooling of a cantilever’s fundamental mode below 5 mK,” Phys. Rev. Lett. 99, 017201 (2007). [CrossRef]  

25. A. Hopkins, K. Jacobs, S. Habib, and K. Schwab, “Feedback cooling of a nanomechanical resonator,” Phys. Rev. B 68, 235328 (2003). [CrossRef]  

26. A. Vinante, A. Kirste, A. Den Haan, O. Usenko, G. Wijts, E. Jeffrey, P. Sonin, D. Bouwmeester, and T. Oosterkamp, “High sensitivity squid-detection and feedback-cooling of an ultrasoft microcantilever,” App. Phys. Lett. 101, 123101 (2012). [CrossRef]  

27. A. H. Ghadimi, D. J. Wilson, and T. J. Kippenberg, “Radiation and internal loss engineering of high-stress silicon nitride nanobeams,” Nano Lett. 17, 3501–3505 (2017). [CrossRef]  

28. L. G. Villanueva and S. Schmid, “Evidence of surface loss as ubiquitous limiting damping mechanism in sin micro- and nanomechanical resonators,” Phys. Rev. Lett. 113, 227201 (2014). [CrossRef]  

29. We follow the method of Reinhardt et al. [4] using a simplified Teflon sample holder for wet etching. Fabrication begins with standard two-sided photolithography followed by dry etch of ${{\rm Si}_3}{{\rm N}_4}$, using plasma Ar+SF$_{6}$, to define the trampoline on one side and a square window on the other. The chip is then wet etched from both sides using a 45% percent KOH solution at 65 C for 21 hours. After the trampoline is released, the gradual dilution method described in [4] is used in lieu of a critical point dryer, in order to suspend the trampoline.

30. P. Sadeghi, M. Tanzer, S. L. Christensen, and S. Schmid, “Influence of clamp-widening on the quality factor of nanomechanical silicon nitride resonators,” J. Appl. Phys. 126, 165108 (2019). [CrossRef]  

31. D. Wilson, C. Regal, S. Papp, and H. Kimble, “Cavity optomechanics with stoichiometric sin films,” Phys. Rev. Lett. 103, 207204 (2009). [CrossRef]  

32. S. Sridaran and S. A. Bhave, “Electrostatic actuation of silicon optomechanical resonators,” Opt. Express 19, 9020–9026 (2011). [CrossRef]  

33. Q. P. Unterreithmeier, E. M. Weig, and J. P. Kotthaus, “Universal transduction scheme for nanomechanical systems based on dielectric forces,” Nature 458, 1001–1004 (2009). [CrossRef]  

34. M. Poggio, C. Degen, H. Mamin, and D. Rugar, “Feedback cooling of a cantilever’s fundamental mode below 5 mK,” Phys. Rev. Lett. 99, 017201 (2007). [CrossRef]  

35. E. Romero, V. M. Valenzuela, A. R. Kermany, F. Iacopi, and W. P. Bowen, “Engineering the dissipation of crystalline micromechanical resonators,” Phys. Rev. Applied 13, 04407 (2020). [CrossRef]  

36. S. A. Fedorov, A. Beccari, N. J. Engelsen, and T. J. Kippenberg, “Fractal-like mechanical resonators with a soft-clamped fundamental mode,” Phys. Rev. Lett. 124, 025502 (2020). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. ${{\rm Si}_3}{{\rm N}_4}$ trampoline resonator. (Top left) Camera image of a typical device. (Top right) Microscope image of the trampoline used in the experiment. (Bottom right) Finite element simulation of the fundamental 40 kHz vibrational mode. (Bottom left) Energy ringdown of the fundamental mode before (red) and after (blue) deposition of a dust particle onto a tether.
Fig. 2.
Fig. 2. Setup for probing the trampoline, consisting of a confocal microscope embedded in a balanced Michelson interferometer. Electronics for stabilizing the interferometer path length (PI = proportional integral controller, Newport LB1005) and for radiation pressure feedback cooling (see main text for details) are indicated in black. An image of the focused optical beam on the trampoline pad is shown at bottom left.
Fig. 3.
Fig. 3. Characterization of interferometer sensitivity. Upper plot: imprecision in noise quanta units versus power, compared to Eq. (8) (blue line). Lower plot: apparent displacement spectrum of the trampoline versus frequency for different optical powers. The dashed line is a model for ${S_\textit{xx}}$. The blue line is obtained by blocking the signal arm of the interferometer at highest power. (Inset: broadband spectrum for highest power).
Fig. 4.
Fig. 4. Radiation pressure feedback cooling. Upper plot: feedback cooling curve for parameters described in the main text. Colored points correspond to models overlaying experimental data. Lower plot: experimental measurements (colored) overlaid with models (dashed curves) using Eq. (9). The solid black curve is a model for $g = 0$ (no feedback).

Equations (10)

Equations on this page are rendered with MathJax. Learn more.

Γ t h = k B T 0 Q 0 < Ω 0 ,
S xx i m p , g s = 4 x z p 2 Γ t h = 2 2 k B T 0 Q 0 m Ω 0 ,
n = k B T e f f Ω 0 S xx i m p S xx i m p , g s 3 × 10 3 .
y = x + x i m p .
m x ¨ + m Γ 0 x ˙ + m Ω 0 2 x = 2 k B T 0 m Γ 0 ξ ( t ) g m Γ 0 y ˙ ,
S xx [ Ω ] 2 S xx z p = | χ g [ Ω ] | 2 ( n t h + g 2 n i m p ) ,
S yy [ Ω ] 2 S xx z p = | χ g [ Ω ] | 2 ( n t h + ( 1 + g ) 2 | χ 0 [ Ω ] | 2 n i m p ) ,
n + 1 2 = x 2 2 x z p 2 = n t h + g 2 n i m p 1 + g 2 n t h n i m p .
S xx i m p , s h o t = c λ 16 π η R m P ,
δ F f b g m Γ 0 ( y ˙ + Ω 0 cot ( ϕ ) y ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.