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

Ground-state cooling of mechanical oscillator via quadratic optomechanical coupling with two coupled optical cavities

Open Access Open Access

Abstract

We present a scheme for the electromagnetically induced transparency (EIT)-like nonlinear ground-state cooling in a double-cavity optomechanical system in which an optical cavity mode is coupled parametrically to the square of the position of a mechanical oscillator, an additional auxiliary cavity is coupled to the optomechanical cavity. The optimum cooling conditions is derived, based on which the heating process can be well suppressed and the mechanical resonator can be cooled with an optimal effect to near its ground state through EIT-like cooling mechanism even in unresolved sideband regime. It is demonstrated by numerical simulations that not only the average phonon number of steady state is lower than that of single-cavity optomechanical system, but also the cooling rate is greatly faster than that of the linear optomechanical coupling due to the two-phonon cooling process in the quadratic coupling. Also, the ground-state cooling is achievable even with a relatively weak quadratic coupling strengthby tunning the coupling between two cavities to reach the optimum cooling conditions, thus provides an solution for overcoming the limitations of weak quadratic coupling rate in experiments. The proposed approach provides a platform for quantum manipulation of macroscopic mechanical devices beyond the resolved sideband limit and linear coupling regime.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Cavity optomechanics as a crucial branch of physics which focuses on the interaction of controllable radiation pressure between light and mechanical motion has aroused wide interest in recent years [1, 2]. Cooling the mechanical resonators to the quantum ground-state provides a crucial application prospect in cavity optomechanics [3, 4], and has been widely used to quantum information processing [5–10], high-precision control and measurement [11–14], and quantum-classical boundary probe of quantum behaviour of macroscopic system [15–17], etc. So far, various methods have been used for cooling a mechanical resonator to its ground-state, such as pure cryogenic cooling [18], feedback cooling [19, 20],and sideband cooling [21–26]. Among them, the sideband cooling is the most common method which require the condition of resolved sideband limit to be fulfilled, i.e. the decay rate of the cavity mode is much less than the mechanical frequency. This is a strict restriction since a high quality factor is needed for the cavity and it is hard to implement in most physical systems. Once the condition of sideband cooling cannot be satisfied by the cavity mode, the Stokes heating cannot be well suppressed, which will result in a failure in cooling. In contrast, the restriction for implementing ground-state cooling is relaxed in the unresolved sideband regime which can be achievable with a bad cavity. Recently some approaches have been proposed such as the dissipative coupling mechanism [27–31], optomechanically induced transparency [32, 33], atom-optomechanical systems [34–36], and coupled-cavity configurations [37–39], etc. These schemes are promising to be realized in unresolved sideband conditions where the decay rate of cavity mode would be larger than the mechanical frequency.

In most optomechanical theoretical and experimental studies, researchers mainly focused on the linear coupling optomechanical system, where the optical cavity mode is parametrically coupled to the position of a mechanical oscillator. However, some studieshave started to focus on the nonlinear mechanical cooling process, for instance, the cooling in high optomechanical coupling regime of the single-photon [40] and the secondary sideband laser cooling of trapped ions [41]. This leads to nonthermal steady state, even nonclassical sub-Poissonian state. Quantum systems far from thermal equilibrium have great prospects for fundamental physics research and the realization of practical quantum devices [42, 43]. At present, a neotype of optomechanical system, quadratic optomechanics (also known as “quadratic nonlinearity”) has been proposed, where the coupling term is proportional to the square of mechanical displacement x2. It’s realizable in the setups of membrane-in-the-middle geometry [44, 45], which is placed at a node or antinode of the cavity field as well as the ultracold atom cloud loaded into the optical cavity, the centroid coordinate of the cloud is regarded as the mechanical degree of freedom [46, 47]. Up to now, theoretical literature has focused on the detection of phonon Fock states by using quadratic coupling, as well as the cooling and squeezing of mechanical oscillators [48–50]. At the same time, electromagnetically induced transparency (EIT) or optomechanically induced transparency (OMIT) have also been investigated in quadratic coupled optomechanical systems, which means two phonon processes [51, 52]. The development of experimental technology has made it become more and more attractive to investigate the nonlinear cooling process caused by intrinsic quadratic coupling. However, the main limitation in experiment comes from the weak quadratic coupling rate when quadratic coupling is used. Hence, many efforts have been devotedto improve the nonlinear coupling strength in the optomechanical system [53]. Enlightened by theses works, here we propose an EIT-like ground-state cooling scheme for the mechanical oscillator in a coupled-cavity optomechanical system with weak quadratic coupling. In our scheme, the first cavity is coupled parametrically to the square of the position of a mechanical oscillator via the radiation pressure force. We demonstrate that the cooling of mechanical oscillator can be achieved even if the optomechanical cavity decay is in unresolved sideband regime due to the coherent coupling to the auxiliary optical cavity mode. By analyzing the optical force spectrum and deriving the optimal cooling conditions, the EIT-like spectrum splits from the single Lorentzian peak of the standard optomechanical cavity into two narrower peaks with a dip emerging between them, which leads to more efficient cooling for a weak optomechanical coupling strength. Compared with the previous works involving single-cavity quadratic optomechanical coupling [54, 55], an auxiliary cavity is included in our present system which brings out the quantum interference effect [56], thus is more efficient and the mechanical mode can be cooled to a lower mean phonon number even when the resolved sideband condition is not fulfilled. In addition, different from the linear optomechanical coupling scheme with double cavities [37], our present scheme with quadratic coupling can cool the oscillator to its ground state with a faster velocity due to the two-phonon absorption. The cooling is achievable even with a relatively weak quadratic coupling strength by tunning the coupling between two cavities to reach the optimum cooling conditions, thus provides an useful way to overcome the limitations of weak quadratic coupling rate in experiments.

