Optical Control of Nanomechanical Brownian Motion Eigenfrequencies in MetamaterialsClick to copy article linkArticle link copied!
- Jinxiang LiJinxiang LiOptoelectronics Research Centre, University of Southampton, Highfield, Southampton SO17 1BJ, United KingdomMore by Jinxiang Li
- Kevin F. MacDonald*Kevin F. MacDonald*Email: [email protected]Optoelectronics Research Centre, University of Southampton, Highfield, Southampton SO17 1BJ, United KingdomMore by Kevin F. MacDonald
- Nikolay I. ZheludevNikolay I. ZheludevOptoelectronics Research Centre, University of Southampton, Highfield, Southampton SO17 1BJ, United KingdomCentre for Disruptive Photonic Technologies and The Photonics Institute, SPMS, Nanyang Technological University Singapore 637371, SingaporeMore by Nikolay I. Zheludev
Abstract
Nanomechanical photonic metamaterials provide a wealth of active switching, nonlinear, and enhanced light-matter interaction functionalities by coupling optically and mechanically resonant subsystems. Thermal (Brownian) motion of the nanostructural components of such metamaterials leads to fluctuations in optical properties, which may manifest as noise, but which also present opportunity to characterize performance and thereby optimize design at the level of individual nanomechanical elements. We show that nanomechanical motion in an all-dielectric metamaterial ensemble of silicon-on-silicon-nitride nanowires can be controlled by light at sub-μW/μm2 intensities. Induced changes in nanowire temperature of just a few Kelvin and nonthermal optical forces generated within the structure change the few-MHz Eigenfrequencies and/or picometric displacement amplitudes of motion, and thereby metamaterial transmission. The tuning mechanism can provide active control of frequency response in photonic metadevices and may serve as a basis for bolometric, mass, and micro/nanostructural stress sensing.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
By virtue of their low mass and fast (MHz–GHz frequency) response times, nanomechanical oscillators actuated and/or interrogated by light are of fundamental and applied interest in numerous applications, ranging from mass and force sensors to photonic data processing and quantum ground state measurements. (1−13) As the dimensions of such systems decrease, their thermal (i.e., Brownian) motion assumes increasing importance. By adding noise to induced/controlled movements that underpin the functionality, it can constrain performance, but it also presents opportunity by directly linking observable (far-field optical) properties to geometry, composition, and temperature at the nanoscale. We show here that such motion can be optically controlled at μW/μm2 intensities in nanomechanical photonic metamaterials. In an array of mechanically independent and (by design) alternately dissimilar dielectric nanowires, which are at the same time of identical bilayer (i.e., asymmetric) material composition and part of an optically resonant ensemble subject to the fundamental constraint of linear transmission reciprocity, dependences of motion Eigenfrequencies and picometric displacement amplitudes on local light-induced temperature changes can be accurately determined. We further show that the amplitude of high-frequency oscillatory motion driven by nonthermal optical (gradient and radiation pressure) forces is resonantly enhanced, leading to larger changes in transmission, when the nanowires’ natural frequencies are photothermally tuned to coincide with the pump modulation frequency.
In the present study, we employ an all-dielectric metamaterial comprising pairs of dissimilar (by length and width) silicon nanobricks on a free-standing array of flexible silicon nitride nanowires (Figure 1a). It is fabricated on a 200 nm thick Si3N4 membrane coated by plasma-enhanced chemical vapor deposition with a 115 nm thick layer of amorphous Si. This bilayer is then structured by focused ion beam milling to define rows of alternately short, narrow (720 nm × 210 nm) and long, wide (780 nm × 300 nm) nanobricks in the Si layer on parallel 21 μm long nanowire beams cut through the Si3N4 layer with a gap size between neighboring nanowires of 170 nm.
The metamaterial structure supports a near-infrared closed mode optical resonance (14) at a wavelength of 1542 nm (see Supplementary Figure S1), underpinned by the excitation of antiparallel displacement currents in adjacent dissimilar silicon nanobricks by incident light polarized parallel to the long axis of the bricks. In the vicinity of this optical resonance, thermal (Brownian) motion of the nanowires, mutual positional fluctuations of pico- to nanometric amplitude, translate to fluctuations of metamaterial transmission (of order 0.1%) at their few MHz natural mechanical resonance frequencies. (15,16) These thermomechanical oscillations are detected as peaks in frequency spectra of transmission amplitude spectral density (Figure 1b,c): the metamaterial is mounted in a vacuum chamber at a pressure of 4 × 10–3 mbar to exclude air damping of mechanical motion. (It should be noted here that while a classical Brownian particle in a fluid is thermally perturbed by “external” collisions with ambient atoms, the thermal motion of objects under vacuum is driven “internally” by momentum transfer from the annihilation, creation, and interference of phonons.) The sample chamber is located between a confocal pair of 20× (NA 0.4) microscope objectives, via which incident light at a wavelength of 1550 nm is focused onto the sample (to a spot of diameter ∼5 μm at the center of the metamaterial array) and transmitted light is collected. The arrangement includes a fiber-optic MEMS switch to enable transmission measurements in both directions through the sample without disturbance of its position/alignment relative to the beam path. (The “forward” direction of light propagation is designated as that for which light is incident on the Si side of the sample; “backward” is the Si3N4 side.)
Figure 1c shows a representative measurement of optical transmission amplitude spectral density (ASD) in which peaks associated with the fundamental out-of-plane flexural modes of a pair of individual nanowires, one narrow and one wide decorated, respectively, with short and long Si nanobricks, are seen. (Attribution to this oscillatory mode is confirmed through computational modeling; see Supporting Information.)
Nanowire displacement ASD can be expressed as (17,18)
An analytical model for optical control of thermomechanical (Brownian) motion resonances, tightly constrained by the requirement to describe the properties of two independent, similar (related) but not identical oscillators - narrow and wide Si3N4/Si bilayer nanowires, under two similar (related) but not identical regimes of optical excitation - forward and backward directions of illumination, provides for accurate quantitative evaluation of light-induced temperature changes in the individual nanowires and of the corresponding relationships between their resonance frequencies/amplitudes, illumination conditions, and the optical properties of the metamaterial array. From Euler–Bernoulli beam theory, (20) the stress-dependent fundamental frequency of a doubly clamped beam of homogeneous rectangular cross-section is
Assuming the presence of a heat source uniformly distributed over the rectangular cross-section at the midpoint of the beam, while the two ends are held at ambient temperature (T0 = 298 K), the equilibrium difference ΔT between average beam and ambient temperatures is (21)
For the purpose of applying these analytical expressions to the present case, we approximate the silicon nitride nanowires decorated with silicon nanobricks as simple rectangular-section beams of the same length with effective values of E, ρ, α, κ, and A derived from the material parameters of Si and Si3N4 and the volume fractions of the two materials in each nanowire type (see Supporting Information). P is taken as the optical power absorbed by a nanowire: Pabs = μγPin, where Pin is the total power incident on the metamaterial, γ is the absorption coefficient of the metamaterial, and μ is the “absorption cross-section” of an individual nanowire.
