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

Enhanced sensing of millicharged particles using nonlinear effects in an optomechanical system

Open Access Open Access

Abstract

Particles with electric charge 10-14 e in bulk mass are not excluded by present experiments. In the present letter we provide a feasible scheme to measure the millicharged particles via the optical cavity coupled to a levitated nanosphere. The results show that the optical probe spectrum of the nano-oscillator presents a tiny shift due to the existence of millicharged particles. Compare to the previous experiment the sensitivity can be improved by the using of a specific geometry to generate an electric field gradient and a pump-probe scheme to read the weak frequency shift. Owing to the very narrow linewidth(10-6 Hz) of the optical Kerr peak on the spectrum, this shift will be more obvious, which makes the millicharges more easy to be detectable. The technique proposed here paves the way for new applications for probing dark matter and nonzero charged neutrino in the condensed matter.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Millicharged particles(MCPs) or epsilon charged particles are the hypothetical particles with an electric charge e=εe, which is much smaller than the elementary charge e. The existence of the particles with small, unquantized electric charge can be introduced into the standard model in a variety of ways [1]. There has been interest in the possibility of a small, nonzero electric charge for the neutrino [2,3]. The possibility that dark matter carries a fractional or epsilon charge has been considered recently [4–6]. The search for the epsilon charges for the range from keV to GeV progressed through the efforts of many direct experiments and indirect observations over many years [7–11]. Now the search for stable particles with millicharges in bulk matter is still on the initial stage. Previous research using magnetic levitation electrometers or Millikan oil drop techniques in matter did not have sensitivity less than 0.1e [12–16]. Recently, Moore et al. report the search for particles with ε105using optically levitated microspheres [17]. Chen et al. propose to use a high-Q whispering-gallery-mode (WGM) optical microresonator to detect the surface and bulk charge of a dielectric nanoparticle [18]. Based the optical trapping scheme, we propose a levitated sensor to detect the epsilon charge in nano-scale with improved sensitivity of ε1014 in this paper by the cavity pump-probe technology.

Cavity optomechanics, which explores the interaction between optical field and mechanical motion, has witnessed rapid advances in recent years [19]. One of the principal advantages of cavity-optomechanical systems is the built-in readout of mechanical motion via the light field transmitted through (or reflected from) the cavity. Various of schemes have been proposed for the measuring of the parametric displacement [20], mechanical Fock-state detection [21], optical feedback cooling [22], extremely sensitive force sensing [23,24] and electrical charges measurement [18,25]. Recently, the optical pump-probe technique has become a popular topic, which offers an effective method to study the light-matter interaction. The optical pump-probe technology includes a strong pump field and a weak probe field. The strong pump laser is used to stimulate the system to generate coherent optical effect, while the weak laser plays the role of probe beam. Therefore, the linear and nonlinear optical effects can be observed on the probe spectrum in the two field's control scheme.

Most recently, this pump-probe scheme has also been realized experimentally in optical cavity [26–28]. In the present paper, we apply the pump-probe fields to the cavity-nanosphere coupling system and study theoretically the optical nonlinear in the optomechanical system. The results show that a Kerr peak with ultra narrow linewidth can be achieved through the high vacuum allowed. Based on the frequency shift induced by the millicharge, the detection of signature of MCPs in bulk matter can be carried out by the transmission spectrum. We expect that the optomechanical detector could improve the search for epsilon charges via an enhanced sensitivity of 108~109 than current experiments [17].

The remainder of the paper is organized as follows. Section two describes the theoretical model. Section three gives discussions about the measurement mechanism: 1) with the charged ring, the charge of the nanosphere is mapped to its mechanical resonant frequency. 2) with optical optomechanical system, very tiny drift in mechanical frequency could be readout optically. Section four studies the constraints and limits of the measurement system, discusses the thermomechanical noise and photon shot noise. We check the frequency stability and radiation damping rate respectively. The last section contains the summary of the main results of the present work.