This paper is organized as follows. In Sec. 2, the quadratic optomechanical system is introduced and the effective Hamiltonian is given. In Sec. 3, the rate equation for the phonon of the mechanical oscillator is given and the properties of optical force spectrum are derived. In Sec. 4, the optimum cooling conditions are obtained by parameters modulation, and the cooling limits in weak quadratic coupling condition are discussed. In Sec. 5, cooling dynamics of the system is investigated by using master equation and numerical simulations. In Sec. 6, we give some discussions on experimental feasibility. Finally, a brief summary of our work is presented in Sec. 7.

 figure: Fig. 1

Fig. 1 Schematic of the optomechanical system. A membrane oscillator is placed in the middle of cavity 1 which is driven by a continuous-wave input laser, the coupling between them is proportional to the square of position of the oscillator. The cavity2 as an auxiliary cavity only directly couples to the optomechanical cavity 1 with coupling strength J.

Download Full Size | PDF

2. Model and Hamiltonian

We consider a optomechanical system including double coupled cavities as shown in Fig. 1, where the optomechanical cavity is coupled parametrically to the square of the position of a mechanical oscillator and driven by an continuous-wave input laser, the second cavity as an auxiliary cavity couples directly to the first one with the coupling strength J. The total system Hamiltonian can be written as (=1)

H=ω1a1a1+ω2a2a2+ωmbb+ga1a1(b+b)2+Ω(a1eiωLt+a1eiωLt)+J(a1a2+a1a2),
in which a1, a2 and b denote the annihilation operators for the two optical cavity modes and the mechanical oscillator, respectively, with the frequencies ω1, ω2 and ωm; g is the quadratic optomechanical coupling coefficient; ωL is the frequency of the input laser and Ω is the pumping strength. In the frame rotating at the input laser frequency ωL, the system Hamiltonian becomes
H=Δ1a1a1+Δ2a2a2+ωmbb+ga1a1(b+b)2+Ω(a1+a1)+J(a1a2+a1a2),
where Δi=ωiωL,(i=1,2) are detunnings between the laser and two cavities. To obtain the linearized Hamiltonian the three modes are split into an average amplitude and a fluctuation term, i.e., a1α1+a1, a2α2+a2 and bβ+b, where α1, α2 and β denote the steady-state displacements of the optical and mechanical modes which can be calculated as
α1=iΩ(iΔ1κ12)+J2iΔ2κ22,α2=iJα1iΔ2κ22,β=0.

After the standard linearization procedure, the effective linearized Hamiltonian can be expressed as

Heff=Δ1a1a1+Δ2a2a2+ωmbb+(Ga1+G*a1)(b+b)2+J(a1a2+a1a2),
where G=α1g is the cavity-enhanced optomechanical coupling strength.

The energy levels of the system are presented in Fig. 2. Due to the coupling between two cavities, quantum interference may take place among different excitation pathways. For example, quantum interference occurs between the excitation paths |1|2 and |1|2|3|3 which is corresponding to the cooling process for the mechanical mode. On the other side, that occurs between the paths |1|2 and |1|2|3|3 which is corresponding to the heating process for the mechanical mode and should be suppressed in order to get a efficient cooling. Next we are going to harness the quantum interference to suppress the heating process and enhance the cooling process, which can be achievable effectively in the current scheme even in the unresolved sideband regime κ1>ωm.

 figure: Fig. 2

Fig. 2 Energy level diagram of the system in the displaced frame. |n1,n2,m donates the state of n1 photons in mode a1, n2 photons in mode a2, and m phonons in mode b. The orange double arrow denotes the couplingbetween two cavities with coupling strength J.

Download Full Size | PDF

3. Optical force spectrum and quantum rate equation

In order to show the cooling regime, we calculate the optical force spectrum. In the weak-coupling regime, the optical force spectrum can be analyzed by using perturbation theory. Regarding the optomechanical coupling as perturbation to the optical field, the optical force spectrum SFF(ω)=F(ω)F(ω)dω can be calculated from the optical part of the linearized Hamiltonian Eq. (4) without coupling to the mechanical oscillator

Hop=Δ1a1a1+Δ2a2a2+J(a1a2+a1a2),
and the optical force is described by F=(G*a1+Ga1)/xZPF2 from the Eq. (4) with xZPF denoting the zero-point fluctuation of the oscillator. Then SFF(ω) can be obtained from the corresponding quantum Langevin equations in the frequency domain
iωa1(ω)=(iΔ1κ12)a1(ω)iJa2(ω)κ1ain,1(ω),iωa2(ω)=(iΔ2κ22)a2(ω)iJa1(ω)κ2ain,2(ω),
where the corresponding noise operators ain,1, ain,2 satisfy the correlation in time domain ain,1(t)ain,1(t)=ain,2(t)ain,2(t)=δ(tt). Calculating from the Eq. (6) we obtain
a1(ω)=κ1ain,1(ω)iJχ2(ω)κ2ain,2(ω)χ(ω),
with
χ1(ω)=1κ1/2i(ωΔ1),χ2(ω)=1κ2/2i(ωΔ2),χ(ω)=1χ1(ω)+J2χ2(ω).

As a result, we have

SFF(ω)=|G|2xZPF4[1χ(ω)+1χ*(ω)].

The two-phonon cooling and amplification rates can be obtained from Fermi’s Golden rule by using the methods given in Ref. [57]. Concentrating on the diagonal terms of the density matrix ρnn=Pn, we write down a set of rate equations

Pn˙=γm[nth(n+1)+(nth+1)n]Pn+γmnthnPn1+γm(nth+1)(n+1)Pn+1[An(n1)+A+(n+2)(n+1)]Pn+A(n+2)(n+1)Pn+2+A+n(n1)Pn2,
where Pn is the phonon number distribution in the n phonon state. The terms in the first two lines are due to the coupling of the mechanical oscillator to its thermal bath with rate γm and thermal phonon number nth. The last two lines describe the transition effects induced by the effective optomechanical coupling, where A=xZPF4SFF(±2ωm) stand for the cooling rate and heating rate for the mechanical oscillator. Then the net cooling rate is given as Γopt=AA+.

4. Cooling manipulation and cooling limits

4.1. Cooling improvement by parameter modulation

In order to obtain the optimal cooling effect, it is necessary to get the optimum optical coupling strength between the two optical cavities as well as the effective optical detunings.

