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

Magnetometry via spin-mechanical coupling in levitated optomechanics

Open Access Open Access

Abstract

We analyze magnetometry using an optically levitated nanodiamond. We consider a configuration where a magnetic field gradient couples the mechanical oscillation of the diamond with its spin degree of freedom provided by a nitrogen vacancy center. First, we investigate the measurement of the position spectrum of the mechanical oscillator. We find that conditions of ultrahigh vacuum and feedback cooling allow a magnetic field gradient sensitivity of 1μTm1/Hz. At high pressure and room temperature, this sensitivity degrades and can attain a value of the order of 100mTm1/Hz. Subsequently, we characterize the magnetic field gradient sensitivity obtainable by maneuvering the spin degrees of freedom using Ramsey interferometry. We find that this technique can offer photon-shot noise and spin-projection noise limited magnetic field gradient sensitivity of 100μTm1/Hz. We conclude that this hybrid levitated nanomechanical magnetometer provides a favorable and versatile platform for sensing applications.

© 2017 Optical Society of America

1. Introduction

Recent developments in cavity opto- and electro-mechanics have actively explored the cooling and manipulation of mechanical motion of nano resonators [1–5]. The spectacular progress in this research has been motivated by ideas for applications in ultrasensitive metrology [6] and for testing macroscopic quantum mechanics [7]. However, experimental investigations in this field suffer from heating and decoherence produced by mechanical clamping of the oscillator. An effective remedy to this problem is to isolate the mechanical oscillator from its environment by means of levitation using optical [8–10] or magnetic fields [11, 12], for example. In particular, optically levitated nanoparticles have been investigated for cooling [13–15], preparation of quantum superposition states [16,17] and precision sensing applications [18,19].

Interestingly, optical levitation of solid state quantum emitters possessing internal degrees of freedom such as Nitrogen-vacancy (NV)-centers provides a promising platform for optomechanical studies [20–22]. Such hybrid nano-mechanical systems are of significant interest for generating macroscopic quantum superposition states [23–26] and testing quantum gravity in vacuum [27]. The key ingredient in these proposals is the coupling between the nanomechanical oscillator and spin degrees of freedom mediated by either a magnetic field gradient (MFG) [28,29] or a lattice strain field [30, 31]. Such couplings have been extensively explored for mechanical cooling [32, 33], mechanical squeezing [34] and spin squeezing [35, 36]. Due to their easy controllability, fast manipulation, as well as unprecedented coherence time under ambient conditions, NV centers are attractive quantum systems for applications in quantum information processing and ultrasensitive magnetometry [37].

More recently, optomechanical system based on whispering gallery modes have been used to realize magnetometers [38, 39] with magnetic field sensitivity 131 pT Hz−1/2 [40]. These magnetometers are miniaturized and offer large dynamic range, unlike superconducting quantum interference devices (SQUID) [41] and spin exchange relaxation free (SERF) magnetometers [42] which however have better sensitivity. All these magnetometers demonstrate sensitivity to a magnetic field. However, the investigation of MFG sensitivity is equally important due to its potential applications in the generation of macroscopic superposition states [23–26], testing of quantum gravity [27], magnetic resonance imaging [43], and cell therapy [44]. In recent years, non-levitated cantilever type of systems have been used for the measurement of MFG (~105 Tm−1 in [29, 45]). However, the analysis for the measurement of MFG sensitivity in levitated systems still needs to be explored. In this paper, we propose the exploration of MFG sensitivity based on an NV-center levitated in an optical dipole trap. We consider two ways in which the measurement of a MFG sensitivity can be carried out - (i) by manipulating the mechanical degrees of freedom, and taking a trace over the spin degrees of freedom, and (ii) by manipulating the spin degrees of freedom and tracing over the mechanical degrees of freedom. As recently predicted, maneuvering of the mechanical degrees of freedom under ultrahigh vacuum is expected to lead to the preparation of the ground state [14]. Working under these conditions, we obtain a MFG sensitivity of the order of 1μTm1/Hz. However, the MFG sensitivity of the proposed magnetometer degrades under less stringent conditions. Specifically, we identify a suitable parameter regime where MFG sensitivity can attain a value of 100mTm1/Hz at room temperature and atmospheric pressure.

On the other hand, in order to manipulate the spin of the NV-center we consider a Ramsey pulse sequence [37]. After initial preparation of the spin state, the spin-mechanical interaction is switched on during the free evolution time, which leads the mechanical states to acquire different phases conditioned on the spin states [24, 25]. Finally, after a suitably chosen interaction time, the acquired phase can be read out optically. Following this procedure, we obtain the photon-shot noise and spin-projection noise limited MFG sensitivity of the order of 100μTm1/Hz. For these conditions, we find that the proposed magnetometer does not suffer appreciably from decoherence under realistic experimental conditions, unlike clamped oscillators [38–40]. Further, cavity optomechanical magnetometers utilize a magnetostrictive material whose geometry is difficult to control causing an undesirably weak coupling between mechanical motion and applied force [38]. Contrary to this, the proposed magnetometer employs a nanodiamond with a single NV-center which are attractive systems for magnetometry [46]. Thus, the ability to sensitively detect magnetic field gradients, portability, small size and suitability for working under both feedback-cooled as well as room temperature conditions, makes the proposed magnetometer appropriate for sensitive applications [47].

2. Model

Recently, nanodiamonds containing single nitrogen-vacancy (NV) center have been optically levitated in vacuum [20]. We consider such a single NV-center diamond of mass m which is trapped in vacuum by a focused Gaussian beam resulting in mechanical oscillations at a frequency ωm, as shown in Fig. 1. For small amplitudes, as considered in this paper, the oscillations along the three spatial directions are uncoupled and may be considered independently. We will consider nanodiamond oscillations in the optical trap along the z-axis. The position of the levitated particle is continuously monitored by means of interferometric techniques. The detected position information is fed back to the trapping beam in order to modulate its optical intensity [20]. Such feedback causes additional damping of the nanodiamond and gives rise to cooling, in addition to also causing some backaction heating. Under optimal conditions, the ground state of the oscillator can be prepared [14]. Neglecting the spin degree of freedom for the moment, the quantum dynamics of the nanodiamond oscillation are modeled by the master equation [14]

