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

Ultrasensitive nanomechanical mass sensor using hybrid opto-electromechanical systems

Open Access Open Access

Abstract

Nanomechanical resonators provide an unparalleled mass sensitivity sufficient to detect single biomolecules, viruses and nanoparticles. In this work we propose a scheme for mass sensing based on the hybrid opto-electromechanical system, where a mechanical resonator is coupled to an optical cavity and a microwave cavity simultaneously. When the two cavities are driven by two pump fields with proper frequencies and powers, a weak probe field is used to scan across the optical cavity resonance frequency. The mass of a single baculovirus landing onto the surface of the mechanical resonator can be measured by tracking the resonance frequency shift in the probe transmission spectrum before and after the deposition. We also propose a nonlinear mass sensor based on the measurement of the four-wave mixing (FWM) spectrum, which can be used to weigh a single 20-nm-diameter gold nanoparticle with sub-femtogram resolution.

© 2014 Optical Society of America

1. Introduction

Nanomechanical resonators are being actively investigated as sensitive mass sensors for applications such as biological and chemical sensing because of their minuscule masses (10−15 − 10−21kg), high resonance frequencies (kHz–GHz), and high quality factors (103 − 107) [1, 2]. The operation of a nanomechanical mass sensor employs tracking the resonance frequency shifts of a nanomechanical resonator due to mass changes caused by accreted particles. Mass sensitivity, an important parameter to relate frequency shifts with added masses, is proportional to the resonance frequency of the resonator and inversely proportional to its mass [3]. This has led to ever better levels of mass resolution–from femtogram (10−15g) [4] through attogram (10−18g) [5] and zeptogram (10−21g) [6] to yoctogram (10−24g) [7]. In the past decade, nanomechanical mass sensors based on cantilevers [8, 9], nanowires [1012], suspended microchannel resonators [13, 14] and carbon nanotubes [15, 16] have been used to weigh single molecules, viruses, and nanoparticles. Recently, Liu et al. demonstrated sub-pg mass sensing with an microtoroid optomechanical oscillator (OMO) supporting whispering-gallery mode (WGM) [17, 18]. Their analysis shows that femtogram level resolution is within reach even with relatively large OMOs. Furthermore, Shao et al. experimentally detected single nanoparticles and lentiviruses based on microtoroid microcavity by monitoring WGM broadening in microcavities, which is immune to both noise from the probe laser and environmental disturbances [19]. Li and Zhu theoretically proposed a scheme for nonlinear mass sensor with a toroidal microcavity optomechanical system [20, 21].

Cavity optomechanics, where the mechanical resonator is coupled to the electromagnetic cavity via radiation pressure force, is drawing considerable research interest in recent years [2224]. The optical response of optomechanical systems is modified due to the mechanical interactions, leading to effects such as normal mode splitting [25] and optomechanically induced transparency [2628]. Likewise, the optomechanical interaction allows for the readout of the mechanical motion with high sensitivity [29, 30], quantum ground state cooling of the mechanical resonators [31, 32], and quantum coherent coupling between light and mechanical resonator [33,34]. More recently, Addrews et al. have successfully coupled a micromechanical resonator to both a microwave cavity and an optical cavity [35], which consists of a hybrid opto-electromechanical system [3638]. The vibration of the resonator formed by a silicon nitride (Si3N4) membrane with low dissipation changes the resonance frequencies of both the optical and microwave cavities. When the cavities are driven by two pump fields with suitable frequencies and powers, respectively, and a weak probe field scans across the optical cavity resonance frequency, the resonance frequency of of the membrane can be obtained from the probe transmission spectrum. Consequently, the accreted mass deposited on the membrane can be weighed easily according to the frequency shift. The extremely narrow linewidth of the membrane [35] makes it a high-resolution mass sensor. Similar to the mass sensing demonstrated by Liu et al. [17, 18], the mass sensor proposed here is also based on monitoring the frequency shift of the mechanical resonator. However, only one laser provides the optical power both for actuation and high resolution monitoring of the mechanical resonant frequency in Liu’s experiment. In the hybrid opto-electromechanical system we consider here, two pump fields are used to drive the two cavities and one weak probe field is applied to track the resonance frequency of the oscillating mechanical mode by detecting the probe transmission spectrum. Such a pump-probe technique has been widely used in optomechanical experiments [2628, 35].

2. Model and theory

The hybrid opto-electromechanical system under consideration is consisted of an optical cavity and a microwave cavity which are coupled to a common mechanical resonator. The schematic of this system is shown in Fig. 1, where the mechanical resonator with resonance frequency ωm and damping rate γm is formed by a thin silicon nitride membrane that is free to vibrate [35]. The optical cavity with resonance frequency ωo is driven by a strong pump beam Eo with frequency Ωo and a weak probe beam Ep with frequency Ωp simultaneously. The microwave cavity with resonance frequency ωe, denoted by equivalent inductance L and equivalent capacitance C, is only driven by a strong pump beam Ee with frequency Ωe. When the membrane vibrates due to radiation pressure force, it moves along the optical intensity standing wave and changes the resonance frequency of the optical cavity [39]. At the same time, the vibrating membrane modulates the capacitance of the microwave cavity, and thus its resonance frequency. The interaction Hamiltonian between the mechanical mode and cavity modes is Hint = −h̄goaa(c + c) − h̄gebb(c + c), where operators a, b, and c are the annihilation operators of optical cavity, microwave cavity, and mechanical resonator, respectively. go (ge) is the single-photon coupling rate between the mechanical mode and the optical (microwave) cavity mode. In a rotating frame at the pump frequency Ωo and Ωe, the Hamiltonian of the hybrid system can be written as H = H0 + Hint + Hdrive, where