2. Models

As shown in Fig. 1(a), a fused silica nanoshpere with radius a=50nm is placed inside a Fabry-Perot cavity and optically trapped at the antinodes of the optical lattice. The dielectric object is subjected to an electric field generated by a homogeneously charged ring with radius R0=50μm. Then we radiate a strong pump field(with frequencyωp) and a weak probe(with frequency ωpr) field on the cavity.

 figure: Fig. 1

Fig. 1 (a)Schematic diagram of the proposed setup for detecting millicharged particles by optical pump-probe technology in the cavity. Our approach involves optically trapping a fabricated nanosphere in the antinode of an optical standing wave. To eliminate the polarization force background, we use a homogeneously charged ring to produce symmetrical electric field near the nanosphere and the nanosphere should be trapped at the center of the ring. (b) Force analysis for the millicharged particles in the nanosphere.

Download Full Size | PDF

For small oscillation amplitudes, the levitated particle experiences a harmonic trapping potential with a trap stiffness(or the spring constant) k=ωn2m, where ωn is the trapping frequency and m is the mass of the nanosphere. The effective spring constant can be defined as k=k+F, here Frepresents the force gradient at the equilibrium position of the levitated nanosphere. If we consider the interaction between the electric field gradient and the single MCP, we get F=εeE, here E is the electric field gradient. The resonance frequency ωn can be converted into ωn=(k+F)/m. If the differentiation is sufficiently small, we get ωnωn, and F/2kωn/ωn1. Assumedω=ωnωn,based on the frequency shift of the resonance mode, one can obtain

F=2kωndω.
Here the frequency shift is dependent linearly on the force gradient via the linearized responsivity 2k/ωn.

To probe the MCPs, we electricize the ring homogeneously with an assumed linear density.υ The main systematic effect limiting the measurement sensitivity in [17] is the interacting primarily between the electric field and the dipole moments of the nanospheres. In order to minimize this effect, we suggest using the homogeneously charged ring. The nanosphere can be trapped at the centre of the ring thereby the applied electric field is distributed symmetrically along the horizontal axis. Thus the resultant polarization force in the nanosphere is zero. In cylindrical coordinates, the position of the center of the nanosphere is (0,0,0) as shown in Fig. 1(b). Now we consider a MCP appears in the nanosphere with the position (r,ϕ,z). The nanosphere will experience a force due to the electromagnetic interaction between the charged ring and the MCP. The electric field along zaxis at the MCP's location can be obtained by

Ez=q4π2Ε02zR02+r2+z2+2R0r(R02+r2+z22R0r)O,
where R0 is the radius of the ring, q=2πR0υis the total quantity of electric charge of the ring, Ε0 is the vacuum permittivity,O=(π/2)[1(1/2)2U2(13/24)2(U4/3)]is a total elliptical integral and the parameter U2=4R0r/(R02+r2+z2+2R0r). Since the scale of the nanosphere is much smaller than the ring, namelyr,zR0, the electric field can be obtained within approximation,
Ezq4π2Ε02zR03.
Then we can obtain the force gradient
Fz=2εeq4π2Ε0R03.
We replace Eq. (1) in the equation for Fz and considering k=ωn2m, it is easy to show that the single MPC will induce a shift as
dω=εeυ2πΕ0R02ωnm.
Interestingly, we find that the shift is independent with the position of the MPC in the nanosphere in this case. That is to say, no matter where it appears, the shift is merely decided by the total electron charge εe. In what follows, we propose a pump-probe scheme to readout the very tiny drift in mechanical frequency optically.

Then let’s denote the cavity mode annihilation (creation) operator by σ(σ+)with the commutation relation[σ,σ+]=1. The bosonic annihilation (creation) operators of the nanosphere oscillator are represented by sand s+with [s,s+]=1.Thus the Hamiltonian of the whole system in a rotating frame can be expressed as follows [29–31]