ρ˙=1i[Hm,ρ](At+Ap+Dp2)D[Qz]ρ(t)Dq2D[Pz]ρ(t)iηf4m[Qz,{Pz,ρ}]+[ρ(t)],
where, the first term represents the unitary evolution of the motion of center-of-mass mode am described by the Hamiltonian Hm=ωmamam. The second (third) term describes the momentum (position) diffusion of the nanodiamond due to collisions with background gas, with momentum (position) diffusion coefficient Dp=2ηfkBTz02/2(Dq=ηf2/24kBTm2z02), where, T is the gas temperature, kB is Boltzmann’s constant and z0=/2mωm is the zero point fluctuation of the mechanical oscillator. Also, At (Ap) represents the heating rate due to trap (probe) beam photon scattering, respectively. The fourth term accounts for gas damping with coefficient of friction ηf = 6πμR, where μ is the dynamic viscosity of the surrounding gas and R is the radius of nanodiamond. The superoperator in the fifth term corresponds to feedback and contains both the respective feedback damping and backaction and is given by [ρ(t)]=iχ2ΦG[Qz3,{Pz,ρ}]χ2ΦG2D[Qz3]ρ. Here, Φ is the average detected flux of photons, G is the feedback gain and χ is the scaled optomechanical coupling. The respective dimensionless position and momentum operators are defined as Qz=am+am and Pz=i(amam). The Lindblad superoperator in Eq. (1) is written as D[Qz]ρ=QzQzρ+ρQzQz2QzρQz and the creation (annihilation) operator of the mechanical oscillator along z-axis is am(am). In earlier works, this master equation was shown to analytically model the phonon dynamics [14] and oscillator phase space distributions [48] observed experimentally.

 figure: Fig. 1

Fig. 1 Schematic of the optically levitated nanodiamond oscillating in dipole trap along z-direction. A green laser of wavelength 532 nm excites the nanodiamond. A MFG along z-direction is applied to engineer spin-mechanical coupling. Microwaves are used to manipulate the ground spin state of nanodiamond.

Download Full Size | PDF

Besides the mechanical motion, levitated optomechanics of the nanodiamond provides a versatile platform for coherent control over spin states, as well [22]. We now consider the addition of the spin degree of freedom to the dynamics. The NV defect is typically available in two different forms, namely the neutral state NV0 and negatively charged state NV [37]. These two forms have different spin and optical properties. However, the negatively charged state of the defect is of particular interest for magnetometry applications, due to the existence of its spin triplet ground state which can be initialized, coherently manipulated and read out by optical means. We consider illumination by green laser of 532 nm wavelength driving the spin-conserving transitions 3 A23 E2 followed by spin-selective intersystem crossing towards an intermediate spin singlet state 1A1 [see Fig. 2(a)]. This leads to a high degree of spin polarization in the ms = 0 state (here, ms denotes the spin projection along the intrinsic quantization axis), which results in higher photoluminescence intensity from the 637 nm zero-phonon line, as shown in Fig. 2(a).

 figure: Fig. 2

Fig. 2 (a) Level diagram for excitation and read-out of NV-spins by optical means. Inset shows the bare energy-level diagram for the triplet ground-state of NV-center in nanodiamond. The microwave fields of Rabi frequency Ω0 drives the transitions |0〉 → |+ 1〉 and |0〉 → | − 1〉 with detuning Δ. (b) The dressed-state description of NV-spin. Phonons of energy ωm drives the transitions |c〉 → |b〉 and |b〉 → |a〉 with effective detunings Δ1 = ωmωbc and Δ2 = ωmωab, respectively.

Download Full Size | PDF

The spin triplet electronic ground state 3A2 of the NV-center with ms = 0 and ms = ±1 is shown in the inset of Fig. 2(a). The bare state Hamiltonian describing the NV spin is written as HNV = − ℏΔ (| + 1〉〈+1| + |−1〉〈−1|), where Δ is the detuning. Now, once initially prepared in the ms = 0 state by optical means, the NV center is subjected to two microwave fields which drive the transitions |0〉 ↔ | + 1〉 and |0〉 ↔ | − 1〉 with both Rabi frequencies equal to Ω0. The corresponding Hamiltonian in the rotating-wave frame of the microwave field is written as Hdrive=Ω02(|0+1|+|01|+h.c.). We now apply a MFG B0=Bz(z)z along the z-direction. This creates a magnetic field B=B0z and generates the coupling between spin and mechanical degrees of freedom of the oscillator, governed by the interaction Hamiltonian Hint=gSz(am+am), where, Sz=|+1|+1|11| is the spin operator aligned along NV symmetry axis, g=glμBB0z0 is coupling strength, gl ≈2 is the Landé factor and μB is the Bohr magneton.

It is to be noted that besides the center of mass motion, the torsional vibrations of levitated nanodiamonds have also been realized [49, 50]. Further, the quantum ground state cooling of torsional mode [49, 51] and torsional matter-wave interferometry [52] is proposed as well. In these works the torsional modes can be coupled to NV center by employing a homogeneous magnetic field and this torsional-spin coupling is proportional to the coordinate of torsional displacement. However, in the present system we work in a regime where homogeneous magnetic field vanishes and only MFG is present, which implies to the lowest order spin-torsional coupling vanishes. Nevertheless, MFG creates a magnetic field which is proportional to the linear vibrational mechanical amplitude. Hence the torsional-spin coupling is proportional to the product of linear times torsional mechanical amplitude which comes out to be very small in the context of present system. Typically, the particle is energetic enough to climb out of the torsional well and rotate freely, the rotational frequency is about an order of magnitude higher than the oscillation frequency and the spin-torsional interaction averages to zero [53].

Now, to accurately describe the spin-mechanical interaction, we consider the resonant interaction of microwave fields with the NV center in the dressed state eigenbasis of HNV + Hdrive [see Fig. 2(b)]. These dressed states with their corresponding eigenvalues can be written as

|a=sin(θ)|0+cos(θ)|+;ωa=(Δ+Δ2+2Ω02)/2,|b=|;ωb=Δ,|c=cos(θ)|0sin(θ)|+;ωc=(ΔΔ2+2Ω02)/2,
where, |±=(|+1±|1)2 and tan(2θ)=2Ω0Δ.

3. Manipulation of mechanical motion

In this section, to analyze the MFG sensitivity, we consider the manipulation of the mechanical degrees of freedom and taking trace over the spin degree of freedom. The physical idea is to investigate the effect of weak coupling of the mechanics to the spin degree of freedom which is the one addressed by the MFG. Specifically, we will analyze the position spectrum of the mechanical oscillator. As explained above, optical illumination leads to electric dipole transitions between the ground and excited states, subsequently followed by fast dissipation. This dissipation controls the populations and coherences for the NV dressed states. It is to be kept in mind that such a quick dissipation mechanism is necessary for the generation of steady state squeezing and dissipation-induced cooling leading to the preparation of ground state of the mechanical oscillator. Taking this into account, we trace over the spin degrees of freedom and include the contribution from various dissipative terms to obtain the following full master equation for the mechanical oscillator [14,34]

ρ˙=1i[Hm,ρ](At+Ap+Dp2)D[Qz]ρ(t)Dq2D[Pz]ρ(t)iηf4m[Qz,{Pz,ρ}]+[ρ(t)]A2D[am]ρA+2D[am]ρiδ2[amam,ρ],
where, A=2g2α1(A+=2g2α2) and δ = 2g2α3 describe dissipation induced cooling (heating) and mechanical frequency shift due to spin mechanical coupling. Here, α1, α2 and α3 can be written as [34]
α1=Re[sin2(θ)N(p2ρcc+p3ρca)+cos2(θ)Np1ρbb],α2=Re[cos2(θ)N(p1ρaa+p3ρac)+sin2(θ)Np2ρbb],α3=Im[sin2(θ)(p2(ρccρbb)+p3ρca)N+cos2(θ)(p1(ρaaρbb)+p3ρac)N],
where, p1=iΔ1+Γ12(1+sin2θ), p2=iΔ2+Γ12(1+cos2θ), p3=Γ14sin(2θ), N=p1p2p32 and Γ1 ≈ Γ2 is the decay of states |+ 1〉 and |− 1〉, respectively. The detailed derivation of population and coherence terms in Eq. (4) is presented in [34].