H0=h¯Δoaa+h¯Δebb+h¯ωmcc,Hdrive=ih¯κo,extEo(aa)+ih¯κe,extEe(bb)+ih¯κo,extEp(aeiδtaeiδt).
Here, Δo = ωo − Ωo and Δe = ωe − Ωe are the corresponding cavity-pump field detunings. Hdrive represents the interaction between the input fields and the cavity fields, where δ = Ωp − Ωo is the detuning between the probe laser beam and the pump laser beam. Eo, Ee, and Ep are the amplitudes of the applied fields, respectively, and they are related to their powers by |Eo|=2Poκo/h¯Ωo, |Ee|=2Peκe/h¯Ωe, and |Ep|=2Ppκo/h¯Ωp, where κo (κe) is the linewidth of the optical (microwave) cavity mode. κo,ext (κe,ext) describes the rate at which energy leaves the optical (microwave) cavity into propagating fields [35]. The quantum Langevin equations governing the system can be obtained by applying the Heisenberg equation and adding the corresponding damping and input noise terms for the cavity and mechanical modes,
a˙=i(ΔogoQ)aκoa+κo,ext(Eo+Epeiδt)+2κoain,
b˙=i(ΔegeQ)bκeB+κe,extEe+2κebin,
Q¨+γmQ˙+ωm2Q=2goωmaa+2geωmbb+ξ,
where Q is defined as Q = c + c, ain and bin are the input vacuum noise with zero mean value, which obey the correlation relation in the time domain
ain(t)ain(t)=bin(t)bin(t)=δ(tt),
ain(t)ain(t)=bin(t)bin(t)=0.
The mechanical mode is affected by a vicious force and by a Brownian stochastic force with zero mean value ξ that has the following correlation function [40]
ξ(t)ξ(t)=γmωmdω2πωeiω(tt)[1+coth(h¯ω2kBT)],
where kB is the Boltzmann constant and T is the temperature of the reservoir of the mechanical resonator. We derive the steady-state solution to Eqs. (2)(4) by setting all the time derivatives to zero, which are given by
as=κo,extEoκo+iΔo,bs=κe,extEeκe+iΔe,Qs=2ωm(go|as|2+ge|bs|2),
where Δ′o = ΔogoQs and Δ′e = ΔegeQs are the effective cavity detunings including radiation pressure effects. Subsequently, we follow the typical procedure and solve Eqs. (2)(4) perturbatively by rewriting each Heisenberg operator as the sum of its steady-state mean value and a small fluctuation with zero mean value,
a=as+δa,b=bs+δb,Q=Qs+δQ.
Substituting them into quantum Langevin equations (2)(4) and assuming that |as| ≫ 1 and |bs| ≫ 1, we can obtain the following linearized Heisenberg-Langevin equations:
δa˙=(κo+iΔo)δa+igoQsδa+igoasδQ+κo,extEpeiδt+2κoain,
δb˙=(κe+iΔe)δb+igeQsδb+igebsδQ+2κebin,
δQ¨+γmδQ˙+ωm2δQ=2ωmgoas(δa+δa)+2ωmgebs(δb+δb)+ξ,
where we have neglected the nonlinear terms δaδa, δbδb, δaδQ, and δbδQ which can result in some interesting phenomena of the optomechanical systems, such as second and higher-order sideband [41]. In the following, since the drives are classical coherent fields, we will identify all operators with their expectation values, and drop the quantum and thermal noise terms [26]. Then the linearized Langevin equations can be written as:
δa˙=(κo+iΔo)δa+igoQsδa+igoasδQ+κo,extEpeiδt,
δb˙=(κe+iΔe)δb+igeQsδb+igebsδQ,
δQ¨+γmδQ˙+ωm2δQ=2ωmgoas(δa+δa)+2ωmgebs(δb+δb),
In order to solve Eqs (13)(15), we make the ansatz [42] 〈δa〉 = a+eiδt + aeiδt, 〈δb〉 = b+eiδt + beiδt, and 〈Q〉 = Q+eiδt + Qeiδt. Upon substituting the above ansatz into Eqs. (13)(15), we can obtain the following solution
a+=κo,extEpκo+iΔoiδigo2nof(δ)κo,extEp(κo+iΔoiδ)2,
a=1f(δ)*(κo+iΔo)2igo2κo,extEo2κo,extEp(κo+iδ)2+Δo2,
where
f(δ)=2Δogo2no(κoiδ)2+Δo2+2Δege2ne(κeiδ)2+Δe2ωm2δ2iδγmωm.
Here, no = |as|2 and ne = |bs|2, approximately equal to the number of pump photons in each cavity, are determined by the following coupled equations
no=κo,extEo2κo2+[Δo2go/ωm(gono+gene)]2,
ne=κe,extEe2κe2+[Δe2ge/ωm(gono+gene)]2.
The output field from the cavity can be obtained by using the standard input-output theory [43] aout(t)=ain(t)κexta(t), where aout (t) is the output field operator. Considering the output field of the optical cavity, we have
aout(t)=(Eoκo,extas)eiΩot+(Epκo,exta+)ei(δ+Ωo)tκo,extaei(δΩo)t=(Eoκo,extas)eiΩot+(Epκo,exta+)eiΩptκo,extaei(2ΩoΩp)t.