For the case of single cavity (i.e. J = 0), the optical force spectrum SFF(ω) has a Lorentzian noise spectrum with single peak localed at ω=Δ1. According to the sideband cooling regime, it is necessary to meet the condition of κ1<ωm to realize ground-state cooling in the resolved sideband. In the unresolved sideband regime (κ1>ωm), the peaks of the cooling spectrum SFF(2ωm) and heating spectrum SFF(2omegam) are comparable, thus leads to the ground-state cooling for the mechanical oscillator unachievable. However, in our present scheme, the optical force spectrum SFF(ω) becomes a complex line shape due to the coupling between the two optical modes. We focus on the the double-cavity optomechanical system in the unresolved sideband regime (κ1>ωm).

 figure: Fig. 3

Fig. 3 (a) The optical force spectrum SFF (ω) (in arbitrary units) for different optical coupling strengths J, with Δ1 = 2ωm, κ2 = 0.01ωm. (b) The optical force spectrum SFF (ω) (in arbitrary units) for different detunings of the second cavity κ2 in the optimum conditions with Δ1 = −4ωm, J=26ωm. The other parameters are taken as Δ2 = −2ωm, κ1 = 10ωm, G = 0.08ωm.

Download Full Size | PDF

In Fig. 3(a), the optical force spectrum SFF(ω) is plotted versus the frequency ω with different optical coupling strength J. It can be seen that the single Lorentzian peak splits into two peaks with a dip between them due to the coupling between two optical cavities. The dip appears similar to the two-photon resonance in the electromagnetically induced transparency (EIT) of a three-level system [58]. Obviously, the minimal value of the optical force spectrum SFF(ω) locates at ω=Δ2=2ωm, corresponding to the two-photon resonance condition in EIT. In this case, the phenomenon of quantum destructive interference happens, thereby the heating process of mechanical oscillator is suppressed effectively. Therefore in order to locate the minimum value of the optical force spectrum SFF(ω) at ω=2ωm, the corresponding optimal condition is taken as Δ2=2ωm. In the meantime, it can be seen from the Fig. 3(a) that the two peaks of the optical force spectrum move with different optical coupling strength. In order to achieve the maximum cooling efficiency, it is necessary to locate the right-hand peak at ω=2ωm corresponding to the cooling process.

Next, we investigate the relation between the detuning Δ1 and the optical coupling strength J to implement the optimal cooling effect. In fact, the two split peaks of the optical force spectrum origin from the normal mode splitting owing to the coupling between the two cavities. So the optimal coupling strength can be ascertained by the eigen-energies of the dressed states of Hamiltonian Hop in Eq. (5),

E+=12(Δ1+Δ2+(Δ1+Δ2)2+4J2),E=12(Δ1+Δ2(Δ1+Δ2)2+4J2).

Here, E+ is corresponding to the right-hand peak of the optical force spectrum. In order to obtain the optimal cooling effect, the relation E+=2ωm should be satisfied. Then the optimal coupling strength can be obtained as

J=4ωm(2ωmΔ1).

With this optimal coupling strength in the Eq. (12) and Δ2=2ωm, in Fig. 3(b), the optical force spectrum SFF(ω) is illustrated under different decay rate κ2 of the auxiliary cavity. Here the effective detuning of the first cavity has been selected as Δ1=4ωm, which means the corresponding optimal coupling strength J=26ωm.

 figure: Fig. 4

Fig. 4 The net cooling rate Γopt = AA+ versus the dacay rate of the second cavity κ2 and the coupling strength J. Here the detuning and the decay rate of the first cavity are taken as Δ1 = −4ωm, κ1 = 10ωm, the detuning of the second cavity is taken as Δ2 = −2ωm, and the optomechanical coupling strength is taken as G = 0.08ωm.

Download Full Size | PDF

It’s worth to be noticed that even if the above optimal conditions are satisfied, i.e., Δ2=2ωm corresponding to a minimal heating effect and E+=2ωm corresponds to a maximal cooling effect, we still have to control the decay rate κ2 appropriately which directly determines the depth of the dip, in order to limit the minimal value of the optical force spectrum to be as close as possible to zero. Then the optimal ground cooling can be implemented. For this purpose the net cooling rate Γopt=AA+ versus the decay rate of the auxiliary cavity κ2 and the coupling strength J is plotted in Fig. 4. As can be seen from that, the net cooling rate increases with decreasing of the decay rate κ2 of the auxiliary cavity. That is because a smaller decay rate κ2 leads to a lower minimum of the dip in the optical force spectrum, even close to zero. In this way, the heating process can be significantly suppressed. According to the above conditions, when Δ1=4ωm, the position of the maximum cooling rate in the Fig. 4 corresponds to the proximity of J=26ωm, which conforms to the optimum conditions we deduced. Compared with the case of single cavity, i.e. J = 0, the net cooling rate is obviously much lower than that of our scheme.

 figure: Fig. 5

Fig. 5 A comparison of the net cooling rates Γopt=AA+ between the current coupled-cavity optomechanical system and the single-cavity optomechanical system. The former is plotted as the green circle curve under the optimum coupling conditions in Eq. (12) versus the detuning of the auxiliary cavity Δ2 for Δ1=4ωm, κ1=30ωm, κ2=0.01ωm, and the latter is plotted as the blue triangle curve versus Δ1 for κ=30ωm. The effective optomechanical coupled strength G=0.08ωm is taken.

Download Full Size | PDF

In order to illustrate the effect the auxiliary cavity, a comparison of the net cooling rates between the current system and the single optomechanical system is shown in the Fig. 5. The net cooling rate Γopt of versus the detuning of the auxiliary cavity Δ2 in our double-cavity system is plotted with the green circle line in the optimum conditions with Δ1=4ωm selected. In this case, the maximum cooling rate can be achieved by modulating the detunning of the auxiliary cavity. Compared with the net cooling rate for the single-cavity optomechanical system (plotted with the blue triangle line), it can be clearly seen that the optimal net cooling rate in the coupled cavity system is much higher than that of single cavity system. In our coupled cavity system, the minimum and maximum net cooling rate is located at the point Δ2=±2ωm, while in the case of single cavity, the peak of net cooling rate moves and becomes smooth under the unresolved sideband regime, thus the mechanical oscillator can hardly be cooled.