3.1. Position spectrum

Now to continuously monitor the position of nanodiamond, the master equation Eq. (3) can be unraveled in terms of the following set of quantum Langevin equations for position and momentum quadratures [14]

Q˙z=(ωm+δ2)Pz(AA+2)Qz,
P˙z=(ωm+δ2)Qz[Γ+(AA+2)]Qz+2mΓ0kBTeffξT(t)+18ηΦnG2Q4ξF(t)+[2A++N(2A+2A)]ξS(t),
where, ξ (t) are white noise terms with zero mean and correlation E [ξ (t) ξ (t′)] = δ (tt′), and Teff is the effective temperature of the total background due to the combination of gas and optical scattering. The parameter Γ = Γ0 + δΓ, where Γ0 represents the gas damping and δΓ12χ2G(N+12). Note that last stochastic term in Eq. (6) arises as a consequence of the spin-mechanical coupling and depends on the mechanical state of the system. Using Eqs. (5,6), the stochastic differential equation for position (qz = z0Qz) can be written as
q¨z+[Γ+AA+]q˙z+{(ωm+δ2)2+[Γ+(AA+2)](AA+2)}qz=(FT+FF+FS)m,
where, FT, FF and Fs are independent stochastic forces due to thermal, feedback backaction heating and spin interaction, respectively with zero mean and correlations FT(t)FT(t)=STδ(tt), FF(t)FF(t)=SFδ(tt) and FS(t)FS(t)=SSδ(tt). Here,
ST=(1+δ2ωm)2mΓ0kBTeff,
SF=(1+δ2ωm)54mωmχ2ΦG2(2N2+2N+1),
SS=(1+δ2ωm)mωm[A++N(A+A)].
The position spectrum of the mechanical oscillator can be obtained by taking the Fourier Transform of Eq. (7) which gives
q˜z(ω)=χm(ω)[F˜T+F˜F+F˜S],
where, χm (ω) represents the optomechanical susceptibility of the oscillator and is given by
χm(ω)=1m{[(ωm+δ2)2+(AA+2)(Γ+AA+2)ω2]i(Γ+AA+)ω}.
The expectation value of q˜z is therefore q˜z(ω)=0 and the positional power spectral density (PSD) is written as
|δq˜z(ω)|2=|χm(ω)|2[|F˜T|2+|F˜F|2+|F˜S|2].
Thus, the position noise spectrum including the contribution from shot-noise is given by
|δq˜z(ω)|2=|χm(ω)|2[ST+SF+SS]+z02χ2Φ.
The position noise spectrum in Eq. (14) depends upon the spin-mechanical coupling and hence on the MFG. We explore this dependence of the noise spectrum to describe the sensitivity of MFG. To do so, we work in a regime of low pressure where cooling from various dissipative mechanisms dominates over heating. For instance, the optimal feedback cooling at pressure 10−5 mbar assists to prepare the ground state of the mechanical oscillator with 〈N〉 < 1 [14]. For the spin part, we consider the cooling transition (|c〉 → |b〉) to be resonant (Δ1 = 0) while the heating transition (|a〉 → |c〉 is assumed to be far off-resonant (Δ2 = −Δ). In this situation, A+=0, δ = 0 and A simplifies to first order as [34] Ag2α, where
α=4Γ1(1+cos(2θ))sin2(2θ)9cos2(2θ).
The cooling term as described above varies with microwave power. It is possible to cool the mechanical oscillator with 〈N〉 < 1 by working in a suitable regime of microwave power. For instance, at Ω0 = 0.8ωm dissipative cooling dominates over heating and the steady state mean phonon number approximates to 0.3 [34].

Working under the conditions stated above, we plot the position noise spectrum from Eq. (14) in Fig. 3. In the absence of spin-mechanical coupling (SS=0), the noise spectrum is dominated by the feedback cooling [14] term proportional to SF in Eq. (14). However, the amplitude of the spectrum decreases at resonance as the spin-mechanical coupling is introduced [see Fig. 3(a)]. Our analysis indicates that the noise spectrum in Fig. 3 is dominated by | χm (ω) |2 [see Eq. (14)], and the reduction in amplitude of position PSD at resonance is a consequence of the increase in the damping of the mechanical oscillator caused by spin-mechanical interaction [54]. The susceptibility at the slightly shifted mechanical resonance ω=ωm2+A2(Γ+A2)12(Γ+A)2ωm attains the following form

|χm(ωm)|max2=[m2{(Γ+A)2[(ωm2+A2(Γ+A2))12(Γ+A)2]}]1.

 figure: Fig. 3

Fig. 3 (a) Position power spectral density for z-component of motion versus frequency. The parameters chosen are ωm = 38kHz, R =50 nm, density=2200 kg/m3, Teff = 4 K, χ = 10−7, G = 1/18, Ω0 = 0.8ωm, Γ1 = 0.25ωm. (b) Position power spectral density versus scaled spin-mechanical coupling at ω = ωm. The inset shows the position PSD corresponding to red ellipse for smaller values of g/ωm.

Download Full Size | PDF

It follows from Eq. (16) that |χm(ωm)|max2 decreases with increase in A and hence with increase in the spin-mechanical coupling (g), [see Eq. (15)]. In Fig. 3(b), we depict the decrease in the amplitude of noise spectrum (at resonance) with spin-mechanical coupling. It follows that there is a maximum value of spin-mechanical coupling (gmax=0.62ωm) at which the reduction in position PSD at resonance is maximum and beyond which noise is larger than the peak value of the position measurement at resonance. This value is indicated by an arrow in Fig. 3(b). We also note that the smaller values of the spin-mechanical coupling would not be useful as for these values it would be difficult to detect the change in the peak of position PSD with respect to the case when coupling is absent. The minimum coupling for which such a change can be detected corresponds to gmin = 2 × 10−3ωm, as shown in the inset of Fig. 3(b).

3.2. Magnetic field gradient sensitivity

In the preceding analysis, we showed that the spin-mechanical coupling created by MFG substantially affects the position noise spectrum. This position PSD can be treated as a signal for which MFG sensitivity can be written as [37,55]

ηB=|δq˜z(ω)|2|δq˜z(ω)|2B0×tm,
where, tm is the measurement time. Under cooling conditions, MFG sensitivity can be written as,
ηB=|δq˜z(ω)|2μBglz0|χm(ω)|2[D+mωmNα],
where, D is given by
D=2m2gα(Γ+A)|χm(ω)|2(ST+SF+SS)[ωm2+ω2+A2(Γ+A2)].