 figure: Fig. 1

Fig. 1 Schematic diagram of the hybrid opto-electromechanical system. A mechanical resonator c couples to both an optical cavity modes a and a microwave cavity mode b denoted by the equivalent inductance L and equivalent capacitance C. The optical cavity is driven by a strong pump beam Eo in the simultaneous presence of a weak probe beam Ep while the microwave cavity is only driven by a pump beam Ee.

Download Full Size | PDF

We can see from Eq. (21) that the output field contains two input components (Ωo and Ωp) and one generated four-wave mixing (FWM) component at the frequency 2Ωo − Ωp. FWM is a third-order nonlinear process [42], where two pump photons at the frequency Ωo are converted, due to the optomechanical intercation between photons and phonons, into one idler and one probe photon at the frequencies Ωidler and Ωp, respectively, in accordance with energy conversation: Ωidler = 2Ωo − Ωp. This process has been experimentally observed in an ultrahigh-Q toroid microcavity [44] and theoretically investigated in a single-mode optomechanical system [45]. Based on the current experimental conditions, FWM can be observed in the hybrid opto-electromechanical systems. The transmission of the probe beam, defined by the ratio of the output and input field amplitudes at the probe frequency [26], is then given by

t(Ωp)=Epκo,exta+Ep=1[κo,extκo+iΔoiδ1f(δ)igo2noκo,ext(κo+iΔoiδ)2].
Likewise, the FWM intensity in terms of the probe field can be defined as
FWM=|κo,extaEp|2=|1f(δ)*(κo+iΔo)2igo2κo,ext2Eo2(κo+iδ)2+iΔo2|2.
In the following, we would propose a scheme for mass sensing based on the measurement of the probe transmission and FWM spectrum, where the resonance frequency of the membrane can be determined. The particles depositing onto the surface of the membrane will result in the resonant frequency shift. The relationship between the frequency shift Δω with the deposited mass Δm is given by [3]
Δm=2meffωmΔω=1Δω,
where = (−2meff/ωm)−1 is defined as the mass responsivity.

3. Results and discussion

For illustration of our numerical results, we choose the experimentally realizable hybrid opto-electromechanical system. The parameters used in the simulation are [31, 35]: ωo = 2π × 282 THz, ωe = 2π × 7.1 GHz, κo = 2π × 1.65 MHz, κe = 2π × 1.6 MHz, κo,ext = 0.76κo, κe,ext = 0.11κe, go = 2π × 27 Hz, ge = 2π × 2.7 Hz, ωm = 2π × 5.6 MHz, γm = 2π × 4 Hz, and the effective mass of the resonator is approximately meff = 45 pg. We take baculovirus and gold nanoparticle (the density of Au is ρAu =19300 kg/m3) as deposition sample, respectively. It should be noted that the hybrid system has to operate under cryogenic temperature to keep the superconducting circuitry working.

Mass sensing is based on detecting the frequency shift of the mechanical resonator before and after the adsorption of the accreted mass. Firstly, we would present a scheme for measuring the resonance frequency of the mechanical resonator based on this hybrid opto-electromechanical system. Figure 2(a) plots the transmission spectrum of the probe field as a function of the probe-cavity detuning Δp with Δo = Δe = 0 and Po = Pe = 0.01 μW. It can be seen clearly from this figure that there is a broad transmission dip when the probe field is resonant with the optical cavity, which corresponds to the cavity absorption. Furthermore, two sharp sideband peaks, representing the resonant amplification and absorption of the mechanical mode, locate exactly at Δp = ±ωm. The spectral width of the sideband peaks is the mechanical damping rate γm/2π = 4 Hz, which can be seen clearly from the enlarged image in Fig. 2(b). This extremely narrow spectral linewidth is beneficial in resolving the frequency shifts due to accreted mass. Therefore, Fig. 2 provides us an convenient method to measure the resonance frequency of the mechanical resonator. The measurement process can be illustrated as follows. (1) We apply a strong optical pump field and a strong microwave pump field to the respective cavity, and fix the pump fields at the resonance of the cavity frequencies (Δo = Δe = 0); (2) we then apply another weak probe field to the optical cavity, and scan the probe frequency across the cavity frequency. By detecting the probe transmission spectrum, one can easily obtain the resonance frequency of the mechanical resonator. The physical mechanism for this phenomenon can be understood as a result of the radiation pressure force oscillating at the beat frequency δ between the optical pump and probe fields. If δ is close to the resonance frequency ωm, the mechanical resonator starts to oscillate coherently. The induced motion leads to a mechanical sideband on the pump field that can interfere with the probe field and hence modifies the probe transmission spectrum.