H=Δcσ+σ+ωns+sgσ+σ(s++s)iΩp(σσ+)iΩpr(σeiΔprtσ+eiΔprt),
whereg is the optomechanical coupling rate, Δc=ωcωp is pump-cavity detuning. We also introduce the Rabi frequency of the pump field and the probe field inside the cavity Ωp=2Ppκ/ωp and Ωpr=2Pprκ/ωpr, where Pp is the pump power, Ppr is the power of the probe field, respectively. We then define the operator n=s++s and the detuning of the probe and pump fieldΔpr=ωprωp. The temporal evolutions ofσandn can be obtained by solving the quantum Langevin equations with the damping terms
dσdt=(iΔc+κ)σ+ignσ+Ωp+ΩpreiΔprt+κa^in,
d2ndt2+γndndt+ωn2n=2ωngσ+σ+ξ^(t),
whereκ is the total energy damping rate of the cavity,γnis the damping rates of the vibrational mode of the nanosphere. In Eq. (7), a^in is the δ-correlated Langevin noise operator of the cavity-field quantum vacuum fluctuation, which has zero mean, a^in(t)a^in+(t)=δ(tt),a^in+(t)a^in(t)=0. The motion of nanosphere are affected by thermal bath of Brownian and non-Markovian stochastic process,ξ^(t) is the Brownian noise force with zero mean value [32,33]. Following standard methods from quantum optics, we derive the steady-state solutions to Eqs. (7) and (8) by setting all the time derivatives to zero,
σ0=Ωp(iΔc+κ)ign0,
n0=2g|σ0|2ωn.
We define the parameter ω0=|σ0|2,hence we can obtain
Ωp2=ω0[κ2+(Δc2g2ω0ωn)2].
One can always rewrite each Heisenberg operator as the sum of its steady-state mean value and a small fluctuation with zero mean value as follows
σ=σ0+δσ,
n=n0+δn.
Inserting these equations into the Langevin equations, one can safely neglect the nonlinear terms such asδσδn. Since the driving fields are weak, we will identify all operators with their expectation values, and drop the quantum and thermal noise terms. Then the linearized Langevin equations can be written as
δσ˙=κδσ+ig(n0δσ+σ0δn)+Ωp+ΩpreiΔprt,
δn¨+γnδn˙+ωn2δn=2ωngσ02.
To solve these equations, we make the ansatz as follows
δσ=σ+eiΔprt+σeiΔprt,
δn=n+eiΔprt+neiΔprt.
Upon substituting these equations into Eqs. (14) and (15), and upon working to the lowest order in Ωpbut to all orders inΩpr, we obtain in the steady state
σ=G(P+κ2)ΩprΩp2B(Y2P2)+2PGω0.
Hereσ is a parameter in analogy with the Kerr nonlinear optical susceptibility withY=iΔc+κ, P=iΔcign0,B=Δpr2iγnΔpr+ωn2, G=2ig2ωn.Using the input-output relationσout(t)=σin(t)2κσ(t),the amplitude of the Kerr nonlinear can be expressed asσout=2κσ. The transmission intensity of the output field is then given by [26]
|Tout|2=2κ|σoutΩpr|2.
The full width at half maximum (FWHM) of the transmission peak can be defined as FWHM=2|δpr|, with|Tout(δpr)|2=(1/2)[|Tout|2]max, where the detuningδpr=Δprωn.

3. Measurement

For numerical work, we use a fused silica nanosphere with the density ofρ=2.2g/cm3and the dielectric constant Ε=2. The trapping frequencyωndepends on the trapping beam intensity, it can be modulated by the intracavity power [29]

ωn=12π(6K2I0ρcReΕ1Ε+2)1/2,
herecis the speed of light, K=2π/λis the wave numbers of the trapping beams with wavelengthλ=1μm.In what follows, we choose the intensityI0=102W/μm2, hence one can obtainωn=150kHz. The quality factor one would get for a levitated nanosphere is solely determined by the air molecule impacts. Random collisions with residual air molecules provide the dampingγn=ωn/Q and thus, the quality factor due to the gas dissipation can be defined as [29,34]
Q=mωnvpA=aρωnv3p.
Herev=kBT/mgas is the thermal velocity of the gas molecules, pthe gas pressure andA the surface area of the resonators. For the nanosphere under the pressurep=1010Torr,we getQ=1.3×1011in the room temperature(T=300K). This value is considered to be within reach of moderate experimental improvements [29,35]. The optomechanical coupling between the cavity field and the nanosphere can be described by the coupling rate [29]
g=3Vs4VcΕ1Ε+2ωc,
whereVs=(4/3)πa3and Vc=(π/4)Lw2 are the nanosphere and the optical mode volumes, respectively. To estimate the coupling, we consider the use of optical microcavities operating at 1064nm, corresponding toωc=280THz, the length L=1mm [36], and mode waist w=50μm. Then the optomechanical coupling rate can be obtained asg=1.4×104Hz.