By substituting eqs 4 and 5 into eq 3 and then fitting that expression simultaneously to all four sets of experimental data points in Figure 3a (see Supporting Information) under constraints that
(i) | μ must be identical for forward and backward directions of illumination for a given nanowire (i.e., the same nanowire will intercept the same fraction of incident light in both directions); | ||||
(ii) | γ must be identical for the two nanowires for a given illumination directions (i.e., as a metamaterial ensemble property, absorption can only have a single value in each direction); | ||||
(iii) | σ0 must take a single fixed value (ensuring degeneracy of nominally forward- and backward-illumination zero-power resonant frequencies for each type of nanowire); |
Light-induced changes in nanowire temperature (Figure 3c) depend upon the direction of illumination and nanowire dimensions, that is, upon the strength of optical absorption and the rate at which heat is dissipated (the latter being lower for the nanowire of smaller cross-section). The narrow nanowire changes temperature at a rate of 27 K/μW of absorbed power, and the wide nanowire at 19 K/μW. These derived dependences of induced temperature change ΔT on laser power map to the theoretical dependences of RMS Brownian motion amplitude presented in Figure 3b via eq 2, using values of meff for the two nanowires established in the above (Figure 1c) calibration of displacement spectral density. The picometrically accurate correlation with experimental data points separately derived from integrals of PSD over frequency is remarkably good.
With the introduction of a second, modulated pump laser, one can observe in probe transmission the interplay between thermomechanical fluctuations and nonthermal optically driven motion, that is, the action of optical forces (Figure 4). Here, we employ a probe laser at a fixed power of 22 μW at a wavelength of 1540 nm, and a pump laser at 1550 nm electro-optically modulated at a fixed frequency of 2.97 MHz. The beams are coincident on the sample in the forward direction (coupled in fiber before the MEMS switch input), and a narrow bandpass filter (at the switch output) ensures that only transmitted probe light reaches the detector. Under these conditions, we observe again in the frequency spectrum of probe transmission peaks relating to Brownian motion at nanowire Eigenfrequencies dependent upon average pump power (i.e., pump-induced temperature change). But also a sharp peak at the pump modulation frequency relating to structural reconfiguration, and thereby probe transmission change, driven by nonthermal optical forces. (There is no coherent temperature oscillation induced by the pump; modulation period is some 2 orders of magnitude shorter than the nanowires’ thermal relaxation time.) The magnitude of this optically induced transmission change is resonantly enhanced by more than an order of magnitude (reaching ∼30%) when the pump modulation frequency coincides with a nanowire’s mechanical eigenfrequency (as shown in the inset in Figure 4).
In summary, we have shown that fluctuations in the resonant optical properties of a photonic metamaterial, which are associated with the mechanically resonant Brownian motion of its constituent elements, can be controlled by light at sub-μW/μm2 intensities. In an all-dielectric metamaterial ensemble of free-standing silicon nitride nanowires (mechanical oscillators) supporting an array of silicon nanobricks (optical resonators), the few MHz Eigenfrequencies and picometric amplitudes of individual nanowires’ motion are directly proportional to incident laser power, changing, respectively, in consequence of light-induced heating by up to 0.7% and 0.9% per K.
An analytical model for the photothermal tuning mechanism, simply but effectively constrained by the requirements of optical transmission reciprocity in a linear medium, links the local, nanoscopic properties and behaviors of individual nanowires (i.e., at subwavelength scale) to the far-field optical properties of the (micro/macroscopic) metamaterial ensemble. It provides for accurate evaluation of light-induced changes in nanowire temperature and of ambient condition (zero-illumination) Brownian motion Eigenfrequencies and displacement amplitudes.
We further illustrate how nonthermal optical forces generated at the near-IR optical resonance of the metamaterial structure, again at sub-μW/μm2 intensities, can be engaged to drive motion at its (photothermally tuned) mechanical Eigenfrequencies and thus to deliver strong light-induced changes in transmission.
The ability to finely tune the nanomechanical resonance characteristics of photonic metamaterials may be beneficial in a variety of metadevice applications where, for example, the frequency of nanostructural oscillation is required to match (or avoid matching) another frequency, such as that of a pulsed laser. The fact that tuning characteristics can (as here) depend strongly upon the direction of light propagation through a metamaterial by simple virtue of bilayer material composition (leading to different levels of reflection and absorption for light incident on opposing sides) may find application in devices to favor/select a single direction of propagation. The accurately quantifiable sensitivity of optical response to nanomechanical properties in such structures also suggests applications to bolometric sensing and detection of changes in mass (e.g., through adsorption/desorption) or micro/nanostructural stress.
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.nanolett.1c04900.
Metamaterial optical properties; Nanowire geometry and mechanical resonance frequencies; Nanowire effective medium parameters; Curve fitting algorithm (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
This work was supported by the Engineering and Physical Sciences Research Council (EPSRC), U.K. (Grants EP/M009122/1 and EP/T02643X/1); the Ministry of Education, Singapore (MOE2016-T3-1-006), and the China Scholarship Council (201708440254). The authors would also like to thank to Jun-Yu Ou for contributions to the design of experimental apparatus.
References
This article references 21 other publications.
- 1Chaste, J.; Eichler, A.; Moser, J.; Ceballos, G.; Rurali, R.; Bachtold, A. A nanomechanical mass sensor with yoctogram resolution. Nat. Nanotechnol. 2012, 7 (5), 301– 304, DOI: 10.1038/nnano.2012.42Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XkvVKitbc%253D&md5=9a8a31f058e85ec94624d906f5b0dd68A nanomechanical mass sensor with yoctogram resolutionChaste, J.; Eichler, A.; Moser, J.; Ceballos, G.; Rurali, R.; Bachtold, A.Nature Nanotechnology (2012), 7 (5), 301-304CODEN: NNAABX; ISSN:1748-3387. (Nature Publishing Group)Nanomech. resonators have been used to weigh cells, biomols. and gas mols., and to study basic phenomena in surface science, such as phase transitions and diffusion. These expts. all rely on the ability of nanomech. mass sensors to resolve small masses. Here, we report mass sensing expts. with a resoln. of 1.7 yg (1 yg = 10-24 g), which corresponds to the mass of one proton. The resonator is a carbon nanotube of length ∼150 nm that vibrates at a frequency of almost 2 GHz. This unprecedented level of sensitivity allows us to detect adsorption events of naphthalene mols. (C10H8), and to measure the binding energy of a xenon atom on the nanotube surface. These ultrasensitive nanotube resonators could have applications in mass spectrometry, magnetometry and surface science.