 figure: Fig. 2

Fig. 2 (a) The probe transmission |t|2 as a function of the probe-cavity detuning Δp with Δo = Δe = 0 and Po = Pe = 0.01 μW, where two sideband peaks locate exactly at Δp = ±ωm. (b) The enlarged sideband peaks in (a), and the spectral width is the mechanical damping rate γm/2π. The other parameters used are ωo = 2π × 282 THz, ωe = 2π × 7.1 GHz, κo = 2π × 1.65 MHz, κe = 2π × 1.6 MHz, κo,ext = 0.76κo, κe,ext = 0.11κe, go = 2π × 27 Hz, ge = 2π × 2.7 Hz, ωm = 2π × 5.6 MHz, and γm = 2π × 4 Hz.

Download Full Size | PDF

After the resonance frequency of the mechanical resonator is determined via the probe transmission spectrum, we can subsequently measure the frequency shift caused by an object landing on the resonator. The frequency shift depends on the mass and position of the deposited object, so trapping an object at a known position would allow its mass to be determined directly. We assume for simplicity that the relationship between the accreted mass and the frequency shift satisfies Eq. (24) [7]. If a single baculovirus lands onto the surface of the mechanical resonator, the total mass of the resonator would be increased, resulting in reduction of the resonance frequency. Figure 3 plots the probe transmission spectrum as a function of the probe-cavity detuning Δp before (solid curve) and after (dashed curve) the adsorption of a single baculovirus in the vicinity of the mechanical resonance frequency. We can see that the resonance frequency shift Δω = −2π × 93 Hz can be resolved in the transmission spectrum due to the increased mass of the resonator. Based on the resonance frequency shift and using Eq. (24), we found the added mass to be 1.5 fg (1 fg = 10−15 g), about the mass of a single baculovirus [46]. Therefore, the hybrid opto-electromechanical system studied here could be used to weigh the mass of a single virus. Though the principle of mass sensing we propose is simple and feasible, it remains challenging to realize such a mass sensor in experiments. For example, the baculovirus and the mechanical resonator have to be dealt with by some specific process before the adsorption of the baculovirus [46]. In addition, mass responsivity is an important parameter to evaluate the performance of the mechanical resonator for mass sensing. The inset of Fig. 3 shows the direct linear relationship between the resonance frequency shifts and the number of the baculovirus landing on the mechanical resonator. Such a linear relationship has been verified by various experiments [6, 8]. The negative slope of the line gives the mass responsivity of the resonator. Both smaller mass and higher resonance frequency of the resonator are crucial in obtaining higher mass responsivity. Sun and Zheng et al. have recently designed and experimentally demonstrated femtogram L3-nanobeam optomechanical cavities by embedding a doubly clamped nanomechanical double-beam resonator with mass as small as 25 fg in a finely tuned two-dimensional (2D) photonic crystal (PC) slab, where optical transduction of the fundamental flexural mode around 1 GHz was demonstrated [4749]. They also pointed out that such femtogram-mass, high-mechanical resonance frequency structures could be used as ultrasensitive sensors of mass, force and displacement.

 figure: Fig. 3

Fig. 3 The simulation results of the probe transmission |t|2 versus the probe-cavity detuning Δp before and after adsorption of a single baculovirus on the mechanical resonator. The frequency shift of Δω/2π = 93 Hz can be easily resolved in the spectrum. Here, we have used the left peak in Fig. 2(b) to demonstrate the validity of our proposed mass sensing scheme. The inset plots the frequency shift as a function of the number of the virus adsorbed on the resonator. Other parameters are the same as in figure 2.

Download Full Size | PDF

Furthermore, the hybrid opto-electromechanical system can be employed as a nonlinear mass sensor based on the third-order nonlinear optical effect. Figure 4 plots the FWM spectrum as a function of the probe-cavity detuning with Δo = Δe = 0 before and after a binding event of a gold nanoparticle of 20 nanometers in diameter. Similarly, a narrow peak appears at Δp = ωm due to the quantum interference effect, which can be used to measure the resonance frequency of the mechanical resonator. After the adsorption of a single nanoparticle, there is a frequency shift Δω = −2π × 5 Hz and the red dashed peak locates at ωm + Δω. According to Eq. (24), we can obtain the mass of the single 20-nm-diameter nanoparticle: Δm=2meffωmΔω=80.4ag (1 ag =10−18 g), which is close to the mass calculated by the density and volume of the gold nanoparticle. Compared with the traditional mass spectrometry, the nonlinear mass sensor based on the four-wave mixing effect has some obvious advantages. First, the particles needn’t to be ionized like traditional mass spectrometer [50], and their masses can be measured from the FWM spectrum conveniently. Second, the simultaneous presence of the pump and probe beams generates a beat wave to drive the mechanical resonator. Therefore, resonators with both high and low resonance frequency are suited for this mass sensing scheme. Third, the use of nonlinear optical spectrum may offer better performance over the linear optical spectrum in the presence of detection noise [51].

 figure: Fig. 4

Fig. 4 Nonlinear probe transmission spectrum (FWM) as a function of the probe-cavity detuning Δp before and after a binding event of a single 20-nm-diameter gold nanoparticle. The other parameters used are ωo = 2π × 282 THz, ωe = 2π × 7.1 GHz, κo = 2π × 1.65 MHz, κe = 2π × 1.6 MHz, κo,ext = 0.76κo, κe,ext = 0.11κe, go = 2π × 27 Hz, ge = 2π × 2.7 Hz, ωm = 2π × 5.6 MHz, γm = 2π × 4 Hz, Δo = Δe = 0, and Po = Pe = 0.01 μW.