The sensitivity spectrum of MFG is shown in Fig. 4. The peak value of MFG sensitivity decreases with increase in spin-mechanical coupling and can attain a value of 1μTm1/Hz at g = 0.11ωm, as shown in Fig. 4(a). Interestingly, there exists an optimum value of spin-mechanical coupling (gopt=0.11ωm) such that for g < gopt MFG sensitivity improves with increase in g and for g > gopt its value start increasing, as depicted in Fig. 4(b). For g < gopt, MFG sensitivity is dominated by high spin-mechanical noise and mechanical susceptibility, both of which diminish with increase in g. This results in the improvement of sensitivity reaching a minimum value at g = gopt. However, for g > gopt, spin-mechanical noise becomes flat and susceptibility decreases causing the MFG sensitivity to increase with increase in spin-mechanical coupling. But still in this regime it can acquire a value of the order of 2μTm1/Hz at g = gmax.

 figure: Fig. 4

Fig. 4 MFG sensitivity versus (a) frequency (b) scaled spin-mechanical coupling for tm=2πωm. Other parameters are same as in Fig. 3.

Download Full Size | PDF

So far, the above analysis is based on the preparation of ground state for a rather low value of pressure (10−5 mbar). However, to illustrate the performance of the proposed magnetometer for higher pressure and temperature, we first write the equation for the phonon number dynamics [14] using Eq. (3)

N˙=2JN2(J+K)N+,
where, J=12[G9G2]χ2Φ, K=ηf/m+J+A+A+ and =Dp+γopt+A+ and γopt governs the optical scattering rate. The steady state phonon number from Eq. (20) can be written as
NssDp+γopt+A+2J.
This approximation holds for N0 = kBTeff/ℏωm ≫ 1, where N0 is the initial number of phonons. Note that MFG sensitivity is influenced by the state of the system. On this note, we exhibit in Fig. 5(a) that steady state phonon number increases with pressure for a particular value of g. However, the increase in spin-mechanical coupling leads to decrease in steady state phonon number for a specific value of pressure [see Fig. 5(a)].

 figure: Fig. 5

Fig. 5 (a) Steady state phonon number versus g/ωm and Pressure (mbar). MFG sensitivity versus (b)g/ωm and Pressure (mbar), (c) g/ωm and Temperature (K). In plots (a) and (b) Teff =300 K, γopt =1 kHz, while in (c) pressure is 0.3 mbar. The other parameters are same as in Fig. 3.

Download Full Size | PDF

Now, the MFG sensitivity from Eq. (17) for high pressure and temperature can be written as,

ηB=μBglz0|δq˜z(ω)|2|χm(ω)|2(D1+D2+D3)+D4(ST+SF+SS),
where, Di (i = 1, 2, 3, 4) are given by
D1=2gα3ωm2mΓ0KBTeff,
D2=108mgωmχ2ΦG2[α3ωm(2N2+2N+1)+(1+δ2ωm)α2J(2+1N)],
D3=mωm2g[α3ωm{A++N(A+A)}+(1+δ2ωm){α2A++N(α2A+α1A)+12Nα2J(A+A)}],
D4=8m2g|χm(ω)|4{[(ωm+δ2)2+(AA+2)(Γ+AA+2)ω2]×[α3(ωm+δ2)+(α1+α22)(Γ+AA+)]+(α1α2)(Γ+AA+)ω2}.

Using Eq. (22), the influence of pressure and temperature on MFG sensitivity is shown in Fig. 5. For small spin-mechanical coupling the increase in pressure degrades the MFG sensitivity, as depicted in Fig. 5(b). However, for a particular value of pressure, MFG sensitivity still improves with spin-mechanical coupling for the same reasons as explained after Eq. (19). For instance at a particular pressure, for a regime of g ≥ 0.4ωm, the MFG sensitivity improves and can even attain a value of the order of 100mTm1/Hz, as shown in Fig. 5(b). Further, the MFG sensitivity is influenced by temperature as exhibited in Fig. 5(c). Again the rise in temperature degrades the MFG sensitivity. However, for a particular temperature, a strong spin-mechanical coupling improves the sensitivity and at room temperature, the improvement is of the order of 500mTm1/Hz for increase in spin-mechanical coupling [see Fig. 5(c)]. Thus, the present hybrid nanomechanical sensor harnesses useful MFG sensitivity even under room temperature and pressure conditions (albeit in the range of mTm1/Hz.

4. Manipulation of spin

In the preceding analysis, the MFG sensitivity was described by manipulating the mechanical degrees of freedom and tracing over the spin degrees of freedom. In this section, we show that manipulation of spin degree of freedom can also provide the MFG sensitivity. The essential idea is to prepare the spin-mechanical system in a separable state and allow it to undergo free evolution in the presence of spin-mechanical coupling. The result of this is that mechanical coherent states acquire different phase depending on the spin state [24, 25]. For a particular evolution of time, the acquired phase can be completely written on the spin states, which can be read by using Ramsey interferometry [37].

Here, we consider a single spin NV center, as shown in Fig. 1, driven by microwave fields and coupled to a mechanical degree of freedom by means of MFG. In diamond substrates, ms = ±1 levels are typically degenerate. However, this degeneracy can be broken either by a strain field or by applying a static magnetic field. Thus, it is sufficient to use two-levels for the spin manipulation. From now onward, we consider only the |0〉 → |+ 1〉 transition of the spin. The Ramsey setup used to manipulate the spin of NV center, is shown schematically in Fig. 6. In the first step, the hybrid system is prepared in a separable state |ψ (0)〉 = |0〉|α〉. Here, the mechanical system is assumed to be in coherent state |α〉. Now, a π/ 2 microwave pulse is applied such that spin state is rotated into the superposition state