- 2Rugar, D.; Budakian, R.; Mamin, H. J.; Chui, B. W. Single spin detection by magnetic resonance force microscopy. Nature 2004, 430 (6997), 329– 332, DOI: 10.1038/nature02658Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXls1eqs78%253D&md5=289d1d5dc5543fd443fa420e5a96a6d1Single spin detection by magnetic resonance force microscopyRugar, D.; Budakian, R.; Mamin, H. J.; Chui, B. W.Nature (London, United Kingdom) (2004), 430 (6997), 329-332CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Magnetic resonance imaging (MRI) is well known as a powerful technique for visualizing subsurface structures with three-dimensional spatial resoln. Pushing the resoln. below 1 μm remains a major challenge, however, owing to the sensitivity limitations of conventional inductive detection techniques. Currently, the smallest vol. elements in an image must contain at least 1012 nuclear spins for MRI-based microscopy, or 107 electron spins for ESR microscopy. Magnetic resonance force microscopy (MRFM) was proposed as a means to improve detection sensitivity to the single-spin level, and thus enable three-dimensional imaging of macromols. (for example, proteins) with at. resoln. MRFM has also been proposed as a qubit readout device for spin-based quantum computers. Here we report the detection of an individual electron spin by MRFM. A spatial resoln. of 25 nm in one dimension was obtained for an unpaired spin in silicon dioxide. The measured signal is consistent with a model in which the spin is aligned parallel or anti-parallel to the effective field, with a rotating-frame relaxation time of 760 ms. The long relaxation time suggests that the state of an individual spin can be monitored for extended periods of time, even while subjected to a complex set of manipulations that are part of the MRFM measurement protocol.
- 3Feng, P. X.-L.; Young, D. J.; Zorman, C. A. MEMS/NEMS devices and applications. In Springer Handbook of Nanotechnology; Springer: 2017; pp 395– 429.Google ScholarThere is no corresponding record for this reference.
- 4Van Thourhout, D.; Roels, J. Optomechanical device actuation through the optical gradient force. Nat. Photonics 2010, 4, 211– 217, DOI: 10.1038/nphoton.2010.72Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXktVWqurY%253D&md5=f02fb96b10b26d8b14c20ff27f470aeeOptomechanical device actuation through the optical gradient forceVan Thourhout, Dries; Roels, JorisNature Photonics (2010), 4 (4), 211-217CODEN: NPAHBY; ISSN:1749-4885. (Nature Publishing Group)A review. Optical forces are widely used to manipulate microparticles such as living cells, DNA and bacteria. The forces used in these optical tweezers originate from the strongly varying electromagnetic field in the focus of a high-power laser beam. This field gradient polarizes the particle, causing the pos. and neg. charged sides of the dipole to experience slightly different forces. It was recently realized that the strong field gradient in the near-field of guided wave structures can also be exploited for actuating optomech. devices, and initial theor. work in this area was followed rapidly by several exptl. demonstrations. The rapid developments in this field are summarized. The origin of the optical gradient force is discussed. Several exptl. demonstrations and approaches for enhancing the strength of the effect are discussed. Some of the possible applications of the effect are reviewed.
- 5Eggleton, B. J.; Poulton, C. G.; Rakich, P. T.; Steel, M. J.; Bahl, G. Brillouin integrated photonics. Nat. Photonics 2019, 13 (10), 664– 677, DOI: 10.1038/s41566-019-0498-zGoogle Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhs1Srtr7I&md5=56781cacb80828c34eedcdad6f2182b9Brillouin integrated photonicsEggleton, Benjamin J.; Poulton, Christopher G.; Rakich, Peter T.; Steel, Michael. J.; Bahl, GauravNature Photonics (2019), 13 (10), 664-677CODEN: NPAHBY; ISSN:1749-4885. (Nature Research)A recent renaissance in Brillouin scattering research has been driven by the increasing maturity of photonic integration platforms and nanophotonics. The result is a new breed of chip-based devices that exploit acousto-optic interactions to create lasers, amplifiers, filters, delay lines and isolators. Here, we provide a detailed overview of Brillouin scattering in integrated waveguides and resonators, covering key concepts such as the stimulation of the Brillouin process, in which the optical field itself induces acoustic vibrations, the importance of acoustic confinement, methods for calcg. and measuring Brillouin gain, and the diversity of materials platforms and geometries. Our Review emphasizes emerging applications in microwave photonics, signal processing and sensing, and concludes with a perspective for future directions.
- 6Zheludev, N. I.; Plum, E. Reconfigurable nanomechanical photonic metamaterials. Nat. Nanotechnol. 2016, 11, 16– 22, DOI: 10.1038/nnano.2015.302Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XjslGqtA%253D%253D&md5=86be7d90217dd13c4aad5a8c2816b7a5Reconfigurable nanomechanical photonic metamaterialsZheludev, Nikolay I.; Plum, EricNature Nanotechnology (2016), 11 (1), 16-22CODEN: NNAABX; ISSN:1748-3387. (Nature Publishing Group)A review. The changing balance of forces at the nanoscale offers the opportunity to develop a new generation of spatially reconfigurable nanomembrane metamaterials in which electromagnetic Coulomb, Lorentz and Ampe´re forces, as well as thermal stimulation and optical signals, can be engaged to dynamically change their optical properties. Individual building blocks of such metamaterials, the metamols., and their arrays fabricated on elastic dielec. membranes can be reconfigured to achieve optical modulation at high frequencies, potentially reaching the gigahertz range. Mech. and optical resonances enhance the magnitude of actuation and optical response within these nanostructures, which can be driven by elec. signals of only a few volts or optical signals with power of only a few milliwatts. The authors envisage switchable, electrooptical, magnetooptical and nonlinear metamaterials that are compact and Si-nanofabrication-technol. compatible with functionalities surpassing those of natural media by orders of magnitude in some key design parameters.
- 7Karvounis, A.; Gholipour, B.; MacDonald, K. F.; Zheludev, N. I. Giant electro-optical effect through electrostriction in a nano-mechanical metamaterial. Adv. Mater. 2019, 31, 1804801, DOI: 10.1002/adma.201804801Google ScholarThere is no corresponding record for this reference.
- 8Zhang, Q.; Plum, E.; Ou, J.-Y.; Pi, H.; Li, J.; MacDonald, K. F.; Zheludev, N. I. Electrogyration in metamaterials: Chirality and polarization rotatory power that depend on applied electric field. Adv. Opt. Mater. 2020, 2001826, DOI: 10.1002/adom.202001826Google ScholarThere is no corresponding record for this reference.