In order to remove the net charge of levitated sensor in trapping, a fiber-coupled xenon flash lamp is used to illuminate the electrode surfaces near the nanosphere [17]. To determine the original frequency, we hit the cavity by the pump-probe pulses and measure the absorption of the probe beam without the applied electric field. According to Eqs. (11) and (18), we depict the Kerr nonlinear susceptibilityσ as a function of probe-pump detuningΔprin Fig. 2(a) with Δc=0. Two sharp Kerr peaks are located at the original frequencies of the nanosphere resonator on the spectrum. Such a phenomenon is attributed to the destructive quantum interference effect between the nanosphere mode and the beat of the two optical fields via the cavity. If the beat frequency of two lasersΔpr is close to the resonance frequency of nanosphere, the mechanical mode starts to oscillate coherently, which results in Stokes-like (Δpr=ωn) and anti-Stokes-like(Δpr=ωn) scattering of light from the strong intracavity field.

 figure: Fig. 2

Fig. 2 (a)The plot of absorption spectrum as a function of probe-pump detuning. The unit of the Stokes scattering absorption intensity is an arbitrary unit(a.u.) which is a relative unit of measurement to show the ratio of amount intensity to a predetermined reference measurement. (b)Nonlinear optical spectrum of the probe field as a function of probe-pump detuning before (black solid line) and after (red dashed line and blue dash line) the action between millicharge and electric field gradient. The signals of the MCP withε=1014andε=5×1014can be well recognized in the spectrum. Other parameters used areΩp=3MHz,Δc=0, and Q=1.3×1011

Download Full Size | PDF

Then we illustrate the responses of Kerr nonlinear to external field by electricize the ring homogeneously with linear densityυ=104C/m. We depict Fig. 2(b) to show the transmission spectrum of the probe laser|Tout|2 as a function of the probe detuningδpr. As a MPC ofε=1014 appears in the matter, we use the black curve to show the Kerr peak in the absence of external field. Then we use the red dash curve to show the shift ofdω=6.4×106Hzcan be obtained on the probe spectrum due to the interaction between the charged ring and MPC. The dash blue sharp peak shows that the enhanced peak will exhibit a larger frequency shift on the transmission spectrum as we consider the MPC ofε=5×1014in the matter. The distance of shift dω as a function ofεfollows a linear relationship, dω=6.4×108ε. This may provide a direct method to measure the millicharges in this coupled system. The ability to improve the resolution depends on the FWHM of the transmission peak which can be estimated asFWHMωn/Q. From Eq. (21) we can see that the Q-factor is mainly limited by the pressure of the residual gas p. We can get FWHM3×106Hz from the condition ofp=1010Torr.

In the scheme we use a nanosphere with the radiusa=50nm, the massm=1.2×1018kg,the trapping frequencyωn=150kHz, and a charged ring with the radiusR0=50μm, linear densityυ=104C/m. Then we can calculate from Eq. (5) that the millicharges can be measured down to a precision ofεmin4.7×1015 in this coupling system.

4. Constraints and limits

M. R. Vanner et al. propose a scheme to realize state preparation and state purification of a mechanical resonator using short optical pulses as discussed in [37]. During a pulsed interaction of time scaleκ1ωn1 the mechanical position is approximately constant, therefore this scheme can be used as an alternative to “cooling via damping” for mechanical state purification. In our shcmem, the pump pulses can also be used for realizing high-purity states of the mechanical resonator. According to [37], the purity can be improved by two pulses with separationtp=π/2ωn, where the initial uncertainty in the momentum becomes the uncertainty in position. Considering the coupling to a thermal bath and starting with an initial thermal state n¯, the effective occupation of the final state after two pulses is

n¯eff12(1+1χ4+πn¯Qχ21).
Here the position measurement strength χ is an important parameter in this work as it quantifies the scaling of the mechanical position information onto the light field. The resulting optimal measurement strength is given by χ=25gNp/κ. We consider the use of pulses of mean photon numberNp=108, the optical microcavity has an amplitude decay rate κ=0.5GHzand the optomechanical couplingg=1.4×104Hz. For Q=1.3×1011, one can obtainn¯eff=0.09,corresponding to the effective temperature of Teff107K,indicating that ideal oscillators can be essentially decoupled from their thermal environment. Levitation under good vacuum conditions can lead to extremely low mechanical damping rates, such an approach should also facilitate the emergence of quantum behavior even in room-temperature environments.