|ψ1=Uπ/2|ψ(0)=(|0i|+12)|α.
Now, the spin-mechanical interaction is introduced by turning on the MFG and letting the system to evolve freely so as to produce the following state
|ψ2(t)=Ut|ψ1=(|0|α(t,0)i|+1|α(t,+1)2),
where, Ut=exp(iHintt). Note that the spin-dependent mechanical coherent states in Eq. (28) acquire different phases which are transferred to the spin states after an oscillation period T=2πωm and the resultant state can be written as [24]
|ψ2(T)=(|0iexp(4ig2Tωm)|+12)|α.
The phase difference written on the spin states can be revealed by applying second π/2 microwave pulse such that the final state becomes
|ψ3=12[(1exp(4ig2Tωm))|0i(1+exp(4ig2Tωm))|1]|α.
Thus, the population of ground state |0 can be written as
P0=12(1cos(4g2Tωm)).
The population in Eq. (31) quantifies the following fluorescence signal [56, 57] which can be monitored to yield the MFG sensitivity
S=β0+β12β0β12cos(4g2Tωm),
where, β0 is the number of photons collected in one measurement in the absence of spin-mechanical interaction i.e. the number of photons detected from each spin in |0〉 state and β1 is the number of photons collected in one measurement when there is a phase accumulation of π during free precision i.e. number of photons detected from each spin in | + 1〉 state. Notably, the performance of the magnetometer is ultimately limited by quantum fluctuations associated with photon shot noise and spin projection noise. Thus, the total noise in the MFG is given by the quadrature sum of these sources of noise [58]
δB0=δB0psn2+δB0spn2,
where δB0psn and δB0spn represent the contributions from photon-shot noise and spin-projection noise, respectively [59]. In NV-magnetometers, a very small number of photons are collected from each NV spin, which give rise to photon-shot noise. Thus, the photon-shot noise limited uncertainty can be approximated as δB0psnβ, where β=β0+β12 is the average number of photons collected per measurement. On the other hand, the spin-projection noise arises due to random projection of spin states into one of the states compatible with the measurement process [59] and is given by [60], δB0spn=σz=ψ3|σz2|ψ3ψ3|σz|ψ32. Using, Eqs. (30,33), the minimum uncertainty in the MFG measurement due to the combination of photonshot noise and spin-projection noise can be written as
(δB0)min=β+1cos4(4g2Tωm)max|SB0|.
Further, using Eq. (32), MFG noise becomes
(δB0)min=ωmβ+1cos4(4g2Tωm)8gμBglz0TCβ,
where, C is the measurement contrast. Thus, the MFG sensitivity limited by photon-shot noise and spin-projection noise is given by
ηB=(δB0)mintmωmβ+1cos4(4g2Tωm)8gμBglz0TCβ,
where, the measurement time tm is approximated by T. Note that the MFG sensitivity in Eq. (36) depends on the number of photons collected per measurement β, resonance contrast C, measurement time, and the spin-mechanical coupling. For an observed contrast of 5% and count rate of 10 kcount s−1 [20], the MFG sensitivity is shown in Fig. 7. It can be inferred from Fig. 7 that increase in spin-mechanical coupling results in decrease in MFG sensitivity with a value reaching upto 100μTm1/Hz at g = 10ωm. Further, we find that the above analysis remains unaffected even if the initial mechanical coherent state is replaced by a thermal state [24, 25].

 figure: Fig. 6

Fig. 6 Schematic of the Ramsey microwave pulse sequence.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 MFG sensitivity versus scaled spin-mechanical coupling. Parameters chosen are C=5%, β=10 kcount s−1, and ωm =38 kHz.

Download Full Size | PDF

4.1. Experimental accessibility

In the above discussion, we did not include the decoherence from the mechanical oscillator as the decoherence times are substantially longer than the free evolution time T. Thus, we can neglect the mechanical decoherence. In our analysis above, we considered a nanodiamond of radius 50 nm optically levitated in a dipole trap in ultrahigh vacuum along with feedback cooling [20] and oscillating at a mechanical frequency ωm = 38 kHz. For such a configuration, the decoherence time corresponding to gas damping is of the order of Tg ~ 106 s which is quite large as compare to T (∼200 μs). So, the decoherence due to gas damping can be neglected. Further, the feedback induced decoherence [20] operates over a longer timescale (Tfb ~ ms) than the free evolution time (200 μs) and its effect can also be neglected. Furthermore, in our model, the decoherence due to optical scattering (Topt=γopt1~ms) can also be ignored for the duration T~200μs. Other sources of decoherence include the dephasing of the NV center due to interaction with other (non-NV) spins in the diamond lattice. However, the associated dephasing time (∼ 400 ms [61]) is still larger so as not to affect the MFG sensitivity.

5. Conclusion

In conclusion, an NV-center based hybrid nanomechanical magnetometer is proposed where a spin degree of freedom is coupled to a mechanical degree of freedom by an MFG. We have described two schemes for MFG sensing: first by analyzing the position spectrum of the mechanical oscillator, and second by manipulating the spin degrees of freedom. The first procedure gives a MFG sensitivity of 1μTm1/Hz whereas the latter provides 100μTm1/Hz. We have explained in detail that such an experimentally accessible nanomagnetometer functions well under both cooling and ambient conditions and thereby provides a platform for sensitive magnetometry applications. Our scheme can be extended to performing vector magnetic gradiometry by coupling the three independent directions of nanodiamond oscillations to the three cartesian components of MFG.

Funding

Office of Naval Research (ONR) (N00014-14-1-0803).

References and links

1. T. J. Kippenberg and K. J. Vahala, “Cavity optomechanics: Back-Action at the mesoscale,” Science 321(5893), 1172–1176 (2008). [CrossRef]   [PubMed]  

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

3. 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,” Nat. Lett. 475(7536), 359–363 (2011). [CrossRef]  

4. J. Chan, T. P. Mayer 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,” Nat. Lett. 478(7367), 89–92 (2011). [CrossRef]  

5. P. Meystre, “A short walk through quantum optomechanics,” Ann. Phys. (Berlin) 525(3), 215–233 (2012). [CrossRef]  

6. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86(4), 1391–1452 (2014). [CrossRef]  

7. A. D. Armour, M. P. Blencowe, and K. C. Schwab, “Entanglement and decoherence of a micromechanical resonator via coupling to a cooper-pair box,” Phys. Rev. Lett. 88(14), 148301 (2002). [CrossRef]   [PubMed]  

8. D. E. Chang, C. A. Regal, S. B. Papp, D. J. Wilson, J. Ye, O. Painter, H. J. Kimble, and P. Zoller, “Cavity optomechanics using optically levitated nanosphere,” Proc. Natl. Acad. Sci. U.S.A. 107(3), 1005–1010 (2009). [CrossRef]  

9. L. P. Neukirch and A. N. Vamivakas, “Nano-optomechanics with optically levitated nanoparticles,” Contemp. Phys. 56(1), 48–62 (2014).

10. Z. Q. Yin, A. A. Geraci, and T. Li, “Optomechanics of levitated dielectric particles,” Int. J. Mod. Phys. B 27(26), 1330018 (2013). [CrossRef]  

11. O. Romero-Isart, L. Clemente, C. Navau, A. Sanchez, and J. I. Cirac, “Quantum magnetomechanics with levitating superconducting microspheres,” Phys. Rev. Lett. 109(14), 147205 (2012). [CrossRef]   [PubMed]  

12. M. Cirio, G. K. Brennen, and J. Twamley, “Quantum magnetomechanics: Ultrahigh-Q-levitated mechanical oscillators,” Phys. Rev. Lett. 109(14), 147206 (2012). [CrossRef]   [PubMed]  

13. T. Li, S. Kheifets, and M. G. Raizen, “Millikelvin cooling of an optically trapped microsphere in vacuum,” Nat. Phys. 7(7), 527–530 (2011). [CrossRef]  

14. B. Rodenburg, L. P. Neukirch, A. N. Vamivakas, and M. Bhattacharya, “Quantum model of cooling and force sensing with an optically trapped nanoparticle,” Optica 3(3), 318 (2016). [CrossRef]  