- 9Ye, F.; Lee, J.; Feng, P. X.-L. Atomic layer MoS2 graphene van der Waals heterostructure nanomechanical resonators. Nanoscale 2017, 9 (46), 18208– 18215, DOI: 10.1039/C7NR04940DGoogle Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsF2rtbjO&md5=81a463ce9b33a257025ff19869628950Atomic layer MoS2-graphene van der Waals heterostructure nanomechanical resonatorsYe, Fan; Lee, Jaesung; Feng, Philip X.-L.Nanoscale (2017), 9 (46), 18208-18215CODEN: NANOHL; ISSN:2040-3372. (Royal Society of Chemistry)Heterostructures play significant roles in modern semiconductor devices and micro/nanosystems in a plethora of applications in electronics, optoelectronics, and transducers. While state-of-the-art heterostructures often involve stacks of cryst. epi-layers each down to a few nanometers thick, the intriguing limit would be hetero-at.-layer structures. Here we report the first exptl. demonstration of freestanding van der Waals heterostructures and their functional nanomech. devices. By stacking single-layer (1L) MoS2 on top of suspended single-, bi-, tri- and four-layer (1L to 4L) graphene sheets, we realize an array of MoS2-graphene heterostructures with varying thickness and size. These heterostructures all exhibit robust nanomech. resonances in the very high frequency (VHF) band (up to ≈100 MHz). We observe that fundamental-mode resonance frequencies of the heterostructure devices fall between the values of graphene and MoS2 devices. Quality (Q) factors of heterostructure resonators are lower than those of graphene but comparable to those of MoS2 devices, suggesting interface damping related to interlayer interactions in the van der Waals heterostructures. This study validates suspended at. layer heterostructures as an effective device platform and provides opportunities for exploiting mech. coupled effects and interlayer interactions in such devices.
- 10Rodriguez, A. W.; Capasso, F.; Johnson, S. G. The Casimir effect in microstructured geometries. Nat. Photonics 2011, 5, 211– 221, DOI: 10.1038/nphoton.2011.39Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXjvFKmu7c%253D&md5=4b65d5dcd90d67344fca892e33329c6eThe Casimir effect in microstructured geometriesRodriguez, Alejandro W.; Capasso, Federico; Johnson, Steven G.Nature Photonics (2011), 5 (4), 211-221CODEN: NPAHBY; ISSN:1749-4885. (Nature Publishing Group)A review. In 1948, Hendrik Casimir predicted that a generalized version of van der Waals forces would arise between two metal plates due to quantum fluctuations of the electromagnetic field. These forces become significant in micromech. systems at submicrometre scales, such as in the adhesion between movable parts. The Casimir force, through a close connection to classical photonics, can depend strongly on the shapes and compns. of the objects, stimulating a decades-long search for geometries in which the force behaves very differently from the monotonic attractive force first predicted by Casimir. Recent theor. and exptl. developments have led to a new understanding of the force in complex microstructured geometries, including through recent theor. predictions of Casimir repulsion between vacuum-sepd. metals, the stable suspension of objects and unusual non-additive and temp. effects, as well as exptl. observations of repulsion in fluids, non-additive forces in nanotrench surfaces and the influence of new material choices.
- 11Khosla, K. E.; Vanner, M. R.; Ares, N.; Laird, E. A. Displacemon electromechanics: how to detect quantum interference in a nanomechanical resonator. Phys. Rev. X 2018, 8 (2), 021052, DOI: 10.1103/PhysRevX.8.021052Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXltFSgt7w%253D&md5=3fd5697cc2c3db92cfb6e534558e2050Displacemon Electromechanics: How to Detect Quantum Interference in a Nanomechanical ResonatorKhosla, K. E.; Vanner, M. R.; Ares, N.; Laird, E. A.Physical Review X (2018), 8 (2), 021052CODEN: PRXHAE; ISSN:2160-3308. (American Physical Society)We introduce the "displacemon" electromech. architecture that comprises a vibrating nanobeam, e.g., a carbon nanotube, flux coupled to a superconducting qubit. This platform can achieve strong and even ultrastrong coupling, enabling a variety of quantum protocols. We use this system to describe a protocol for generating and measuring quantum interference between trajectories of a nanomech. resonator. The scheme uses a sequence of qubit manipulations and measurements to cool the resonator, to apply two effective diffraction gratings, and then to measure the resulting interference pattern. We demonstrate the feasibility of generating a spatially distinct quantum superposition state of motion contg. more than 106 nucleons using a vibrating nanotube acting as a junction in this new superconducting qubit configuration.
- 12O’Connell, A. D.; Hofheinz, M.; Ansmann, M.; Bialczak, R. C.; Lenander, M.; Lucero, E.; Neeley, M.; Sank, D.; Wang, H.; Weides, M.; Wenner, J.; Martinis, J. M.; Cleland, A. N. Quantum ground state and single-phonon control of a mechanical resonator. Nature 2010, 464 (7289), 697– 703, DOI: 10.1038/nature08967Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXjsFSqtbc%253D&md5=6d73284662e28f35a1d4e44f51031803Quantum ground state and single-phonon control of a mechanical resonatorO'Connell, A. D.; Hofheinz, M.; Ansmann, M.; Bialczak, Radoslaw C.; Lenander, M.; Lucero, Erik; Neeley, M.; Sank, D.; Wang, H.; Weides, M.; Wenner, J.; Martinis, John M.; Cleland, A. N.Nature (London, United Kingdom) (2010), 464 (7289), 697-703CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Quantum mechanics provides a highly accurate description of a wide variety of phys. systems. However, a demonstration that quantum mechanics applies equally to macroscopic mech. systems has been a long-standing challenge, hindered by the difficulty of cooling a mech. mode to its quantum ground state. The temps. required are typically far below those attainable with std. cryogenic methods, so significant effort has been devoted to developing alternative cooling techniques. Once in the ground state, quantum-limited measurements must then be demonstrated. Here, using conventional cryogenic refrigeration, we show that we can cool a mech. mode to its quantum ground state by using a microwave-frequency mech. oscillator-a quantum drum'-coupled to a quantum bit, which is used to measure the quantum state of the resonator. We further show that we can controllably create single quantum excitations (phonons) in the resonator, thus taking the first steps to complete quantum control of a mech. system.
- 13Aspelmeyer, M. In Gravitational quantum physics, or: How to avoid the appearance of the classical world in gravity experiments? 8th International Topical Meeting on Nanophotonics and Metamaterials; Seefeld-in-Tirol, Austria, 28–31 March 2022; Seefeld-in-Tirol, Austria, 2022.Google ScholarThere is no corresponding record for this reference.
- 14Fedotov, V. A.; Rose, M.; Prosvirnin, S. L.; Papasimakis, N.; Zheludev, N. I. Sharp trapped-mode resonances in planar metamaterials with a broken structural symmetry. Phys. Rev. Lett. 2007, 99, 147401, DOI: 10.1103/PhysRevLett.99.147401Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtFemsLbI&md5=6b4302eadb48ef45564f2a765da5ecd0Sharp Trapped-Mode Resonances in Planar Metamaterials with a Broken Structural SymmetryFedotov, V. A.; Rose, M.; Prosvirnin, S. L.; Papasimakis, N.; Zheludev, N. I.Physical Review Letters (2007), 99 (14), 147401/1-147401/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We report that a resonance response with a very high quality factor can be achieved in a planar metamaterial by introducing symmetry breaking in the shape of its structural elements, which enables excitation of trapped modes, i.e., modes that are weakly coupled to free space.