4.2. Cooling in weak quadratical coupling limits

Here we’re going to discuss the cooling limits of the quadratic optomechanical coupling. The infinite set of rate equations Eq. (10) can be replaced by a single differential equation for the generating function F(z,t)=n=0Pn(t)zn [59]. By solving the rate equations in the limit of strong optical damping and small thermal temperature, i.e. γmnthA±, only pure two-phonon transitions be obtained, we have

P2n+j(2)=(1m)mn(γ+j1)(1)j1,     j=0,1,
where the additional parameter γ is determined by the initial conditions, characterizing a relative weight of the odd oscillator phonon states, m=A+/A. When the above optimum conditions are satisfied, which will lead to AA+, i.e. m1. This means that the distribution of phonon number is only concentrated in zero-phonon state P0=1γ and one-phonon state P1=γ. This indicates that the cooling process of two-phonon maintains the parity of phonon number, making the initial odd phonon state cool to one-phonon state and even phonon state cool to zero-phonon state.

By using the optimal conditions Eq. (12), in the strong two-phonon absorption condition AγmnthA+, we obtain [59]

P0=1+2ξ1+3ξ+O(γmnthA),P1=ξ1+3ξ+O(γmnthA), ξ=nthnth+1,
with all other phonon number distribution Pn = 0 and the minimal mean phonon number n^=1/(4+1/nth).

5. Quantum master equation and numerical simulation

The quantum master equation of the system reads

ρ˙=i[Heff,ρ]+κ12(2a1ρa1a1a1ρρa1a1)+κ22(2a2ρa2a2a2ρρa2a2)+γm2(nth+1)(2bρbbbρρbb)+γm2nth(2bρbbbρρbb),
here, Heff is the linearized effective system Hamiltonian given by Eq. (4). nth=(exp(ωm/kBT)1)1 corresponds to the bath thermal phonon number at the environmental temperature T.

 figure: Fig. 6

Fig. 6 (a) The average phonon number bb of the steady state versus the detuning of the second cavity Δ2 and the optomechanical coupling strength G. The detuning of first cavity Δ1 and the coupling strength between two cavities J meet the optimum condition in Eq. (12). (b) The average phonon number bb versus the effective detuning of the first cavity Δ1 and the coupling strength J. Here the effective detuning of the second cavity Δ2=2ωm is taken. The white dotted curve satisfies the Eq. (12). The other parameters are taken as κ1=10ωm, κ2=0.01ωm, γm=106ωm, G=0.08ωm and nth=10.

Download Full Size | PDF

By employing the master equation, the numerical simulation for the average phonon number nf of the steady state versus the detunings of second optical cavities Δ2 and the optomechanical coupling strength G are shown in Fig. 6(a),where Δ1 satisfies the relation of optimal coupling strength in the Eq. (12). It can be found that the average phonon number bb is far less than unity for Δ2=2ωm, i.e. we can obtain the optimal cooling effect. In addition, it can be seen from the figure that the optomechanical coupling strength satisfies the weak coupling mechanism, i.e. Gκ1. When Δ2 satisfies the optimum condition, the optomechanical coupling strength also has an optimum range, that is, G/ωm<0.1. To further verify the above optimum condition, in the density plot Fig. 6(b), the average phonon number bb is shown as a function of effective detuning of the first cavity Δ1 and the coupling strength J. The white dotted line in this figure represents the derived optimum condition in the Eq. (12). Obviously, the optimal cooling effect obtained by numerical simulation always meet with this optimum cooling line.

 figure: Fig. 7

Fig. 7 A comparison of the cooling dynamics between the current coupled-cavity optomechanical system and the single-cavity optomechanical system. The numerical results of the average phonon number bb for the coupled-cavity system are plotted with the magenta diamond and green square curves, corresponding to the initial even and odd phonon numbers nth=10 and 11, respectively, with Δ1=4ωm, Δ2=2ωm, J=26ωm. The numerical results for single-cavity optomechanical system are plotted with orange circle and blue triangle curves corresponding to initial even and odd phonon numbers nth=10 and 11, respectively, with Δ1=2ωm. The shadow area denotes the optimal cooling region for bb<1.

Download Full Size | PDF

For the high-Q mechanical oscillator and lower initial temperature, we can ignore the effects of thermal damping, and the pure two-phonon transitions occur. Therefore, based on the characteristics of the two-phonon transitions, the quadratic coupling scheme proposed here has a faster cooling rate than the linear coupled double-cavity system [37], which comes from the nature of the quadratic coupling. In Fig. 7, the time evolutions of the final average phonon number bb in the cases of double-cavity and of single-cavity are numerically simulated, respectively. As mentioned above, the final mean phonon number distribution is concentrated in one-phonon state or zero-phonon state. As shown in Fig. 7, the two lines below represent our coupled-cavity system, in which the magenta diamond line represents the case of initial even phonon number which finally can be cooled to near 0; similarly, the green square line represents the case of initial odd phonon number which finally can be cooled to near 1. The upper two lines represent the situation of single-cavity optomechanical system [54, 55], from which we can see that the ground-state cooling can not be achieved in unresolved sideband regime. Obviously, the cooling effect in the current coupled-cavity system is significantly improved compared to that in single-cavity system due to the interference effect caused by the participation of the auxiliary cavity, which suppresses the Stokes heating processes in unresolved sideband regime and weak coupling mechanism. This will be benefit to the practical experiments in weak coupling limit even beyond the resolved sideband regime.

 figure: Fig. 8

Fig. 8 Exact numerical result of the average phonon number bb in quadratic optomechanical system for different initial thermal phonon number nth=10 (green pentagrams), nth=20 (magenta circles), nth=30 (blue triangles). The other parameters are taken as Δ1=4ωm, Δ2=2ωm, κ1=10ωm, κ2=0.01ωm, γm=106ωm, J=26ωm and G=0.08ωm. The shadow area denotes the optimal cooling region for bb<1.