Download Full Size | PDF

Finally, it should be pointed out that we do not take into account the noise processes in our theoretical simulations. Actually, there are various noise sources which impose the ultimate mass sensitivity limits for the resonator, such as thermomechanical noise generated by the internal loss mechanisms in the resonator, adsorption-desorption noise from residual gas molecules [3, 52], and detection noise in the readout circuitry [51]. In the hybrid opto-electromechanical system we study here, the thermal noise of the mechanical motion is the dominant noise source. If the experiment could be done at the dilution refrigerator temperature of 40 millikelvin, the thermal noise would be greatly eliminated. In addition, added vibrational noise can be reduced by using mechanical resonators with higher quality factors [35].

4. Conclusion

In conclusion, we have theoretically demonstrated that the hybrid opto-electromechanical system, consisted of an optical cavity and a microwave cavity coupled to a common mechanical resonator, can be employed as an ultrasensitive mass sensor. Due to the quantum interference between the mechanical mode and the beat of the two optical fields, the resonance frequency of the mechanical resonator can be determined from the probe transmission spectrum as well as the FWM spectrum. Therefore, the mass of the accreted particles such as single viruses and gold nanoparticles landing onto the resonator can be obtained according to the relationship between the added mass and the corresponding frequency shifts. The scheme proposed here could be achievable in current experiments.

Acknowledgments

The authors gratefully acknowledge support from National Natural Science Foundation of China (Grant Nos. 11304110 and 11274230), Jiangsu Natural Science Foundation (Grant No. BK20130413), and Natural Science Foundation of the Jiangsu Higher Education Institutions of China (Grant No. 13KJB140002).

References and links

1. K. C. Schwab and M. L. Roukes, “Putting mechanics into quantum mechanics,” Phys. Today 58, 36–42 (2005). [CrossRef]  

2. J. L. Arlett, E. B. Myers, and M.L. Roukes, “Comparative advantages of mechanical biosensors,” Nature Nanotech. 6, 203–215 (2011). [CrossRef]  

3. K. L. Ekinci, Y. T. Tang, and M. L. Roukes, “Ultimate limits to inertial mass sensing based upon nanoelectromechanical systems,” J. Appl. Phys. 95, 2682–2689 (2004). [CrossRef]  

4. N. V. Lavrik and P. G. Datskos, “Femtogram mass detection using photothermally actuated nanomechanical resonators,” Appl. Phys. Lett. 82, 2697–2699 (2003). [CrossRef]  

5. B. Ilic, H. G. Craighead, S. Krylov, W. Senaratne, C. Ober, and P. Neuzil, “Attogram detection using nanoelectromechanical oscillators,” J. Appl. Phys. 95, 3694–3703 (2004). [CrossRef]  

6. Y. T. Yang, C. Callegari, X. L. Feng, K. L. Ekinci, and M. L. Roukes, “Zeptogram-scale nanomechanical mass sensing,” Nano Lett. 6, 583–586 (2006). [CrossRef]   [PubMed]  

7. J. Chaste, A. Eichler, J. Moser, G. Ceballos, R. Rurali, and A. Bachtold, “A nanomechanical mass sensor with yoctogram resolution,” Nature Nanotech. 7, 301–304 (2012). [CrossRef]  

8. A. Gupta, D. Akin, and R. Bashir, “Single virus particle mass detection using microresonators with nanoscale thickness,” Appl. Phys. Lett. 84, 1976–1978 (2004). [CrossRef]  

9. M. Li, H. X. Tang, and M. L. Roukes, “Ultra-sensitive NEMS-based cantilevers for sensing, scanned probe and very high-frequency applications,” Nature Nanotech. 2, 114–120 (2007). [CrossRef]  

10. X. L. Feng, R. He, P. D. Yang, and M. L. Roukes, “Very high frequency silicon nanowire electromechanical resonators,” Nano Lett. 7, 1953–1959 (2007). [CrossRef]  

11. A. K. Naik, M. S. Hanay, W. K. Hiebert, X. L. Feng, and M. L. Roukes, “Towards single-molecule nanomechanical mass spectrometry,” Nature Nanotech. 4, 445–450 (2009). [CrossRef]  

12. E. Gil-Santos, D. Ramos, J. Martínez, M. Fernández-Regúlez, R. García, Á. S. Paulo, M. Calleja, and J. Tamayo, “Nanomechanical mass sensing and stiffness spectrometry based on two-dimensional vibrations of resonant nanowires,” Nature Nanotech. 5, 641–645 (2010). [CrossRef]  