- 15Li, J.; Papas, D.; Liu, T.; Ou, J.-Y.; MacDonald, K. F.; Plum, E.; Zheludev, N. I. Thermal fluctuations of the optical properties of nanomechanical photonic metamaterials. Adv. Opt. Mater. 2022, 10, 2101591, DOI: 10.1002/adom.202101591Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XisFKg&md5=a92a7731a2426d88e3ef7914a69e5da2Thermal Fluctuations of the Optical Properties of Nanomechanical Photonic MetamaterialsLi, Jinxiang; Papas, Dimitrios; Liu, Tongjun; Ou, Jun-Yu; MacDonald, Kevin F.; Plum, Eric; Zheludev, Nikolay I.Advanced Optical Materials (2022), 10 (5), 2101591CODEN: AOMDAX; ISSN:2195-1071. (Wiley-VCH Verlag GmbH & Co. KGaA)The combination of optical and mech. resonances offers strong hybrid nonlinearities, bistability, and the ability to efficiently control the optical response of nanomech. photonic metamaterials with elec. and magnetic field. While optical resonances can be characterized in routine transmission and reflection expts., mapping the high-frequency mech. resonances of complex metamaterial structures is challenging. Here, it is reported that high-frequency time-domain fluctuations in the optical transmission and reflection spectra of nanomech. photonic metamaterials are directly linked to thermal motion of their components and can give information on the fundamental frequencies and damping of the mech. modes. This is demonstrated by analyzing time-resolved fluctuations in the transmission and reflection of dielec. and plasmonic nanomembrane metamaterials at room temp. and low ambient gas pressure. These measurements reveal complex mech. responses, understanding of which is essential for optimization of such functional photonic materials. At room temp. the magnitude of the obsd. metamaterial transmission and reflection fluctuations is of order 0.1% but may exceed 1% at optical resonances.
- 16Papas, D.; Ou, J.-Y.; Plum, E.; Zheludev, N. I. Optomechanical metamaterial nanobolometer. APL Photon. 2021, 6, 126110, DOI: 10.1063/5.0073583Google ScholarThere is no corresponding record for this reference.
- 17Hauer, B. D.; Doolin, C.; Beach, K. S. D.; Davis, J. P. A general procedure for thermomechanical calibration of nano/micro-mechanical resonators. Ann. Phys. - New York 2013, 339, 181– 207, DOI: 10.1016/j.aop.2013.08.003Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhvFSktr%252FJ&md5=a82af351d80f5764072b2c625f333b64A general procedure for thermomechanical calibration of nano/micro-mechanical resonatorsHauer, B. D.; Doolin, C.; Beach, K. S. D.; Davis, J. P.Annals of Physics (Amsterdam, Netherlands) (2013), 339 (), 181-207CODEN: APNYA6; ISSN:0003-4916. (Elsevier B.V.)We describe a general procedure to calibrate the detection of a nano/micro-mech. resonator's displacement as it undergoes thermal Brownian motion. A brief introduction to the equations of motion for such a resonator is presented, followed by a detailed derivation of the corresponding power spectral d. (PSD) function, which is identical in all situations aside from a system-dependent effective mass value. The effective masses for a no. of different resonator geometries are detd. using both finite element method (FEM) modeling and anal. calcns.
- 18Clerk, A. A.; Devoret, M. H.; Girvin, S. M.; Marquardt, F.; Schoelkopf, R. J. Introduction to quantum noise, measurement, and amplification. Rev. Mod. Phys. 2010, 82 (2), 1155– 1208, DOI: 10.1103/RevModPhys.82.1155Google ScholarThere is no corresponding record for this reference.
- 19Cleland, A. N. Foundations of nanomechanics: from solid-state theory to device applications; Springer Science & Business Media, 2013.Google ScholarThere is no corresponding record for this reference.
- 20Bokaian, A. Natural frequencies of beams under tensile axial loads. Journal of Sound and Vibration 1990, 142 (3), 481– 498, DOI: 10.1016/0022-460X(90)90663-KGoogle ScholarThere is no corresponding record for this reference.
- 21Yamada, S.; Schmid, S.; Larsen, T.; Hansen, O.; Boisen, A. Photothermal Infrared Spectroscopy of Airborne Samples with Mechanical String Resonators. Anal. Chem. 2013, 85 (21), 10531– 10535, DOI: 10.1021/ac402585eGoogle Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsFGkurjJ&md5=d10c4581bdf74f26051cf97bc561082aPhotothermal Infrared Spectroscopy of Airborne Samples with Mechanical String ResonatorsYamada, Shoko; Schmid, Silvan; Larsen, Tom; Hansen, Ole; Boisen, AnjaAnalytical Chemistry (Washington, DC, United States) (2013), 85 (21), 10531-10535CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)Micromech. photothermal IR spectroscopy is a promising technique, where absorption-related heating is detected by frequency detuning of microstring resonators. We present photothermal IR spectroscopy with mech. string resonators providing rapid identification of femtogram-scale airborne samples. Airborne sample material is directly collected on the microstring with an efficient nondiffusion limited sampling method based on inertial impaction. Resonance frequency shifts, proportional to the absorbed heat in the microstring, are recorded as monochromatic IR light is scanned over the mid-IR range. As a proof-of-concept, we sample and analyze polyvinylpyrrolidone (PVP) and the IR spectrum measured by photothermal spectroscopy matches the ref. IR spectrum measured by an FTIR spectrometer. We further identify the org. surface coating of airborne TiO2 nanoparticles with a total mass of 4 pg. With an estd. detection limit of 44 fg, the presented sensor demonstrates a new paradigm in ultrasensitive vibrational spectroscopy for identification of airborne species.
Cited By
This article is cited by 2 publications.
- Yanru Chen, Wei Chen, Xiaowen Xiang, Lianfu Deng, Jihong Qian, Wenguo Cui, Hao Chen. Pollen‐Inspired Shell–Core Aerosol Particles Capable of Brownian Motion for Pulmonary Vascularization. Advanced Materials 2023, 4 , 2207744. https://doi.org/10.1002/adma.202207744
- Yang Wang, Yujie Liu, Zeyu Zhong, Xiangliang Jin, Yan Peng, Jun Luo, Jun Yang. Gate-controlled silicon controlled rectifier with adjustable clamping voltage using a photoelectric mechanism. Optics Express 2023, 31
(1)
, 651. https://doi.org/10.1364/OE.475142
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 21 other publications.