The ultimate sensitive limits for nanomechanical resonators operating in vacuum are imposed by a number of fundamental physical noise processes. In our scheme, the thermal noise of the mechanical motion is the dominant noise source. In order to get the fundamental limits, we need to consider the minimum measurable frequency shiftdωminthat can be resolved in a realistic noise system. For the case ofQ1 [38],

dωmin[kBTeffωnΔfEcQ]1/2.
Here, the measurement bandwidthΔf=1/2πτ, which is dependent upon the measurement averaging time τ.Ec=mωn2xc2represents the maximum drive energy.xcis the maximum rms level which is still consistent with producing a predominantly linear response. For a Gaussian field distribution, the nonlinear coefficients are given byζ=2/w2 [35],where w=50μmis the beam waist radius. For small displacements|xc||ζ|1/2=3.5×105m,the nonlinearity is negligible. In our considerations, xcis taken to be 3 orders of magnitude smaller, we choose xc108m.Considering the quality factorQ=1.3×1011,Teff0.1μK,we can achieve thatdωmin2.7×106Hz<FWHMfor the measurement averaging timeτ=1/2πΔf0.01s.The result shows that the thermomechanical noise can be controlled at a level comparable to the claimed sensitivity at a short measuring time.

The photon recoil noise should play a role for nano-scale sensors [39]. The scattered power is calculated asP=σsI0withI0the trapping beam intensity andσsthe scattering cross section. The scattering cross sectionσs=|α|2K4/6πΕ02.The particle polarizabilityα=4πΕ0a3(ns21)/(ns2+2) andnsis the index of refraction of the nanosphere. Then we can obtain radiation damping rateγradP/mc.2For the parameters used in our scheme, I0=102W/μm2,we findγrad109dωmin.The result shows that the minimal detectable shift is above the limit set by the photon recoil noise. Therefore, we would expect that the induced frequency shift can be resolvable in actual experiment by using the subwavelength silica nanospheres. The recoil heating rate can be given byΓrecoil=Pωc/5mc2ωn.Usingωc=280THz andωn=150kHz,we predicts a reheating rate of Γrecoil=0.24Hz. The time between two cooling pulses should be larger than the inverse of the reheating rate, which is defined as the time required to increment one quantumωnof energy in the quantum harmonic oscillator. Since tp=π/2ωnΓrecoil1,indicating that oscillators can be essentially decoupled from their thermal environment.

At last we conservatively calculate limits on the abundance per nucleon Nof charged particles withε in Fig. 3, and compare to previous limits from optically levitated nanospheres [17] and magnetic levitometer [14] experiments. The rectangle region A and B denotes the space directly excluded by the previous experiments above their single particle thresholds. The region C denotes the constraints for our scheme. As shown in Fig. 3, the method we have described provides the direct search for MCP in condensed matter with sensitivity ofε1014 to single particle. Over this full range, the upper limit on the abundance per nucleon is at mostN109.The lines extending from each region show limits on the abundance below the single particle threshold for comparison. The yellow dash line shows the constraints form a high-Q whispering-gallery-mode (WGM) optical nanoresonator by the use of Cu2O nanoparticle [18]. The method we have described promises to be the most sensitive approach for detecting millicharged particles.

 figure: Fig. 3

Fig. 3 Constraints on the abundance of millicharged particles per nucleon N versus the epsilon charge. The results from this work (red rectangle C) are compared to previous results from levitated nanospheres (blue rectangle B) [17], magnetic levitometer (green rectangle A) [14] experiments and WGM nanoresonator [18](dash yellow line). The lines extending from each region show the upper limits below the single particle threshold.

Download Full Size | PDF

5. Conclusion

In this work, we investigate the Kerr nonlinear spectrum in the optomechanical system where a nanosphere coupled to a cavity. We have proposed an optical method to detect the existence of MCPs in bulk matter via pump-probe technology. A direct scheme to determine the electric charge is also demonstrated. The result shows that a sharp peak can be obtained at the resonant frequency. Based on the frequency shift of the vibrational mode, the MCPs can be measured on the spectrum. The Kerr nonlinearity with an ultra narrow linewidth gives rise to an improved sensitivity of order1014e,which is much smaller than that in current experiments and many other theoretical expectations. In the resonant millicharges detection, the strain sensitivity of our setup is limited by the thermal motion of the sensor. We would expect that the claimed sensitivity can be still achieved in actual noise system in the ultrahigh vacuum.