13. T. P. Burg, M. Godin, S. M. Knudsen, W. Shen, G. Carlson, J. S. Foster, K. Babcock, and S. R. Manalis, “Weighing of biomolecules, single cells and single nanoparticles in fluid,” Nature (London) 446, 1066–1069 (2007). [CrossRef]  

14. S. Olcuma, N. Cermak, S. C. Wasserman, K. S. Christine, H. Atsumi, K. R. Payer, W. Shen, J. Lee, A. M. Belcher, S. N. Bhati, and S. R. Manalis, “Weighing nanoparticles in solution at the attogram scale,” Proc. Natl. Acd. Sci. U.S.A. 111, 1310–1315 (2014). [CrossRef]  

15. B. Lassagne, D. Garcia-Sanchez, A. Aguasca, and A. Bachtold, “Ultrasensitive mass sensing with a nanotube electromechanical resonator,” Nano Lett. 8, 3735–3738 (2008). [CrossRef]   [PubMed]  

16. K. Jensen, K. Kim, and A. Zettl, “An atomic-resolution nanomechanical mass sensor,” Nature Nanotech. 3, 533–537 (2008). [CrossRef]  

17. F. Liu and M. Hossein-Zadeh, “Mass sensing with optomechanical oscillation,” IEEE Sensors 13, 146–147 (2013). [CrossRef]  

18. F. Liu, S. Alaie, Z. C. Leseman, and M. Hossein-Zadeh, “Sub-pg mass sensing and measurement with an optomechanical oscillator,” Opt. Express 21, 19555–19567 (2013). [CrossRef]   [PubMed]  

19. L. Shao, X.-F. Jiang, X.-C. Yu, B.-B. Li, W. R. Clements, F. Vollmer, W. Wang, Y.-F. Xiao, and Q. Gong, “Detection of single nanoparticles and lentiviruses using microcavity resonance broadening,” Adv. Mater. 25(39), 5616–5620 (2013). [CrossRef]   [PubMed]  

20. J.-J. Li and K.-D. Zhu, “Nonlinear optical mass sensor with an optomechanical microresonator,” Appl. Phys. Lett. 101, 141905 (2012). [CrossRef]  

21. J.-J. Li and K.-D. Zhu, “All-optical mass sensing with coupled mechanical resonator systems,” Phys. Rep. 525, 223–254 (2013). [CrossRef]  

22. T. J. Kippenberg and K. J. Vahala, “Cavity optomechanics: back-action at the mesoscale,” Science 321, 1172–1176 (2008). [CrossRef]   [PubMed]  

23. F. Marquardt and S. M. Girvin, “Optomechanics,” Physics 2, 40 (2009). [CrossRef]  

24. M. Aspelmeyer, P. Meystre, and K. Schwab, “Quantum optomechanics,” Phys. Today 65, 29–35 (2012). [CrossRef]  

25. S. Gröblacher, K. Hammerer, M. R. Vanner, and M. Aspelmeyer, “Observation of strong coupling between a micromechanical resonator and an optical cavity field,” Nature (London) 460, 724–727 (2009). [CrossRef]  

26. S. Weis, R. Rivière, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, and T. J. Kippenberg, “Optomechanically induced transparency,” Science 330, 1520–1523 (2010). [CrossRef]   [PubMed]  

27. A. H. Safavi-Naeini, T. P. Mayer Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature (London) 472, 69–73 (2011). [CrossRef]  

28. J. D. Teufel, D. Li, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, and R. W. Simmonds, “Circuit cavity electromechanics in the strong-coupling regime,” Nature (London) 471, 204–208 (2011). [CrossRef]  

29. O. Arcizet, P.-F. Cohadon, T. Briant, M. Pinard, A. Heidmann, J.-M. Mackowski, C. Michel, L. Pinard, O. Francais, and L. Rousseau, “High-sensitivity optical monitoring of a micromechanical resonator with a quantum-limited optomechanical sensor,” Phys. Rev. Lett. 97, 133601 (2006). [CrossRef]   [PubMed]  

30. J. D. Teufel, T. Donner, M. A. Castellanos-Beltran, J. W. Harlow, and K. W. Lehnert, “Nanomechanical motion measured with an imprecision below that at the standard quantum limit,” Nature Nanotech. 4, 820–823 (2009). [CrossRef]  

31. J. D. Teufel, T. Donner, D. Li, J. W. Harlow, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W. Simmonds, “Sideband cooling of micromechanical motion to the quantum ground state,” Nature (London) 475, 359–363 (2011). [CrossRef]  

32. J. Chan, T. P. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer, and O. Painter, “Laser cooling of a nanomechanical oscillator into its quantum ground state,” Nature (London) 478, 89–92 (2011). [CrossRef]  

33. E. Verhagen, S. Deléglise, S. Weis, A. Schliesser, and T. J. Kippenberg, “Quantum-coherent coupling of a mechanical oscillator to an optical cavity mode,” Nature (London) 482, 63–67 (2012). [CrossRef]  

34. T. A. Palomaki, J. W. Harlow, J. D. Teufel, R. W. Simmonds, and K. W. Lehnert, “Coherent state transfer between itinerant microwave fields and a mechanical oscillator,” Nature (London) 495, 210–214 (2013). [CrossRef]  