- 1Chaste, J.; Eichler, A.; Moser, J.; Ceballos, G.; Rurali, R.; Bachtold, A. A nanomechanical mass sensor with yoctogram resolution. Nat. Nanotechnol. 2012, 7 (5), 301– 304, DOI: 10.1038/nnano.2012.421https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XkvVKitbc%253D&md5=9a8a31f058e85ec94624d906f5b0dd68A nanomechanical mass sensor with yoctogram resolutionChaste, J.; Eichler, A.; Moser, J.; Ceballos, G.; Rurali, R.; Bachtold, A.Nature Nanotechnology (2012), 7 (5), 301-304CODEN: NNAABX; ISSN:1748-3387. (Nature Publishing Group)Nanomech. resonators have been used to weigh cells, biomols. and gas mols., and to study basic phenomena in surface science, such as phase transitions and diffusion. These expts. all rely on the ability of nanomech. mass sensors to resolve small masses. Here, we report mass sensing expts. with a resoln. of 1.7 yg (1 yg = 10-24 g), which corresponds to the mass of one proton. The resonator is a carbon nanotube of length ∼150 nm that vibrates at a frequency of almost 2 GHz. This unprecedented level of sensitivity allows us to detect adsorption events of naphthalene mols. (C10H8), and to measure the binding energy of a xenon atom on the nanotube surface. These ultrasensitive nanotube resonators could have applications in mass spectrometry, magnetometry and surface science.
- 2Rugar, D.; Budakian, R.; Mamin, H. J.; Chui, B. W. Single spin detection by magnetic resonance force microscopy. Nature 2004, 430 (6997), 329– 332, DOI: 10.1038/nature026582https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXls1eqs78%253D&md5=289d1d5dc5543fd443fa420e5a96a6d1Single spin detection by magnetic resonance force microscopyRugar, D.; Budakian, R.; Mamin, H. J.; Chui, B. W.Nature (London, United Kingdom) (2004), 430 (6997), 329-332CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Magnetic resonance imaging (MRI) is well known as a powerful technique for visualizing subsurface structures with three-dimensional spatial resoln. Pushing the resoln. below 1 μm remains a major challenge, however, owing to the sensitivity limitations of conventional inductive detection techniques. Currently, the smallest vol. elements in an image must contain at least 1012 nuclear spins for MRI-based microscopy, or 107 electron spins for ESR microscopy. Magnetic resonance force microscopy (MRFM) was proposed as a means to improve detection sensitivity to the single-spin level, and thus enable three-dimensional imaging of macromols. (for example, proteins) with at. resoln. MRFM has also been proposed as a qubit readout device for spin-based quantum computers. Here we report the detection of an individual electron spin by MRFM. A spatial resoln. of 25 nm in one dimension was obtained for an unpaired spin in silicon dioxide. The measured signal is consistent with a model in which the spin is aligned parallel or anti-parallel to the effective field, with a rotating-frame relaxation time of 760 ms. The long relaxation time suggests that the state of an individual spin can be monitored for extended periods of time, even while subjected to a complex set of manipulations that are part of the MRFM measurement protocol.
- 3Feng, P. X.-L.; Young, D. J.; Zorman, C. A. MEMS/NEMS devices and applications. In Springer Handbook of Nanotechnology; Springer: 2017; pp 395– 429.There is no corresponding record for this reference.
- 4Van Thourhout, D.; Roels, J. Optomechanical device actuation through the optical gradient force. Nat. Photonics 2010, 4, 211– 217, DOI: 10.1038/nphoton.2010.724https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXktVWqurY%253D&md5=f02fb96b10b26d8b14c20ff27f470aeeOptomechanical device actuation through the optical gradient forceVan Thourhout, Dries; Roels, JorisNature Photonics (2010), 4 (4), 211-217CODEN: NPAHBY; ISSN:1749-4885. (Nature Publishing Group)A review. Optical forces are widely used to manipulate microparticles such as living cells, DNA and bacteria. The forces used in these optical tweezers originate from the strongly varying electromagnetic field in the focus of a high-power laser beam. This field gradient polarizes the particle, causing the pos. and neg. charged sides of the dipole to experience slightly different forces. It was recently realized that the strong field gradient in the near-field of guided wave structures can also be exploited for actuating optomech. devices, and initial theor. work in this area was followed rapidly by several exptl. demonstrations. The rapid developments in this field are summarized. The origin of the optical gradient force is discussed. Several exptl. demonstrations and approaches for enhancing the strength of the effect are discussed. Some of the possible applications of the effect are reviewed.
- 5Eggleton, B. J.; Poulton, C. G.; Rakich, P. T.; Steel, M. J.; Bahl, G. Brillouin integrated photonics. Nat. Photonics 2019, 13 (10), 664– 677, DOI: 10.1038/s41566-019-0498-z5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhs1Srtr7I&md5=56781cacb80828c34eedcdad6f2182b9Brillouin integrated photonicsEggleton, Benjamin J.; Poulton, Christopher G.; Rakich, Peter T.; Steel, Michael. J.; Bahl, GauravNature Photonics (2019), 13 (10), 664-677CODEN: NPAHBY; ISSN:1749-4885. (Nature Research)A recent renaissance in Brillouin scattering research has been driven by the increasing maturity of photonic integration platforms and nanophotonics. The result is a new breed of chip-based devices that exploit acousto-optic interactions to create lasers, amplifiers, filters, delay lines and isolators. Here, we provide a detailed overview of Brillouin scattering in integrated waveguides and resonators, covering key concepts such as the stimulation of the Brillouin process, in which the optical field itself induces acoustic vibrations, the importance of acoustic confinement, methods for calcg. and measuring Brillouin gain, and the diversity of materials platforms and geometries. Our Review emphasizes emerging applications in microwave photonics, signal processing and sensing, and concludes with a perspective for future directions.
- 6Zheludev, N. I.; Plum, E. Reconfigurable nanomechanical photonic metamaterials. Nat. Nanotechnol. 2016, 11, 16– 22, DOI: 10.1038/nnano.2015.3026https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XjslGqtA%253D%253D&md5=86be7d90217dd13c4aad5a8c2816b7a5Reconfigurable nanomechanical photonic metamaterialsZheludev, Nikolay I.; Plum, EricNature Nanotechnology (2016), 11 (1), 16-22CODEN: NNAABX; ISSN:1748-3387. (Nature Publishing Group)A review. The changing balance of forces at the nanoscale offers the opportunity to develop a new generation of spatially reconfigurable nanomembrane metamaterials in which electromagnetic Coulomb, Lorentz and Ampe´re forces, as well as thermal stimulation and optical signals, can be engaged to dynamically change their optical properties. Individual building blocks of such metamaterials, the metamols., and their arrays fabricated on elastic dielec. membranes can be reconfigured to achieve optical modulation at high frequencies, potentially reaching the gigahertz range. Mech. and optical resonances enhance the magnitude of actuation and optical response within these nanostructures, which can be driven by elec. signals of only a few volts or optical signals with power of only a few milliwatts. The authors envisage switchable, electrooptical, magnetooptical and nonlinear metamaterials that are compact and Si-nanofabrication-technol. compatible with functionalities surpassing those of natural media by orders of magnitude in some key design parameters.