Funding

National Natural Science Foundation of China (11274230, 11574206); the Basic Research Program of the Committee of Science and Technology of Shanghai (14JC1491700).

References and links

1. S. Davidson, S. Hannestad, and G. Raffelt, “Updated bounds on milli-charged particles,” J. High Energy Phys. 5, 003 (2000).

2. A. Y. Ignatiev and G. C. Joshi, “Neutrino electric charge and the possible anisotropy of the solar neutrino flux,” Phys. Rev. D Part. Fields 51(5), 2411–2420 (1995). [CrossRef]   [PubMed]  

3. R. Foot, G. C. Joshi, H. Lew, and R. R. Volkas, “Charge quantization in the standard model and some of its extensions,” Mod. Phys. Lett. A 5(32), 2721–2731 (1990). [CrossRef]  

4. J. M. Cline, Z. Liu, and W. Xue, “Millicharged atomic dark matter,” Phys. Rev. D Part. Fields Gravit. Cosmol. 85(10), 101302 (2012). [CrossRef]  

5. E. Gabrielli, L. Marzola, M. Raidal, and H. Veermae, “Dark matter and spin-1 milli-charged particles,” J. High Energy Phys. 8(8), 150 (2015). [CrossRef]  

6. S. Profumo, “GeV dark matter searches with the NEWS detector,” Phys. Rev. D Part. Fields Gravit. Cosmol. 93(5), 055036 (2016). [CrossRef]  

7. T. Mitsui, R. Fujimoto, Y. Ishisaki, Y. Ueda, Y. Yamazaki, S. Asai, and S. Orito, “Search for invisible decay of orthopositronium,” Phys. Rev. Lett. 70(15), 2265–2268 (1993). [CrossRef]   [PubMed]  

8. A. A. Prinz, R. Baggs, J. Ballam, S. Ecklund, C. Fertig, J. A. Jaros, K. Kase, A. Kulikov, W. G. J. Langeveld, R. Leonard, T. Marvin, T. Nakashima, W. R. Nelson, A. Odian, M. Pertsova, G. Putallaz, and A. Weinstein, “Search for Millicharged Particles at SLAC,” Phys. Rev. Lett. 81(6), 1175–1178 (1998). [CrossRef]  

9. S. N. Gninenko, N. V. Krasnikov, and A. Rubbia, “New limit on millicharged particles from reactor neutrino experiments and the PVLAS anomaly,” Phys. Rev. D Part. Fields Gravit. Cosmol. 75(7), 075014 (2007). [CrossRef]  

10. A. D. Dolgov, S. L. Dubovsky, G. I. Rubtsov, and I. I. Tkachev, “Constraints on millicharged particles from Planck data,” Phys. Rev. D Part. Fields Gravit. Cosmol. 88(11), 117701 (2013). [CrossRef]  

11. A. Haas, C. S. Hill, E. Izaguirre, and I. Yavin, “Looking for milli-charged particles with a new experiment at the LHC,” Phys. Lett. B 746, 117–120 (2015). [CrossRef]  

12. P. C. Kim, E. R. Lee, I. T. Lee, M. L. Perl, V. Halyo, and D. Loomba, “Search for Fractional-Charge Particles in Meteoritic Material,” Phys. Rev. Lett. 99(16), 161804 (2007). [CrossRef]   [PubMed]  

13. I. T. Lee, S. Fan, V. Halyo, E. R. Lee, P. C. Kim, M. L. Perl, H. Rogers, D. Loomba, K. S. Lackner, and G. Shaw, “Large bulk matter search for fractional charge particles,” Phys. Rev. D Part. Fields 66(1), 012002 (2002). [CrossRef]  

14. M. Marinelli and G. Morpurgo, “Searches of fractionally charged particles in matter with the magnetic levitation technique,” Phys. Rep. 85(4), 161–258 (1982). [CrossRef]  

15. P. F. Smith, “Searches for fractional electric charge in terrestrial materials,” Annu. Rev. Nucl. Part. Sci. 39(1), 73–111 (1989). [CrossRef]  

16. G. S. LaRue, J. D. Phillips, and W. M. Fairbank, “Observation of Fractional Charge of (1/3)e on Matter,” Phys. Rev. Lett. 46(15), 967–970 (1981). [CrossRef]  

17. D. C. Moore, A. D. Rider, and G. Gratta, “Search for Millicharged Particles Using Optically Levitated Microspheres,” Phys. Rev. Lett. 113(25), 251801 (2014). [CrossRef]   [PubMed]  