Download Full Size | PDF

In Fig. 8, the cooling dynamics under different temperatures are plotted with blue triangles, magenta circles and green stars standing for different thermal phonon number nth=10, 20, 30. It can be seen from the figure that our current quadratic cooling scheme is robust to the bath noise under the appropriate temperature. We also notice in most cases, that the number of phonons is actually arbitrary in practical experiments. It is very difficult for people to really control phonons to integers. Assuming the initial thermal phonon number is nth=2k+k0 or nth=(2k+1)+k0 with k being an integer and 0<k0<1, the corresponding final average phonon number will be cooled to k0 or k0+1, respectively. So in order to enhance the cooling efficiency as much as possible, cryogenic precooling is still required to reduce the environmental temperature.

6. Experimental feasibility

The present cooling scheme can be implemented with different kinds of optomechanical cavities. For instance, a membrane-in-the middle setup which formed by a high-finesse Fabry-Perot cavity with a thin semi-transparent SiN membrane inside can be used like described in the Ref. [60]. They demonstrated that the position and orientation of the membrane in the cavity can be used to fine-tune the frequency of the cavity modes, and make these modes couple with the vibration modes of the membrane. In most membrane positions, the frequency shifts of the cavity mode is linearly related to the membrane deformation, so there is a traditional radiation pressure coupling between the optical and mechanical modes. However, at the nodes and antinodes of the cavity field, the linear term disappears, and one has a dispersive interaction, which is quadratic in the position operator of the mechanical mode.

In existing experiments with the membrane-in-the-middle geometry work in the resolved sideband limit at large thermal phonon number, the optomechanical coupling is small compared to the cavity linewidth G/κ=105 [61] for the parameters κ=105Hz, ωm=106Hz, γ=0.1Hz, and nth=107 at T=300K. Therefore, many experiments are devoted to improve the optomechanical coupling strength. Experiments with ultracold atoms in optical resonators have already realized quadratic optomechanical coupling G/κ1 at small thermal phonon number nth [62]. In single cavity optomechanical system, ground-state cooling is achieved for the coupling strength within the range of 0.5G/κ0.8 [54]. In our current system, the cavity decay rate and the optomechanical coupling strength are taken as κ1=30ωm (or 10ωm) and G=0.08ωm for numerical simulations, i.e., G/κ1103, which is achievable based on the above experimental parameters. Meanwhile, it is noticeable that the ground-state cooling can be realized in our current scheme without the request for a large quadratic coupling strength due to the participation of the second cavity.

7. Conclusions

In conclusions, we have proposed a EIT-like nonlinear ground-state cooling scheme for the mechanical resonator with weak quadratic optomechanical coupling in unresolved sideband regime. The optical force spectrum splits into a heating peak and a cooling peak from original single Lorentz peak due to the auxiliary cavity introduced. Considering the normal mode splitting, the optimum cooling condition can be obtained. Based on the destructive interference between different transitions and the optimum cooling conditions, the heating process is well suppressed, and the mechanical resonator can be cooled to near its ground state even in unresolved sideband regime. It is demonstrated by the numerical simulations that in our current scheme the mechanical oscillator can be cooled to its ground state with a faster rate than in the linear-coupling system due to the two-phonon absorption. Moreover, given the condition of quadratic optomechanical coupling, the cooling effect is greatly improved compared to the single-cavity system due to contribution of the second auxiliary cavity. By tunning the coupling between two cavities to fulfill the optimum cooling conditions, the ground-state cooling can be achieved even with a weak quadratic coupling strength, thus providesan solution for overcoming the limitations of weak quadratic coupling rate in experiments. The scheme is feasible and promising to be used in quantum manipulation of macroscopic mechanical devices beyond the resolved sideband limit and linear coupling regime.

Funding

National Natural Science Foundation of China under Grant 11564041, 61822114.

References

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

2. K. C. Schwab and M. L. Roukes, “Putting mechanics into quantum mechanics,” Phys. Today 58 (7), 36 (2005). [CrossRef]  

3. T. J. Kippenberg and K. J. Vahala, “Cavity opto-mechanics,” Opt. Express 15, 17172 (2007). [CrossRef]   [PubMed]  

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

5. F. Xue, L. Zhong, Y. Li, and C. P. Sun, “Analogue of cavity quantum electrodynamics for coupling between spin and a nanomechanical resonator: Dynamic squeezing and coherent manipulations,” Phys. Rev. B 75, 033407 (2007). [CrossRef]  

6. V. Fiore, Y. Yang, M. C. Kuzyk, R. Barbour, L. Tian, and H. Wang, “Storing an optical pulse as a mechanical excitation in a silica pptomechanical resonator,” Phys. Rev. Lett. 107, 133601 (2011). [CrossRef]  

7. Y. D. Wang and A. A. Clerk, “Using interference for high fidelity quantum state transfer in optomechanics,” Phys. Rev. Lett. 108, 153603 (2012). [CrossRef]   [PubMed]  

8. Sh. Barzanjeh, M. Abdi, G. J. Milburn, P. Tombesi, and D. Vitali, “A reversible optical to microwave quantum interface,” Phys. Rev. Lett. 109, 130503 (2012). [CrossRef]  

9. H. K. Li, X. X. Ren, Y. C. Liu, and Y. F. Xiao, “Effective photon-photon interactions in largely detuned optomechanics,” Phys. Rev. A 88, 053850 (2013). [CrossRef]  

10. Y. Yan, W. J. Gu, and G. X. Li, “Entanglement transfer from two-mode squeezed vacuum light to spatially separated mechanical oscillators via dissipative optomechanical coupling,” Sci. China Phys. Mech. Astron. 58, 050306 (2015). [CrossRef]  

11. M. D. LaHaye, O. Buu, B. Camarota, and K. C. Schwab, “Approaching the quantum limit of a nanomechanical resonator,” Science 304, 74 (2004). [CrossRef]   [PubMed]  