35. R. W. Andrews, R. W. Peterson, T. P. Purdy, K. Cicak, R. W. Simmonds, C. A. Regal, and K. W. Lehnert, “Bidirectional and efficient conversion between microwave and optical light,” Nature Phys. 10, 321–326 (2014). [CrossRef]  

36. C. A. Regal and K. W. Lehnert, “From cavity electromechanics to cavity optomechanics,” J. Phys. Conf. Ser. 264, 012025 (2011). [CrossRef]  

37. X.-Y. Lü, W.-M. Zhang, S. Ashhab, Y. Wu, and F. Nori, “Quantum-criticality-induced strong Kerr nonlinearities in optomechanical systems,” Sci. Rep. 3, 2943 (2013). [CrossRef]   [PubMed]  

38. K. N. Qu and G. S. Agarwal, “Phonon-mediated electromagnetically induced absorption in hybrid opto-electromechanical systems,” Phys. Rev. A 87, 031802(R) (2013). [CrossRef]  

39. J. D. Thompson, B. M. Zwickl, A. M. Jayich, Florian Marquardt, S. M. Girvin, and J. G. E. Harris, “Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane,” Nature (London) 452, 72–75 (2008). [CrossRef]  

40. 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]  

41. H. Xiong, L.-G. Si, A.-S. Zheng, X. X. Yang, and Y. Wu, “Higher-order sidebands in optomechanically induced transparency,” Phys. Rev. A 86, 013815 (2012). [CrossRef]  

42. R. W. Boyd, Nonlinear Optics (Academic, 2008).

43. C. W. Gardiner and P. Zoller, Quantum Noise (Springer) (2004).

44. T. J. Kippenberg, S. M. Spillane, and K. J. Vahala, “Kerr-nonlinearity optical parametric oscillation in an ultrahigh-Q toroid microcavity,” Phys. Rev. Lett. 93, 083904 (2004). [CrossRef]   [PubMed]  

45. S. Huang and G. S. Agarwal, “Normal-mode splitting and antibunching in Stokes and anti-Stokes processes in cavity optomechanics: Radiation-pressure-induced four-wave-mixing cavity optomechanics,” Phys. Rev. A 81, 033830 (2010). [CrossRef]  

46. B. Ilic, Y. Yang, and H. G. Craighead, “Virus detection using nanoelectromechanical devices,” Appl. Phys. Lett. 85, 2604–2606 (2004). [CrossRef]  

47. X. Sun, J. Zheng, M. Poot, C. W. Wong, and H. X. Tang, “Femtogram doubly clamped nanomechanical resonators embedded in a high-Q two-dimensional photonic crystal nanocavity,” Nano Lett. 12, 2299 (2012). [CrossRef]   [PubMed]  

48. J. Zheng, X. Sun, M. Poot, Y. Li, A. Dadgar, H. X. Tang, and C. W. Wong, “Dispersive coupling and optimization of femtogram L3-nanobeam optomechanical cavities,” Frontiers in Optics (2012).

49. J. Zheng, X. Sun, Y. Li, M. Poot, A. Dadgar, N. N. Shi, W. H. P. Pernice, H. X. Tang, and C. W. Wong, “Femtogram dispersive L3-nanobeam optomechanical cavities: design and experimental comparison,” Opt. Express 20, 26486–26498 (2012). [CrossRef]   [PubMed]  

50. A. Boisen, “Nanoelectromechanical systems: Mass spec goes nanomechanical,” Nature Nanotech. 4, 404–405 (2009). [CrossRef]  

51. Z. Yie, M. A. Zielke, C. B. Burgner, and K. L. Turner, “Comparison of parametric and linear mass detection in the presence of detection noise,” J. Micromech. Microeng. 21, 025027 (2011). [CrossRef]  

52. A. N. Cleland and M. L. Roukes, “Noise processes in nanomechanical resonators,” J. Appl. Phys. 92, 2758–2769 (2002). [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 Schematic diagram of the hybrid opto-electromechanical system. A mechanical resonator c couples to both an optical cavity modes a and a microwave cavity mode b denoted by the equivalent inductance L and equivalent capacitance C. The optical cavity is driven by a strong pump beam Eo in the simultaneous presence of a weak probe beam Ep while the microwave cavity is only driven by a pump beam Ee.
Fig. 2
Fig. 2 (a) The probe transmission |t|2 as a function of the probe-cavity detuning Δp with Δo = Δe = 0 and Po = Pe = 0.01 μW, where two sideband peaks locate exactly at Δp = ±ωm. (b) The enlarged sideband peaks in (a), and the spectral width is the mechanical damping rate γm/2π. The other parameters used are ωo = 2π × 282 THz, ωe = 2π × 7.1 GHz, κo = 2π × 1.65 MHz, κe = 2π × 1.6 MHz, κo,ext = 0.76κo, κe,ext = 0.11κe, go = 2π × 27 Hz, ge = 2π × 2.7 Hz, ωm = 2π × 5.6 MHz, and γm = 2π × 4 Hz.
Fig. 3
Fig. 3 The simulation results of the probe transmission |t|2 versus the probe-cavity detuning Δp before and after adsorption of a single baculovirus on the mechanical resonator. The frequency shift of Δω/2π = 93 Hz can be easily resolved in the spectrum. Here, we have used the left peak in Fig. 2(b) to demonstrate the validity of our proposed mass sensing scheme. The inset plots the frequency shift as a function of the number of the virus adsorbed on the resonator. Other parameters are the same as in figure 2.
Fig. 4
Fig. 4 Nonlinear probe transmission spectrum (FWM) as a function of the probe-cavity detuning Δp before and after a binding event of a single 20-nm-diameter gold nanoparticle. The other parameters used are ωo = 2π × 282 THz, ωe = 2π × 7.1 GHz, κo = 2π × 1.65 MHz, κe = 2π × 1.6 MHz, κo,ext = 0.76κo, κe,ext = 0.11κe, go = 2π × 27 Hz, ge = 2π × 2.7 Hz, ωm = 2π × 5.6 MHz, γm = 2π × 4 Hz, Δo = Δe = 0, and Po = Pe = 0.01 μW.