18. Y. L. Chen, W. L. Jin, Y. F. Xiao, and X. Zhang, “Measuring the charge of a single dielectric nanoparticle using a High-Q optical microresonator,” Phys. Rev. Appl. 6(4), 044021 (2016). [CrossRef]  

19. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86(4), 1391–1445 (2014). [CrossRef]  

20. A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, and R. J. Schoelkopf, “Introduction to quantum noise, measurement, and amplification,” Rev. Mod. Phys. 82(2), 1155–1208 (2010). [CrossRef]  

21. Z. Yin, T. Li, X. Zhang, and L. M. Duan, “Large quantum superpositions of a levitated nanodiamond through spin-optomechanical coupling,” Phys. Rev. A 88(3), 033614 (2013). [CrossRef]  

22. P. F. Barker and M. N. Shneider, “Cavity cooling of an optically trapped nanoparticle,” Phys. Rev. A 81(2), 023826 (2010). [CrossRef]  

23. A. A. Geraci, S. B. Papp, and J. Kitching, “Short-Range Force Detection Using Optically Cooled Levitated Microspheres,” Phys. Rev. Lett. 105(10), 101101 (2010). [CrossRef]   [PubMed]  

24. J. Liu and K. D. Zhu, “Nanogravity gradiometer based on a sharp optical nonlinearity in a levitated particle optomechanical system,” Phys. Rev. D Part. Fields Gravit. Cosmol. 95(4), 044014 (2017). [CrossRef]  

25. H. Xiong, L. G. Si, and Y. Wu, “Precision measurement of electrical charges in an optomechanical system beyond linearized dynamics,” Appl. Phys. Lett. 110(17), 171102 (2017). [CrossRef]  

26. S. Weis, R. Rivière, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, and T. J. Kippenberg, “Optomechanically induced transparency,” Science 330(6010), 1520–1523 (2010). [CrossRef]   [PubMed]  

27. 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 475(7356), 359–363 (2011). [CrossRef]   [PubMed]  

28. A. H. Safavi-Naeini, T. P. M. 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 472(7341), 69–73 (2011). [CrossRef]   [PubMed]  

29. D. E. Chang, C. A. Regal, S. B. Papp, D. J. Wilson, J. Ye, O. Painter, H. J. Kimble, and P. Zoller, “Cavity opto-mechanics using an optically levitated nanosphere,” Proc. Natl. Acad. Sci. U.S.A. 107(3), 1005–1010 (2010). [CrossRef]   [PubMed]  

30. A. C. Pflanzer, O. Romero-Isart, and J. I. Cirac, “Master-equation approach to optomechanics with arbitrary dielectrics,” Phys. Rev. A 86(1), 013802 (2012). [CrossRef]  

31. 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(3), 033804 (2008). [CrossRef]  

32. C. Gardiner and P. Zoller, Quantum Noise, 2nd ed. (Springer, Berlin, 2000), p. 425.

33. V. Giovannetti and D. Vitali, “Phase-noise measurement in a cavity with a movable mirror undergoing quantum Brownian motion,” Phys. Rev. A 63(2), 023812 (2001). [CrossRef]  

34. F. R. Blom, S. Bouwstra, M. Elwenspoek, and J. H. J. Fiuitman, “Dependence of the quality factor of micromachined silicon beam resonators on pressure and geometry,” J. Vac. Sci. Technol. B 10(1), 19–26 (1992). [CrossRef]  

35. J. Gieseler, L. Novotny, and R. Quidant, “Thermal nonlinearities in a nanomechanical oscillator,” Nat. Phys. 9(12), 806–810 (2013). [CrossRef]  

36. N. Kiesel, F. Blaser, U. Delić, D. Grass, R. Kaltenbaek, and M. Aspelmeyer, “Cavity cooling of an optically levitated submicron particle,” Proc. Natl. Acad. Sci. U.S.A. 110(35), 14180–14185 (2013). [CrossRef]   [PubMed]  

37. M. R. Vanner, I. Pikovski, G. D. Cole, M. S. Kim, C. Brukner, K. Hammerer, G. J. Milburn, and M. Aspelmeyer, “Pulsed quantum optomechanics,” Proc. Natl. Acad. Sci. U.S.A. 108(39), 16182–16187 (2011). [CrossRef]   [PubMed]  