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

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

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

15. 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, optomechanical superpositions via nested interferometry,” Phys. Rev. Lett. 107, 020405 (2011). [CrossRef]  

16. B. Pepper, R. Ghobadi, E. Jeffrey, C. Simon, and D. Bouwmeester, “Optomechanical superpositions via nested interferometry,” Phys. Rev. Lett. 109, 023601 (2012). [CrossRef]   [PubMed]  

17. P. Sekatski, M. Aspelmeyer, and N. Sangouard, “Macroscopic optomechanics from displaced single-photon entanglement,” Phys. Rev. Lett. 112, 080502 (2014). [CrossRef]  

18. A. D. O’Connell, M. Hofheinz, M. Ansmann, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang, M. Weides, J. Wenner, J. M. Martinis, and A. N. Cleland, “Quantum ground state and single-phonon control of a mechanical resonator,” Nature (London) 464, 697 (2010). [CrossRef]  

19. S. Mancini, D. Vitali, and P. Tombesi, “Optomechanical cooling of a macroscopic oscillator by homodyne feedback,” Phys. Rev. Lett. 80, 688 (1998). [CrossRef]  

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

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

22. 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]  

23. R. W. Peterson, T. P. Purdy, N. S. Kampel, R.W. Andrews, P. L. Yu, K. W. Lehnert, and C.A. Regal, “Laser cooling of a micromechanical membrane to the quantum backaction limit,” Phys. Rev. Lett. 116, 063601 (2016). [CrossRef]   [PubMed]  

24. J. B. Clark, F. Lecocq, R. W. Simmonds, J. Aumentado, and J. D. Teufel, “Sideband cooling beyond the quantum backaction limit with squeezed light,” Nature 541, 191 (2017). [CrossRef]   [PubMed]  

25. L. Qiu, I. Shomroni, P. Seidler, and T. J. Kippenberg, “High-fidelity laser cooling to the quantum ground state of a silicon nanomechanical oscillator,” arXiv:1903.10242 [quant-ph].

26. J. Chan, T. P. M. 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 478, 89 (2011). [CrossRef]   [PubMed]  

27. F. Elste, S. M. Girvin, and A. A. Clerk, “Quantum noise interference and backaction cooling in cavity nanomechanics,” Phys. Rev. Lett. 102, 207209 (2009). [CrossRef]   [PubMed]  

28. M. Li, W. H. P. Pernice, and H. X. Tang, “Reactive cavity optical force on microdisk-coupled nanomechanical beam waveguides,” Phys. Rev. Lett. 103, 223901 (2009). [CrossRef]  

29. T. Weiss and A. Nunnenkamp, “Quantum limit of laser cooling in dispersively and dissipatively coupled optomechanical systems,” Phys. Rev. A 88, 023850 (2013). [CrossRef]  

30. M. Y. Yan, H. K. Li, Y. C. Liu, W. L. Jin, and Y. F. Xiao, “Dissipative optomechanical coupling between a single-wall carbon nanotube and a high-Q microcavity,” Phys. Rev. A 88, 023802 (2013). [CrossRef]  

31. A. Xuereb, R. Schnabel, and K. Hammerer, “Dissipative optomechanics in a Michelson-Sagnac interferometer,” Phys. Rev. Lett. 107, 213604 (2011). [CrossRef]   [PubMed]  

32. T. Ojanen and K. Børkje, “Ground state cooling of mechanical motion in the unresolved sideband regime by use of optomechanically induced transparency,” Phys. Rev. A 90, 013824 (2014). [CrossRef]  

33. Y. C. Liu, Y. F. Xiao, X. S. Luan, and C. W. Wong, “Optomechanically-induced-transparency cooling of massive mechanical resonators to the quantum ground state,” Sci. China Phys. Mech. Astron. 58, 050305 (2015). [CrossRef]  

34. C. Genes, H. Ritsch, and D. Vitali, “Micromechanical oscillator ground-state cooling via resonant intracavity optical gain or absorption,” Phys. Rev. A 80, 061803 (2009). [CrossRef]  

35. K. Hammerer, K. Stannigel, C. Genes, P. Zoller, P. Treutlein, S. Camerer, D. Hunger, and T. W. Hansch, “Optical lattices with micromechanical mirrors,” Phys. Rev. A 82, 021803 (2010). [CrossRef]  

36. S. Zhang, J. Q. Zhang, J. Zhang, C. W. Wu, W. Wu, and P. X. Chen, “Ground state cooling of an optomechanical resonator assisted by a Λ-type atom,” Opt. Express 22, 28118 (2014). [CrossRef]   [PubMed]  

37. Y. Guo, K. Li, W. Nie, and Y. Li, “Electromagnetically-induced-transparency-like ground-state cooling in a double-cavity optomechanical system,” Phys. Rev. A 90, 053841 (2014). [CrossRef]  

38. Y. C. Liu, Y. F. Xiao, X. Luan, Q. Gong, and C. W. Wong, “Coupled cavities for motional ground-state cooling and strong optomechanical coupling,” Phys. Rev. A 91, 033818 (2015). [CrossRef]  

39. W. J. Gu and G. X. Li, “Quantum interference effects on ground-state optomechanical cooling,” Phys. Rev. A 87, 025804 (2013). [CrossRef]  

40. A. Nunnenkamp, K. Børkje, and S. M. Girvin, “Cooling in the single-photon regime of optomechanics,” Phys. Rev. A 85, 051803 (2012). [CrossRef]  

41. R. L. de Matos Filho and W. Vogel, “Second-sideband laser cooling and nonclassical motion of trapped ions,” Phys. Rev. A 50, R1988 (1994). [CrossRef]   [PubMed]  

42. M. Koch, C. Sames, M. Balbach, H. Chibani, A. Kubanek, K. Murr, T. Wilk, and G. Rempe, “Three-photon correlations in a strongly driven atom-cavity system,” Phys. Rev. Lett. 107, 023601 (2011). [CrossRef]   [PubMed]  

43. X. Lü, Y. Wu, J. R. Johansson, H. Jing, J. Zhang, and F. Nori, “Squeezed optomechanics with phase-matched amplification and dissipation,” Phys. Rev. Lett. 114, 093602 (2015). [CrossRef]   [PubMed]  