15. 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]  

16. O. Romero-Isart, M. L. Luan, R. Quidant, and J. I. Cirac, “Toward quantum superposition of living organisms,” New J. Phys. 12, 033015 (2010). [CrossRef]  

17. O. Romero-Isart, A. C. Pflanzer, F. Blaser, R. Kaltenbaek, N. Kiesel, M. Aspelmeyer, and J. I. Cirac, “Large quantum superpositions and interference of massive nanometer-sized objects,” Phys. Rev. Lett. 107(2), 020405 (2011). [CrossRef]   [PubMed]  

18. 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]  

19. G. Ranjit, D. P. Atherton, J. H. Stutz, M. Cunningham, and A. A. Geraci, “Attonewton force detection using micro-spheres in a dual-beam optical trap in high vacuum,” Phys. Rev. A 91(5), 051805 (2015). [CrossRef]  

20. L. P. Neukirch, E. Hartmaan, J. M. Rosenholm, and A. N. Vamivakas, “Multi-dimensional single-spin nano-optomechanics with levitated nanodiamond,” Nat. Phton. 9, 653 (2015). [CrossRef]  

21. T. M. Hoang, J. Ahn, J. Bang, and T. Li, “Electron spin control of optically levitated nanodiamonds in vacuum,” Nat. Commun. 7, 12250 (2016). [CrossRef]   [PubMed]  

22. R. M. Pettit, L. P. Neukirch, Y. Zhang, and A. N. Vamivakas, “Coherent control of a single nitrogen-vacancy center spin in optically levitated nanodiamond,” J. Opt. Soc. Am. B 34(6), C31–C35 (2017). [CrossRef]  

23. 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]  

24. M. Scala, M. S. Kim, G. W. Morley, P. F. Barker, and S. Bose, “Matter-wave interferometry of a levitated thermal nano-oscillator induced and probed by a spin,” Phys. Rev. Lett. 111(18), 180403 (2013). [CrossRef]   [PubMed]  

25. C. Wan, M. Scala, S. Bose, A. C. Frangeskou, ATM A. Rahman, G. W. Morley, P. F. Barker, and M. S. Kim, “Tolerence in the Ramsey interference of a trapped nanodiamond,” Phys. Rev. A 93(4), 043852 (2016). [CrossRef]  

26. Q. Hou, W. Yang, C. Chen, and Z. Yin, “Generation of macroscopic Schrödinger cat states in diamond mechanical resonator,” Sci. Rep. 6, 37542 (2016). [CrossRef]  

27. A. Albrecht, A. Retzker, and M. B. Plenio, “Testing quantum gravity by nanodiamond interferometry with nitrogen-vacancy centers,” Phys. Rev. A 90(3), 033834 (2014). [CrossRef]  

28. O. Arcizet, V. Jacques, A. Siria, P. Poncharal, P. Vincent, and S. Seidelin, “A single nitrogen-vacancy defect coupled to a nanomechanical oscillator,” Nat. Phys. 7(11), 879–883 (2011). [CrossRef]  

29. S. Kolkowitz, A. C. B. Jayich, Q. P. Unterreithmeier, S. D. Bennett, P. Rabl, J. G. E. Harris, and M. D. Lukin, “Coherent sensing of a mechanical resonator with a single-spin qubit,” Science 335(6076), 1603–1606 (2012). [CrossRef]   [PubMed]  

30. P. Ovartchaiyapong, K. W. Lee, B. A. Myers, and A. C. B. Jayich, “Dynamic strain-mediated coupling of a single diamond spin to a mechanical resonator,” Nat. Commun. 5, 4429 (2014). [CrossRef]   [PubMed]  

31. D. A. Golter, T. Oo, M. Amezcua, K. A. Stewart, and H. Wang, “Optomechanical quantum control of nitrogen-vacancy center in diamond,” Phys. Rev. Lett. 116(14), 143602 (2016). [CrossRef]   [PubMed]  

32. P. Rabl, P. Cappellaro, M. V. Gurudev Dutt, L. Jiang, J. R. Maze, and M. D. Lukin, “Strong magnetic coupling between an electronic spin qubit and a mechanical resonator,” Phys. Rev. B 79(4), 041302 (2009). [CrossRef]  

33. J. Q. Zhang, S. Zhang, J. H. Zou, L. Chen, W. Yang, Y. Li, and M. Feng, “Fast optical cooling of nanomechanical cantilever with the dynamic zeeman effect,” Opt. Express 21(24), 29695–29710 (2013). [CrossRef]  

34. W. Ge and M. Bhattacharya, “Single and two-mode mechanical squeezing of an optically levitated nanodiamond via dressed-state coherence,” New J. Phys. 18, 103002 (2016). [CrossRef]  

35. S. D. Bennett, N. Y. Yao, J. Otterbach, P. Zoller, P. Rabl, and M. D. Lukin, “Phonon-induced spin-spin interactions in diamond nanostructures: Application to spin squeezing,” Phys. Rev. Lett. 110(15), 156402 (2013). [CrossRef]   [PubMed]  

36. K. Xia and J. Twamley, “Generating spin squeezing states and Greenberger-Horne-Zeilinger entanglement using hybrid phonon-spin ensemble in diamond,” Phys. Rev. B 94(20), 205118 (2016). [CrossRef]  

37. L. Rondin, J. -P. Tetienne, T. Hingant, J. -F. Roch, P. Maletinsky, and V. Jacques, “Magnetometry with nitrogen-vacancy defects in diamond,” Rep. Prog. Phys. 77, 056503 (2014). [CrossRef]   [PubMed]  

38. S. Forstner, S. Prams, J. Knittel, E. D. van Ooijen, J. D. Swain, G. I. Harris, A. Szorkovszky, W. P. Bowen, and H. R. Dunlop, “Cavity optomechanical magnetometer,” Phys. Rev. Lett. 108(12), 120801 (2012). [CrossRef]   [PubMed]  

39. S. Forstner, E. Sheridan, J. Knittel, C. L. Humphreys, G. A. Brawley, H. R. Dunlop, and W. P. Bowen, “Ultrasensitive optomechanical magnetometry,” Adv. Mater. 26(36), 6348–6353 (2014). [CrossRef]   [PubMed]  

40. C. Yu, J. Janousek, E. Sheridan, D. L. McAuslan, H. R. Dunlop, P. K. Lam, Y. Zhang, and W. P. Bowen, “Optomechanical magnetometry with a macroscopic resonator,” Phys. Rev. App. 5(4), 044007 (2016). [CrossRef]  

41. M. V. Romalis and H. B. Dang, “Atomic magnetometers for materials characterization,” Mater. Today 14(6), 258–262 (2011). [CrossRef]  

42. D. Budker, D. F. Kimball, S. M. Rochester, V. V. Yashchuk, and M. Zolotorev, “Sensitive magnetometry based on nonlinear magneto-optical rotation,” Phys. Rev. A 62(4), 043403 (2000). [CrossRef]  