38. K. L. Ekinci, Y. T. Yang, and M. L. Roukes, “Ultimate limits to inertial mass sensing based upon nanoelectromechanical systems,” J. Appl. Phys. 95(5), 2682–2689 (2004). [CrossRef]  

39. V. Jain, J. Gieseler, C. Moritz, C. Dellago, R. Quidant, and L. Novotny, “Direct measurement of photon recoil from a levitated nanoparticle,” Phys. Rev. Lett. 116(24), 243601 (2016). [CrossRef]   [PubMed]  

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 (3)

Fig. 1
Fig. 1 (a)Schematic diagram of the proposed setup for detecting millicharged particles by optical pump-probe technology in the cavity. Our approach involves optically trapping a fabricated nanosphere in the antinode of an optical standing wave. To eliminate the polarization force background, we use a homogeneously charged ring to produce symmetrical electric field near the nanosphere and the nanosphere should be trapped at the center of the ring. (b) Force analysis for the millicharged particles in the nanosphere.
Fig. 2
Fig. 2 (a)The plot of absorption spectrum as a function of probe-pump detuning. The unit of the Stokes scattering absorption intensity is an arbitrary unit(a.u.) which is a relative unit of measurement to show the ratio of amount intensity to a predetermined reference measurement. (b)Nonlinear optical spectrum of the probe field as a function of probe-pump detuning before (black solid line) and after (red dashed line and blue dash line) the action between millicharge and electric field gradient. The signals of the MCP with ε= 10 14 and ε=5× 10 14 can be well recognized in the spectrum. Other parameters used are Ω p =3MHz, Δ c =0, and Q=1.3× 10 11
Fig. 3
Fig. 3 Constraints on the abundance of millicharged particles per nucleon N versus the epsilon charge. The results from this work (red rectangle C) are compared to previous results from levitated nanospheres (blue rectangle B) [17], magnetic levitometer (green rectangle A) [14] experiments and WGM nanoresonator [18](dash yellow line). The lines extending from each region show the upper limits below the single particle threshold.

Equations (24)

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

F= 2k ω n dω.
E z = q 4 π 2 Ε 0 2z R 0 2 + r 2 + z 2 +2 R 0 r ( R 0 2 + r 2 + z 2 2 R 0 r) O,
E z q 4 π 2 Ε 0 2z R 0 3 .
F z = 2εeq 4 π 2 Ε 0 R 0 3 .
dω= εeυ 2π Ε 0 R 0 2 ω n m .
H= Δ c σ + σ+ ω n s + sg σ + σ( s + +s)i Ω p (σ σ + )i Ω pr (σ e i Δ pr t σ + e i Δ pr t ),
dσ dt =(i Δ c +κ)σ+ignσ+ Ω p + Ω pr e i Δ pr t + κ a ^ in ,
d 2 n d t 2 + γ n dn dt + ω n 2 n=2 ω n g σ + σ+ ξ ^ (t),
σ 0 = Ω p (i Δ c +κ)ig n 0 ,
n 0 = 2g | σ 0 | 2 ω n .
Ω p 2 = ω 0 [ κ 2 + ( Δ c 2 g 2 ω 0 ω n ) 2 ].
σ= σ 0 +δσ,
n= n 0 +δn.
δ σ ˙ =κ δσ +ig( n 0 δσ + σ 0 δn )+ Ω p + Ω pr e i Δ pr t ,
δ n ¨ + γ n δ n ˙ + ω n 2 δn =2 ω n g σ 0 2 .
δσ = σ + e i Δ pr t + σ e i Δ pr t ,
δn = n + e i Δ pr t + n e i Δ pr t .
σ = G (P+ κ 2 ) Ω pr Ω p 2 B( Y 2 P 2 )+2PG ω 0 .
| T ou t | 2 =2κ | σ ou t Ω pr | 2 .
ω n = 1 2π ( 6 K 2 I 0 ρc Re Ε1 Ε+2 ) 1/2 ,
Q= m ω n v pA = aρ ω n v 3p .
g= 3 V s 4 V c Ε1 Ε+2 ω c ,
n ¯ eff 1 2 ( 1+ 1 χ 4 + π n ¯ Q χ 2 1).
d ω min [ k B T eff ω n Δf E c Q ] 1/2 .
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.