- 7Karvounis, A.; Gholipour, B.; MacDonald, K. F.; Zheludev, N. I. Giant electro-optical effect through electrostriction in a nano-mechanical metamaterial. Adv. Mater. 2019, 31, 1804801, DOI: 10.1002/adma.201804801There is no corresponding record for this reference.
- 8Zhang, Q.; Plum, E.; Ou, J.-Y.; Pi, H.; Li, J.; MacDonald, K. F.; Zheludev, N. I. Electrogyration in metamaterials: Chirality and polarization rotatory power that depend on applied electric field. Adv. Opt. Mater. 2020, 2001826, DOI: 10.1002/adom.202001826There is no corresponding record for this reference.
- 9Ye, F.; Lee, J.; Feng, P. X.-L. Atomic layer MoS2 graphene van der Waals heterostructure nanomechanical resonators. Nanoscale 2017, 9 (46), 18208– 18215, DOI: 10.1039/C7NR04940D9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsF2rtbjO&md5=81a463ce9b33a257025ff19869628950Atomic layer MoS2-graphene van der Waals heterostructure nanomechanical resonatorsYe, Fan; Lee, Jaesung; Feng, Philip X.-L.Nanoscale (2017), 9 (46), 18208-18215CODEN: NANOHL; ISSN:2040-3372. (Royal Society of Chemistry)Heterostructures play significant roles in modern semiconductor devices and micro/nanosystems in a plethora of applications in electronics, optoelectronics, and transducers. While state-of-the-art heterostructures often involve stacks of cryst. epi-layers each down to a few nanometers thick, the intriguing limit would be hetero-at.-layer structures. Here we report the first exptl. demonstration of freestanding van der Waals heterostructures and their functional nanomech. devices. By stacking single-layer (1L) MoS2 on top of suspended single-, bi-, tri- and four-layer (1L to 4L) graphene sheets, we realize an array of MoS2-graphene heterostructures with varying thickness and size. These heterostructures all exhibit robust nanomech. resonances in the very high frequency (VHF) band (up to ≈100 MHz). We observe that fundamental-mode resonance frequencies of the heterostructure devices fall between the values of graphene and MoS2 devices. Quality (Q) factors of heterostructure resonators are lower than those of graphene but comparable to those of MoS2 devices, suggesting interface damping related to interlayer interactions in the van der Waals heterostructures. This study validates suspended at. layer heterostructures as an effective device platform and provides opportunities for exploiting mech. coupled effects and interlayer interactions in such devices.
- 10Rodriguez, A. W.; Capasso, F.; Johnson, S. G. The Casimir effect in microstructured geometries. Nat. Photonics 2011, 5, 211– 221, DOI: 10.1038/nphoton.2011.3910https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXjvFKmu7c%253D&md5=4b65d5dcd90d67344fca892e33329c6eThe Casimir effect in microstructured geometriesRodriguez, Alejandro W.; Capasso, Federico; Johnson, Steven G.Nature Photonics (2011), 5 (4), 211-221CODEN: NPAHBY; ISSN:1749-4885. (Nature Publishing Group)A review. In 1948, Hendrik Casimir predicted that a generalized version of van der Waals forces would arise between two metal plates due to quantum fluctuations of the electromagnetic field. These forces become significant in micromech. systems at submicrometre scales, such as in the adhesion between movable parts. The Casimir force, through a close connection to classical photonics, can depend strongly on the shapes and compns. of the objects, stimulating a decades-long search for geometries in which the force behaves very differently from the monotonic attractive force first predicted by Casimir. Recent theor. and exptl. developments have led to a new understanding of the force in complex microstructured geometries, including through recent theor. predictions of Casimir repulsion between vacuum-sepd. metals, the stable suspension of objects and unusual non-additive and temp. effects, as well as exptl. observations of repulsion in fluids, non-additive forces in nanotrench surfaces and the influence of new material choices.
- 11Khosla, K. E.; Vanner, M. R.; Ares, N.; Laird, E. A. Displacemon electromechanics: how to detect quantum interference in a nanomechanical resonator. Phys. Rev. X 2018, 8 (2), 021052, DOI: 10.1103/PhysRevX.8.02105211https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXltFSgt7w%253D&md5=3fd5697cc2c3db92cfb6e534558e2050Displacemon Electromechanics: How to Detect Quantum Interference in a Nanomechanical ResonatorKhosla, K. E.; Vanner, M. R.; Ares, N.; Laird, E. A.Physical Review X (2018), 8 (2), 021052CODEN: PRXHAE; ISSN:2160-3308. (American Physical Society)We introduce the "displacemon" electromech. architecture that comprises a vibrating nanobeam, e.g., a carbon nanotube, flux coupled to a superconducting qubit. This platform can achieve strong and even ultrastrong coupling, enabling a variety of quantum protocols. We use this system to describe a protocol for generating and measuring quantum interference between trajectories of a nanomech. resonator. The scheme uses a sequence of qubit manipulations and measurements to cool the resonator, to apply two effective diffraction gratings, and then to measure the resulting interference pattern. We demonstrate the feasibility of generating a spatially distinct quantum superposition state of motion contg. more than 106 nucleons using a vibrating nanotube acting as a junction in this new superconducting qubit configuration.
- 12O’Connell, A. D.; Hofheinz, M.; Ansmann, M.; Bialczak, R. C.; Lenander, M.; Lucero, E.; Neeley, M.; Sank, D.; Wang, H.; Weides, M.; Wenner, J.; Martinis, J. M.; Cleland, A. N. Quantum ground state and single-phonon control of a mechanical resonator. Nature 2010, 464 (7289), 697– 703, DOI: 10.1038/nature0896712https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXjsFSqtbc%253D&md5=6d73284662e28f35a1d4e44f51031803Quantum ground state and single-phonon control of a mechanical resonatorO'Connell, A. D.; Hofheinz, M.; Ansmann, M.; Bialczak, Radoslaw C.; Lenander, M.; Lucero, Erik; Neeley, M.; Sank, D.; Wang, H.; Weides, M.; Wenner, J.; Martinis, John M.; Cleland, A. N.Nature (London, United Kingdom) (2010), 464 (7289), 697-703CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Quantum mechanics provides a highly accurate description of a wide variety of phys. systems. However, a demonstration that quantum mechanics applies equally to macroscopic mech. systems has been a long-standing challenge, hindered by the difficulty of cooling a mech. mode to its quantum ground state. The temps. required are typically far below those attainable with std. cryogenic methods, so significant effort has been devoted to developing alternative cooling techniques. Once in the ground state, quantum-limited measurements must then be demonstrated. Here, using conventional cryogenic refrigeration, we show that we can cool a mech. mode to its quantum ground state by using a microwave-frequency mech. oscillator-a quantum drum'-coupled to a quantum bit, which is used to measure the quantum state of the resonator. We further show that we can controllably create single quantum excitations (phonons) in the resonator, thus taking the first steps to complete quantum control of a mech. system.