43. A. Trabesinger, “Magnetic resonance imaging: Take a close look,” Nat. Phys. 3, 302 (2007). [CrossRef]  

44. B. Polyak, I. Fishbein, M. Chorny, I. Alferiev, D. Williams, B. Yellen, G. Friedman, and R. J. Levy, “High field gradient targeting of magnetic nanoparticle-loaded endothelial cells to the surfaces of steel stents,” Proc. Nat. Acad. Sci. U.S.A. 105(2), 698–703 (2007). [CrossRef]  

45. D. Rugar, R. Budakian, H. J. Mamin, and B. W. Chui, “Single spin detection by magnetic resonance force microscopy,” Nat. Lett. 430, 329–332 (2004). [CrossRef]  

46. J. R. Maze, P. L. Stanwix, J. S. Hodges, S. Hong, J. M. Tylor, P. Cappellaro, L. Jiang, M. V. G. Dutt, E. Togan, A. S. Zibrov, A. Yacoby, R. L. Walsworth, and M. D. Lukin, “Nanoscale magnetic sensing with an individual electronic spin in diamond,” Nat. Lett. 455, 644–647 (2008). [CrossRef]  

47. D. Budker and M. Romalis, “Optical magnetometry,” Nat. Phys. 3(4), 227–234 (2007). [CrossRef]  

48. W. Ge, B. Rodenburg, and M. Bhattacharya, “Feedback-induced bistability of an optically levitated nanoparticle: A Fokker-Planck treatment,” Phys. Rev. A 94(2), 023808 (2016). [CrossRef]  

49. T. M. Hoang, Y. Ma, J. Ahn, J. Bang, F. Robicheaux, Z.-Qi Yin, and T. Li, “Torsional optomechanics of a levitated nonspherical nanoparticle,” Phys. Rev. Lett. 117(12), 123604 (2016). [CrossRef]   [PubMed]  

50. T. Delord, L. Nicolas, L. Schwab, and G. Hétet, “Electron spin resonance from NV centers in diamonds levitating in an ion trap,” New. J. Phys. 19, 033031 (2017). [CrossRef]  

51. T. Delord, L. Nicolas, Y. Chassagneux, and G. Hétet, “Strong coupling between a single NV spin and the torsional mode of diamonds levitating in an ion trap,” arXiv:1702.00774.

52. Y. Ma, T. M. Hoang, M. Gong, T. Li, and Z.-q Yin, “Quantum many-body simulation and torsional matter-wave interferometry with a levitated nanodiamond,” arxiv:1611.05599.

53. Y. Arita, M. Mazilu, and K. Dholakia, “Laser induced rotation and cooling of a trapped microgyroscope in vacuum,” Nat. Commun. 4, 2374 (2013). [CrossRef]  

54. D. Vitali, S. Mancini, L. Ribichini, and P. Tombesi, “Mirror quiescence and high-sensitivity position measurements with feedback,” Phys. Rev. A 65(6), 063803 (2002). [CrossRef]  

55. X. Xu and J. M. Taylor, “Squeezing in a coupled two-mode optomechanical system for force sensing below the standard quantum limit,” Phys. Rev. A 90(4), 043848 (2014). [CrossRef]  

56. L. M. Pham, “Magnetic field sensing with nitrogen-vacancy color centers in diamond,” (Doctoral Dissertation, Harvard Uni., 2013).

57. J. M. Taylor, P. Cappellaro, L. Childress, L. Jiang, D. Budker, P. R. Hemmer, A. Yacoby, R. Walsworth, and M. D. Lukin, “High-sensitivity diamond magnetometer with nanoscale resolution,” Nat. Phys. 4, 810–816 (2008). [CrossRef]  

58. I. M. Savukov, S. J. Seltzer, M. V. Romalis, and K. L. Sauer, “Tunable atomic magnetometer for detection of radio-frequency magnetic fields,” Phys. Rev. Lett. 95(6), 063004 (2005). [CrossRef]   [PubMed]  

59. W. M. Itano, J. C. Bergquist, J. J. Bollinger, J. M. Gilligan, D. J. Heinzen, F. L. Moore, M. G. Raizen, and D. J. Wineland, “Quantum projection noise: population fluctuations in two-level systems,” Phys. Rev. A 47(5), 3554–3570 (1993). [CrossRef]   [PubMed]  

60. K. Fang, V. M. Acosta, C. Santori, Z. Huang, K. M. Itoh, H. Watanabe, S. Shikata, and R. G. Beausoleil, “High-sensitivity magnetometry based on quantum beats in diamond nitrogen-vacancy centers,” Phys. Rev. Lett. 110(13), 130802 (2013). [CrossRef]   [PubMed]  

61. B. Naydenov, F. Dolde, L. T. Hall, C. Shin, H. Fedder, L. C. L. Hollenberg, F. Jelezko, and J. Wrachtrup, “Dynamical decoupling of a single-electron spin at room temperature,” Phys. Rev. B 83(8), 081201 (2011). [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 (7)

Fig. 1
Fig. 1 Schematic of the optically levitated nanodiamond oscillating in dipole trap along z-direction. A green laser of wavelength 532 nm excites the nanodiamond. A MFG along z-direction is applied to engineer spin-mechanical coupling. Microwaves are used to manipulate the ground spin state of nanodiamond.
Fig. 2
Fig. 2 (a) Level diagram for excitation and read-out of NV-spins by optical means. Inset shows the bare energy-level diagram for the triplet ground-state of NV-center in nanodiamond. The microwave fields of Rabi frequency Ω0 drives the transitions |0〉 → |+ 1〉 and |0〉 → | − 1〉 with detuning Δ. (b) The dressed-state description of NV-spin. Phonons of energy ωm drives the transitions |c〉 → |b〉 and |b〉 → |a〉 with effective detunings Δ1 = ωmωbc and Δ2 = ωmωab, respectively.
Fig. 3
Fig. 3 (a) Position power spectral density for z-component of motion versus frequency. The parameters chosen are ωm = 38kHz, R =50 nm, density=2200 kg/m3, Teff = 4 K, χ = 10−7, G = 1/18, Ω0 = 0.8ωm, Γ1 = 0.25ωm. (b) Position power spectral density versus scaled spin-mechanical coupling at ω = ωm. The inset shows the position PSD corresponding to red ellipse for smaller values of g/ωm.
Fig. 4
Fig. 4 MFG sensitivity versus (a) frequency (b) scaled spin-mechanical coupling for t m = 2 π ω m. Other parameters are same as in Fig. 3.
Fig. 5
Fig. 5 (a) Steady state phonon number versus g/ωm and Pressure (mbar). MFG sensitivity versus (b)g/ωm and Pressure (mbar), (c) g/ωm and Temperature (K). In plots (a) and (b) Teff =300 K, γopt =1 kHz, while in (c) pressure is 0.3 mbar. The other parameters are same as in Fig. 3.
Fig. 6
Fig. 6 Schematic of the Ramsey microwave pulse sequence.
Fig. 7
Fig. 7 MFG sensitivity versus scaled spin-mechanical coupling. Parameters chosen are C = 5 %, β=10 kcount s−1, and ωm =38 kHz.