44. J. D. Thompson, B. M. Zwickl, and A. M. Jayich, “Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane,” Nature (London) 452, 72 (2008). [CrossRef]  

45. A. M. Jayich, J. C. Sankey, B. M. Zwickl, C. Yang, J. D. Thompson, S. M. Girvin, A. A. Clerk, F. Marquardt, and J. G. E. Harris, “Dispersive optomechanics: a membrane inside a cavity,” New J. Phys. 10, 095008 (2008). [CrossRef]  

46. K. W. Murch, K. L. Moore, and S. Gupta, “Observation of quantum-measurement backaction with an ultracold atomic gas,” Nat. Phys. 4, 561 (2008). [CrossRef]  

47. T. P. Purdy, D. W. C. Brooks, T. Botter, N. Brahms, Z. Y. Ma, and D. M. Stamper-Kurn, “Tunable cavity optomechanics with ultracold atoms,” Phys. Rev. Lett. 105, 133602 (2010). [CrossRef]  

48. F. Helmer, M. Mariantoni, E. Solano, and F. Marquardt, “Quantum nondemolition photon detection in circuit QED and the quantum Zeno effect,” Phys. Rev. A 79, 052115 (2009). [CrossRef]  

49. H. Miao, S. Danilishin, T. Corbitt, and Y. Chen, “Standard quantum Limit for probing mechanical energy quantization,” Phys. Rev. Lett. 103, 100402 (2009). [CrossRef]   [PubMed]  

50. A. A. Clerk, F. Marquardt, and J. G. E. Harris, “Optomechanical cavity cooling of an atomic ensemble,” Phys. Rev. Lett. 104, 213603 (2010). [CrossRef]  

51. M. Asjad, G. S. Agarwal, M. S. Kim, P. Tombesi, G. D. Giuseppe, and D. Vitali, “Robust stationary mechanical squeezing in a kicked quadratic optomechanical system,” Phys. Rev. A 89, 023849 (2014). [CrossRef]  

52. S. Huang and G. S. Agarwal, “Electromagnetically induced transparency from two phonon processes in quadratically coupled membranes,” Phys. Rev. A 83, 023823 (2011). [CrossRef]  

53. T. K. Paraïso, M. Kalaee, L. Zang, H. Pfeifer, F. Marquardt, and O. Painter, “Position-squared coupling in a tunable photonic crystal optomechanical cavity,” Phys. Rev. X 5, 041024 (2015).

54. A. Nunnenkamp, K. Børkje, J. G. E. Harris, and S. M. Girvin, “Cooling and squeezing via quadratic optomechanical coupling,” Phys. Rev. A 82, 021806 (2010). [CrossRef]  

55. W. J. Gu, Z. Yi, L. H. Sun, and D. H. Xu, “Mechanical cooling in single-photon optomechanics with quadratic nonlinearity,” Phys. Rev. A 92, 023811 (2015). [CrossRef]  

56. J. S. Feng, L. Tan, H. Q. Gu, and W. M. Liu, “Auxiliary-cavity-assisted ground-state cooling of an optically levitated nanosphere in the unresolved-sideband regime,” Phys. Rev. A 96, 063818 (2017). [CrossRef]  

57. A. A. Clerk, S. M. Girvin, and F. Marquardt, “Introduction to quantum noise, measurement, and amplification,” Rev. Mod. Phys. 82, 1155 (2010). [CrossRef]  

58. L. He, Y. X. Liu, S. Yi, C. P. Sun, and F. Nori, “Control of photon propagation via electromagnetically induced transparency in lossless media,” Phys. Rev. A 75, 063818 (2007). [CrossRef]  

59. V. V. Dodonov and S. S. Mizrahi, “Exact stationary photon distributions due to competition between one- and two-photon absorption and emission,” J. Phys. A 30, 5657 (1997). [CrossRef]  

60. M. Karuza, M. Galassi, C. Biancofiore, C. Molinelli, R. Natali, P. Tombesi, G. Di. Giuseppe, and D. Vitali, “Tunable linear and quadratic optomechanical coupling for a tilted membrane within an optical cavity: theory and experiment,” Physics 15, 205 (2011).

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

62. J. C. Sankey, C. Yang, B. M. Zwickl, A. M. Jayich, and J. G. E. Harris, “Strong and tunable nonlinear optomechanical coupling in a low-loss system,” Nat. Phys. 6, 707 (2010). [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 (8)