Equations (24)

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

H 0 = h ¯ Δ o a a + h ¯ Δ e b b + h ¯ ω m c c , H drive = i h ¯ κ o , ext E o ( a a ) + i h ¯ κ e , ext E e ( b b ) + i h ¯ κ o , ext E p ( a e i δ t a e i δ t ) .
a ˙ = i ( Δ o g o Q ) a κ o a + κ o , ext ( E o + E p e i δ t ) + 2 κ o a in ,
b ˙ = i ( Δ e g e Q ) b κ e B + κ e , ext E e + 2 κ e b in ,
Q ¨ + γ m Q ˙ + ω m 2 Q = 2 g o ω m a a + 2 g e ω m b b + ξ ,
a in ( t ) a in ( t ) = b in ( t ) b in ( t ) = δ ( t t ) ,
a in ( t ) a in ( t ) = b in ( t ) b in ( t ) = 0 .
ξ ( t ) ξ ( t ) = γ m ω m d ω 2 π ω e i ω ( t t ) [ 1 + coth ( h ¯ ω 2 k B T ) ] ,
a s = κ o , ext E o κ o + i Δ o , b s = κ e , ext E e κ e + i Δ e , Q s = 2 ω m ( g o | a s | 2 + g e | b s | 2 ) ,
a = a s + δ a , b = b s + δ b , Q = Q s + δ Q .
δ a ˙ = ( κ o + i Δ o ) δ a + i g o Q s δ a + i g o a s δ Q + κ o , ext E p e i δ t + 2 κ o a in ,
δ b ˙ = ( κ e + i Δ e ) δ b + i g e Q s δ b + i g e b s δ Q + 2 κ e b in ,
δ Q ¨ + γ m δ Q ˙ + ω m 2 δ Q = 2 ω m g o a s ( δ a + δ a ) + 2 ω m g e b s ( δ b + δ b ) + ξ ,
δ a ˙ = ( κ o + i Δ o ) δ a + i g o Q s δ a + i g o a s δ Q + κ o , ext E p e i δ t ,
δ b ˙ = ( κ e + i Δ e ) δ b + i g e Q s δ b + i g e b s δ Q ,
δ Q ¨ + γ m δ Q ˙ + ω m 2 δ Q = 2 ω m g o a s ( δ a + δ a ) + 2 ω m g e b s ( δ b + δ b ) ,
a + = κ o , ext E p κ o + i Δ o i δ i g o 2 n o f ( δ ) κ o , ext E p ( κ o + i Δ o i δ ) 2 ,
a = 1 f ( δ ) * ( κ o + i Δ o ) 2 i g o 2 κ o , ext E o 2 κ o , ext E p ( κ o + i δ ) 2 + Δ o 2 ,
f ( δ ) = 2 Δ o g o 2 n o ( κ o i δ ) 2 + Δ o 2 + 2 Δ e g e 2 n e ( κ e i δ ) 2 + Δ e 2 ω m 2 δ 2 i δ γ m ω m .
n o = κ o , ext E o 2 κ o 2 + [ Δ o 2 g o / ω m ( g o n o + g e n e ) ] 2 ,
n e = κ e , ext E e 2 κ e 2 + [ Δ e 2 g e / ω m ( g o n o + g e n e ) ] 2 .
a out ( t ) = ( E o κ o , ext a s ) e i Ω o t + ( E p κ o , ext a + ) e i ( δ + Ω o ) t κ o , ext a e i ( δ Ω o ) t = ( E o κ o , ext a s ) e i Ω o t + ( E p κ o , ext a + ) e i Ω p t κ o , ext a e i ( 2 Ω o Ω p ) t .
t ( Ω p ) = E p κ o , ext a + E p = 1 [ κ o , ext κ o + i Δ o i δ 1 f ( δ ) i g o 2 n o κ o , ext ( κ o + i Δ o i δ ) 2 ] .
FWM = | κ o , ext a E p | 2 = | 1 f ( δ ) * ( κ o + i Δ o ) 2 i g o 2 κ o , ext 2 E o 2 ( κ o + i δ ) 2 + i Δ o 2 | 2 .
Δ m = 2 m eff ω m Δ ω = 1 Δ ω ,
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.