Equations (36)

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

ρ ˙ = 1 i [ H m , ρ ] ( A t + A p + D p 2 ) D [ Q z ] ρ ( t ) D q 2 D [ P z ] ρ ( t ) i η f 4 m [ Q z , { P z , ρ } ] + [ ρ ( t ) ] ,
| a = sin ( θ ) | 0 + cos ( θ ) | + ; ω a = ( Δ + Δ 2 + 2 Ω 0 2 ) / 2 , | b = | ; ω b = Δ , | c = cos ( θ ) | 0 sin ( θ ) | + ; ω c = ( Δ Δ 2 + 2 Ω 0 2 ) / 2 ,
ρ ˙ = 1 i [ H m , ρ ] ( A t + A p + D p 2 ) D [ Q z ] ρ ( t ) D q 2 D [ P z ] ρ ( t ) i η f 4 m [ Q z , { P z , ρ } ] + [ ρ ( t ) ] A 2 D [ a m ] ρ A + 2 D [ a m ] ρ i δ 2 [ a m a m , ρ ] ,
α 1 = Re [ sin 2 ( θ ) N ( p 2 ρ c c + p 3 ρ c a ) + cos 2 ( θ ) N p 1 ρ b b ] , α 2 = Re [ cos 2 ( θ ) N ( p 1 ρ a a + p 3 ρ a c ) + sin 2 ( θ ) N p 2 ρ b b ] , α 3 = Im [ sin 2 ( θ ) ( p 2 ( ρ c c ρ b b ) + p 3 ρ c a ) N + cos 2 ( θ ) ( p 1 ( ρ a a ρ b b ) + p 3 ρ a c ) N ] ,
Q ˙ z = ( ω m + δ 2 ) P z ( A A + 2 ) Q z ,
P ˙ z = ( ω m + δ 2 ) Q z [ Γ + ( A A + 2 ) ] Q z + 2 m Γ 0 k B T e f f ξ T ( t ) + 18 η Φ n G 2 Q 4 ξ F ( t ) + [ 2 A + + N ( 2 A + 2 A ) ] ξ S ( t ) ,
q ¨ z + [ Γ + A A + ] q ˙ z + { ( ω m + δ 2 ) 2 + [ Γ + ( A A + 2 ) ] ( A A + 2 ) } q z = ( F T + F F + F S ) m ,
S T = ( 1 + δ 2 ω m ) 2 m Γ 0 k B T e f f ,
S F = ( 1 + δ 2 ω m ) 54 m ω m χ 2 Φ G 2 ( 2 N 2 + 2 N + 1 ) ,
S S = ( 1 + δ 2 ω m ) m ω m [ A + + N ( A + A ) ] .
q ˜ z ( ω ) = χ m ( ω ) [ F ˜ T + F ˜ F + F ˜ S ] ,
χ m ( ω ) = 1 m { [ ( ω m + δ 2 ) 2 + ( A A + 2 ) ( Γ + A A + 2 ) ω 2 ] i ( Γ + A A + ) ω } .
| δ q ˜ z ( ω ) | 2 = | χ m ( ω ) | 2 [ | F ˜ T | 2 + | F ˜ F | 2 + | F ˜ S | 2 ] .
| δ q ˜ z ( ω ) | 2 = | χ m ( ω ) | 2 [ S T + S F + S S ] + z 0 2 χ 2 Φ .
α = 4 Γ 1 ( 1 + cos ( 2 θ ) ) sin 2 ( 2 θ ) 9 cos 2 ( 2 θ ) .
| χ m ( ω m ) | m a x 2 = [ m 2 { ( Γ + A ) 2 [ ( ω m 2 + A 2 ( Γ + A 2 ) ) 1 2 ( Γ + A ) 2 ] } ] 1 .
η B = | δ q ˜ z ( ω ) | 2 | δ q ˜ z ( ω ) | 2 B 0 × t m ,
η B = | δ q ˜ z ( ω ) | 2 μ B g l z 0 | χ m ( ω ) | 2 [ D + m ω m N α ] ,
D = 2 m 2 g α ( Γ + A ) | χ m ( ω ) | 2 ( S T + S F + S S ) [ ω m 2 + ω 2 + A 2 ( Γ + A 2 ) ] .
N ˙ = 2 J N 2 ( J + K ) N + ,
N s s D p + γ o p t + A + 2 J .
η B = μ B g l z 0 | δ q ˜ z ( ω ) | 2 | χ m ( ω ) | 2 ( D 1 + D 2 + D 3 ) + D 4 ( S T + S F + S S ) ,
D 1 = 2 g α 3 ω m 2 m Γ 0 K B T e f f ,
D 2 = 108 m g ω m χ 2 Φ G 2 [ α 3 ω m ( 2 N 2 + 2 N + 1 ) + ( 1 + δ 2 ω m ) α 2 J ( 2 + 1 N ) ] ,
D 3 = m ω m 2 g [ α 3 ω m { A + + N ( A + A ) } + ( 1 + δ 2 ω m ) { α 2 A + + N ( α 2 A + α 1 A ) + 1 2 N α 2 J ( A + A ) } ] ,
D 4 = 8 m 2 g | χ m ( ω ) | 4 { [ ( ω m + δ 2 ) 2 + ( A A + 2 ) ( Γ + A A + 2 ) ω 2 ] × [ α 3 ( ω m + δ 2 ) + ( α 1 + α 2 2 ) ( Γ + A A + ) ] + ( α 1 α 2 ) ( Γ + A A + ) ω 2 } .
| ψ 1 = U π / 2 | ψ ( 0 ) = ( | 0 i | + 1 2 ) | α .
| ψ 2 ( t ) = U t | ψ 1 = ( | 0 | α ( t , 0 ) i | + 1 | α ( t , + 1 ) 2 ) ,
| ψ 2 ( T ) = ( | 0 i exp ( 4 i g 2 T ω m ) | + 1 2 ) | α .
| ψ 3 = 1 2 [ ( 1 exp ( 4 i g 2 T ω m ) ) | 0 i ( 1 + exp ( 4 i g 2 T ω m ) ) | 1 ] | α .
P 0 = 1 2 ( 1 cos ( 4 g 2 T ω m ) ) .
S = β 0 + β 1 2 β 0 β 1 2 cos ( 4 g 2 T ω m ) ,
δ B 0 = δ B 0 p s n 2 + δ B 0 s p n 2 ,
( δ B 0 ) min = β + 1 cos 4 ( 4 g 2 T ω m ) max | S B 0 | .
( δ B 0 ) min = ω m β + 1 cos 4 ( 4 g 2 T ω m ) 8 g μ B g l z 0 T C β ,
η B = ( δ B 0 ) min t m ω m β + 1 cos 4 ( 4 g 2 T ω m ) 8 g μ B g l z 0 T C β ,
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.