Fig. 1
Fig. 1 Schematic of the optomechanical system. A membrane oscillator is placed in the middle of cavity 1 which is driven by a continuous-wave input laser, the coupling between them is proportional to the square of position of the oscillator. The cavity2 as an auxiliary cavity only directly couples to the optomechanical cavity 1 with coupling strength J.
Fig. 2
Fig. 2 Energy level diagram of the system in the displaced frame. | n 1 , n 2 , m donates the state of n1 photons in mode a1, n2 photons in mode a2, and m phonons in mode b. The orange double arrow denotes the couplingbetween two cavities with coupling strength J.
Fig. 3
Fig. 3 (a) The optical force spectrum SFF (ω) (in arbitrary units) for different optical coupling strengths J, with Δ1 = 2ωm, κ2 = 0.01ωm. (b) The optical force spectrum SFF (ω) (in arbitrary units) for different detunings of the second cavity κ2 in the optimum conditions with Δ1 = −4ωm, J = 2 6 ω m. The other parameters are taken as Δ2 = −2ωm, κ1 = 10ωm, G = 0.08ωm.
Fig. 4
Fig. 4 The net cooling rate Γopt = AA+ versus the dacay rate of the second cavity κ2 and the coupling strength J. Here the detuning and the decay rate of the first cavity are taken as Δ1 = −4ωm, κ1 = 10ωm, the detuning of the second cavity is taken as Δ2 = −2ωm, and the optomechanical coupling strength is taken as G = 0.08ωm.
Fig. 5
Fig. 5 A comparison of the net cooling rates Γ opt = A A + between the current coupled-cavity optomechanical system and the single-cavity optomechanical system. The former is plotted as the green circle curve under the optimum coupling conditions in Eq. (12) versus the detuning of the auxiliary cavity Δ2 for Δ 1 = 4 ω m, κ 1 = 30 ω m, κ 2 = 0.01 ω m, and the latter is plotted as the blue triangle curve versus Δ1 for κ = 30 ω m. The effective optomechanical coupled strength G = 0.08 ω m is taken.
Fig. 6
Fig. 6 (a) The average phonon number b b of the steady state versus the detuning of the second cavity Δ2 and the optomechanical coupling strength G. The detuning of first cavity Δ1 and the coupling strength between two cavities J meet the optimum condition in Eq. (12). (b) The average phonon number b b versus the effective detuning of the first cavity Δ1 and the coupling strength J. Here the effective detuning of the second cavity Δ 2 = 2 ω m is taken. The white dotted curve satisfies the Eq. (12). The other parameters are taken as κ 1 = 10 ω m, κ 2 = 0.01 ω m, γ m = 10 6 ω m, G = 0.08 ω m and n th = 10.
Fig. 7
Fig. 7 A comparison of the cooling dynamics between the current coupled-cavity optomechanical system and the single-cavity optomechanical system. The numerical results of the average phonon number b b for the coupled-cavity system are plotted with the magenta diamond and green square curves, corresponding to the initial even and odd phonon numbers n th = 10 and 11, respectively, with Δ 1 = 4 ω m, Δ 2 = 2 ω m, J = 2 6 ω m. The numerical results for single-cavity optomechanical system are plotted with orange circle and blue triangle curves corresponding to initial even and odd phonon numbers n th = 10 and 11, respectively, with Δ 1 = 2 ω m. The shadow area denotes the optimal cooling region for b b < 1.
Fig. 8
Fig. 8 Exact numerical result of the average phonon number b b in quadratic optomechanical system for different initial thermal phonon number n th = 10 (green pentagrams), n th = 20 (magenta circles), n th = 30 (blue triangles). The other parameters are taken as Δ 1 = 4 ω m, Δ 2 = 2 ω m, κ 1 = 10 ω m, κ 2 = 0.01 ω m, γ m = 10 6 ω m, J = 2 6 ω m and G = 0.08 ω m. The shadow area denotes the optimal cooling region for b b < 1.

Equations (15)

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

H = ω 1 a 1 a 1 + ω 2 a 2 a 2 + ω m b b + g a 1 a 1 ( b + b ) 2 + Ω ( a 1 e i ω L t + a 1 e i ω L t ) + J ( a 1 a 2 + a 1 a 2 ) ,
H = Δ 1 a 1 a 1 + Δ 2 a 2 a 2 + ω m b b + g a 1 a 1 ( b + b ) 2 + Ω ( a 1 + a 1 ) + J ( a 1 a 2 + a 1 a 2 ) ,
α 1 = i Ω ( i Δ 1 κ 1 2 ) + J 2 i Δ 2 κ 2 2 , α 2 = i J α 1 i Δ 2 κ 2 2 , β = 0.
H eff = Δ 1 a 1 a 1 + Δ 2 a 2 a 2 + ω m b b + ( G a 1 + G * a 1 ) ( b + b ) 2 + J ( a 1 a 2 + a 1 a 2 ) ,
H op = Δ 1 a 1 a 1 + Δ 2 a 2 a 2 + J ( a 1 a 2 + a 1 a 2 ) ,
i ω a 1 ( ω ) = ( i Δ 1 κ 1 2 ) a 1 ( ω ) i J a 2 ( ω ) κ 1 a in , 1 ( ω ) , i ω a 2 ( ω ) = ( i Δ 2 κ 2 2 ) a 2 ( ω ) i J a 1 ( ω ) κ 2 a in , 2 ( ω ) ,
a 1 ( ω ) = κ 1 a in , 1 ( ω ) i J χ 2 ( ω ) κ 2 a in , 2 ( ω ) χ ( ω ) ,
χ 1 ( ω ) = 1 κ 1 / 2 i ( ω Δ 1 ) , χ 2 ( ω ) = 1 κ 2 / 2 i ( ω Δ 2 ) , χ ( ω ) = 1 χ 1 ( ω ) + J 2 χ 2 ( ω ) .
S FF ( ω ) = | G | 2 x ZPF 4 [ 1 χ ( ω ) + 1 χ * ( ω ) ] .
P n ˙ = γ m [ n th ( n + 1 ) + ( n th + 1 ) n ] P n + γ m n th n P n 1 + γ m ( n th + 1 ) ( n + 1 ) P n + 1 [ A n ( n 1 ) + A + ( n + 2 ) ( n + 1 ) ] P n + A ( n + 2 ) ( n + 1 ) P n + 2 + A + n ( n 1 ) P n 2 ,
E + = 1 2 ( Δ 1 + Δ 2 + ( Δ 1 + Δ 2 ) 2 + 4 J 2 ) , E = 1 2 ( Δ 1 + Δ 2 ( Δ 1 + Δ 2 ) 2 + 4 J 2 ) .
J = 4 ω m ( 2 ω m Δ 1 ) .
P 2 n + j ( 2 ) = ( 1 m ) m n ( γ + j 1 ) ( 1 ) j 1 ,       j = 0 , 1 ,
P 0 = 1 + 2 ξ 1 + 3 ξ + O ( γ m n t h A ) , P 1 = ξ 1 + 3 ξ + O ( γ m n t h A ) ,   ξ = n t h n t h + 1 ,
ρ ˙ = i [ H eff , ρ ] + κ 1 2 ( 2 a 1 ρ a 1 a 1 a 1 ρ ρ a 1 a 1 ) + κ 2 2 ( 2 a 2 ρ a 2 a 2 a 2 ρ ρ a 2 a 2 ) + γ m 2 ( n th + 1 ) ( 2 b ρ b b b ρ ρ b b ) + γ m 2 n th ( 2 b ρ b b b ρ ρ b b ) ,
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.