- 13Aspelmeyer, M. In Gravitational quantum physics, or: How to avoid the appearance of the classical world in gravity experiments? 8th International Topical Meeting on Nanophotonics and Metamaterials; Seefeld-in-Tirol, Austria, 28–31 March 2022; Seefeld-in-Tirol, Austria, 2022.There is no corresponding record for this reference.
- 14Fedotov, V. A.; Rose, M.; Prosvirnin, S. L.; Papasimakis, N.; Zheludev, N. I. Sharp trapped-mode resonances in planar metamaterials with a broken structural symmetry. Phys. Rev. Lett. 2007, 99, 147401, DOI: 10.1103/PhysRevLett.99.14740114https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtFemsLbI&md5=6b4302eadb48ef45564f2a765da5ecd0Sharp Trapped-Mode Resonances in Planar Metamaterials with a Broken Structural SymmetryFedotov, V. A.; Rose, M.; Prosvirnin, S. L.; Papasimakis, N.; Zheludev, N. I.Physical Review Letters (2007), 99 (14), 147401/1-147401/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We report that a resonance response with a very high quality factor can be achieved in a planar metamaterial by introducing symmetry breaking in the shape of its structural elements, which enables excitation of trapped modes, i.e., modes that are weakly coupled to free space.
- 15Li, J.; Papas, D.; Liu, T.; Ou, J.-Y.; MacDonald, K. F.; Plum, E.; Zheludev, N. I. Thermal fluctuations of the optical properties of nanomechanical photonic metamaterials. Adv. Opt. Mater. 2022, 10, 2101591, DOI: 10.1002/adom.20210159115https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XisFKg&md5=a92a7731a2426d88e3ef7914a69e5da2Thermal Fluctuations of the Optical Properties of Nanomechanical Photonic MetamaterialsLi, Jinxiang; Papas, Dimitrios; Liu, Tongjun; Ou, Jun-Yu; MacDonald, Kevin F.; Plum, Eric; Zheludev, Nikolay I.Advanced Optical Materials (2022), 10 (5), 2101591CODEN: AOMDAX; ISSN:2195-1071. (Wiley-VCH Verlag GmbH & Co. KGaA)The combination of optical and mech. resonances offers strong hybrid nonlinearities, bistability, and the ability to efficiently control the optical response of nanomech. photonic metamaterials with elec. and magnetic field. While optical resonances can be characterized in routine transmission and reflection expts., mapping the high-frequency mech. resonances of complex metamaterial structures is challenging. Here, it is reported that high-frequency time-domain fluctuations in the optical transmission and reflection spectra of nanomech. photonic metamaterials are directly linked to thermal motion of their components and can give information on the fundamental frequencies and damping of the mech. modes. This is demonstrated by analyzing time-resolved fluctuations in the transmission and reflection of dielec. and plasmonic nanomembrane metamaterials at room temp. and low ambient gas pressure. These measurements reveal complex mech. responses, understanding of which is essential for optimization of such functional photonic materials. At room temp. the magnitude of the obsd. metamaterial transmission and reflection fluctuations is of order 0.1% but may exceed 1% at optical resonances.
- 16Papas, D.; Ou, J.-Y.; Plum, E.; Zheludev, N. I. Optomechanical metamaterial nanobolometer. APL Photon. 2021, 6, 126110, DOI: 10.1063/5.0073583There is no corresponding record for this reference.
- 17Hauer, B. D.; Doolin, C.; Beach, K. S. D.; Davis, J. P. A general procedure for thermomechanical calibration of nano/micro-mechanical resonators. Ann. Phys. - New York 2013, 339, 181– 207, DOI: 10.1016/j.aop.2013.08.00317https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhvFSktr%252FJ&md5=a82af351d80f5764072b2c625f333b64A general procedure for thermomechanical calibration of nano/micro-mechanical resonatorsHauer, B. D.; Doolin, C.; Beach, K. S. D.; Davis, J. P.Annals of Physics (Amsterdam, Netherlands) (2013), 339 (), 181-207CODEN: APNYA6; ISSN:0003-4916. (Elsevier B.V.)We describe a general procedure to calibrate the detection of a nano/micro-mech. resonator's displacement as it undergoes thermal Brownian motion. A brief introduction to the equations of motion for such a resonator is presented, followed by a detailed derivation of the corresponding power spectral d. (PSD) function, which is identical in all situations aside from a system-dependent effective mass value. The effective masses for a no. of different resonator geometries are detd. using both finite element method (FEM) modeling and anal. calcns.
- 18Clerk, A. A.; Devoret, M. H.; Girvin, S. M.; Marquardt, F.; Schoelkopf, R. J. Introduction to quantum noise, measurement, and amplification. Rev. Mod. Phys. 2010, 82 (2), 1155– 1208, DOI: 10.1103/RevModPhys.82.1155There is no corresponding record for this reference.
- 19Cleland, A. N. Foundations of nanomechanics: from solid-state theory to device applications; Springer Science & Business Media, 2013.There is no corresponding record for this reference.
- 20Bokaian, A. Natural frequencies of beams under tensile axial loads. Journal of Sound and Vibration 1990, 142 (3), 481– 498, DOI: 10.1016/0022-460X(90)90663-KThere is no corresponding record for this reference.
- 21Yamada, S.; Schmid, S.; Larsen, T.; Hansen, O.; Boisen, A. Photothermal Infrared Spectroscopy of Airborne Samples with Mechanical String Resonators. Anal. Chem. 2013, 85 (21), 10531– 10535, DOI: 10.1021/ac402585e21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsFGkurjJ&md5=d10c4581bdf74f26051cf97bc561082aPhotothermal Infrared Spectroscopy of Airborne Samples with Mechanical String ResonatorsYamada, Shoko; Schmid, Silvan; Larsen, Tom; Hansen, Ole; Boisen, AnjaAnalytical Chemistry (Washington, DC, United States) (2013), 85 (21), 10531-10535CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)Micromech. photothermal IR spectroscopy is a promising technique, where absorption-related heating is detected by frequency detuning of microstring resonators. We present photothermal IR spectroscopy with mech. string resonators providing rapid identification of femtogram-scale airborne samples. Airborne sample material is directly collected on the microstring with an efficient nondiffusion limited sampling method based on inertial impaction. Resonance frequency shifts, proportional to the absorbed heat in the microstring, are recorded as monochromatic IR light is scanned over the mid-IR range. As a proof-of-concept, we sample and analyze polyvinylpyrrolidone (PVP) and the IR spectrum measured by photothermal spectroscopy matches the ref. IR spectrum measured by an FTIR spectrometer. We further identify the org. surface coating of airborne TiO2 nanoparticles with a total mass of 4 pg. With an estd. detection limit of 44 fg, the presented sensor demonstrates a new paradigm in ultrasensitive vibrational spectroscopy for identification of airborne species.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.nanolett.1c04900.
Metamaterial optical properties; Nanowire geometry and mechanical resonance frequencies; Nanowire effective medium parameters; Curve fitting algorithm (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.