Steady-state entanglement, cooling, and tristability in ... - OSA Publishing

32 downloads 0 Views 2MB Size Report
2Quantum Optics Group, Department of Physics, Faculty of Science, University of Isfahan, Hezar Jerib, 81746-73441. Isfahan, Iran. 3School of Science and ...
Shahidani et al.

Vol. 31, No. 5 / May 2014 / J. Opt. Soc. Am. B

1087

Steady-state entanglement, cooling, and tristability in a nonlinear optomechanical cavity S. Shahidani,1,* M. H. Naderi,1,2 M. Soltanolkotabi,1,2 and S. Barzanjeh3 2

1 Department of Physics, Faculty of Science, University of Isfahan, Hezar Jerib, 81746-73441 Isfahan, Iran Quantum Optics Group, Department of Physics, Faculty of Science, University of Isfahan, Hezar Jerib, 81746-73441 Isfahan, Iran 3 School of Science and Technology, Physics Division, Universita di Camerino, I-62032 Camerino (MC), Italy *Corresponding author:[email protected]

Received January 8, 2014; revised March 17, 2014; accepted March 18, 2014; posted March 19, 2014 (Doc. ID 204378); published April 16, 2014 The interaction of a single-mode field with both a weak Kerr medium and a parametric nonlinearity in an intrinsically nonlinear optomechanical system is studied. The nonlinearities due to the optomechanical coupling and Kerr-down conversion lead to bistability and tristability in the mean intracavity photon number. Also, our work demonstrates that the lower bound of the resolved sideband regime and the minimum attainable phonon number can be less than those of a bare cavity by controlling the parametric nonlinearity and the phase of the driving field. Moreover, we find that in the system under consideration the degree of entanglement between the mechanical and optical modes is dependent on the two stability parameters of the system. For both cooling and entanglement, while parametric nonlinearity increases the optomechanical coupling, the weak Kerr nonlinearity is very useful for extending the domain of the stability region to the desired range in which the minimum effective temperature and maximal entanglement are attainable. Also, as shown in this paper, the present scheme allows us to have significant entanglement in the tristable regime for the lower and middle branches, which makes the current scheme distinct from the bare optomechanical system. © 2014 Optical Society of America OCIS codes: (270.0270) Quantum optics; (200.4880) Optomechanics; (190.0190) Nonlinear optics. http://dx.doi.org/10.1364/JOSAB.31.001087

1. INTRODUCTION Considerable interest has recently been focused on the optomechanical system as an excellent candidate for studying the transition of a macroscopic degree of freedom from the classical to the quantum regime. This system also provides novel routes for practical applications, such as detection and interferometry of gravitational waves [1] and quantum limited displacement sensing [2]. The standard and simplest setup of this system is a Fabry–Perot cavity in which one of the mirrors is much lighter than the other, so that it can move under the effect of the radiation pressure force. State-of-the-art technology allows experimental demonstration of cooling of the vibrational mode of the mechanical oscillator to its ground state [3–5] and strong optomechanical coupling between the vibrational mode of the mechanical oscillator and the cavity field [6–9]. This coupling is intrinsically nonlinear since the length of the cavity depends upon the intensity of the field in an analogous way to the optical length of a Kerr material [10]. Therefore it enables pondermotive squeezing of the field [11,12], photon blockade [13], generation of nonclassical states of the mechanical and optical mode [14], optical bistability [15,16], and phonon–photon entanglement in the bistable regime [17]. Besides this intrinsic nonlinearity, the presence of an optical parametric amplifier (OPA) [18,19] or the optical Kerr medium [20] inside the cavity has opened up a new domain for combining nonlinear optics and optomechanics toward the enhancement of quantum effects. It has been predicted [18] that the presence of an OPA in a single-mode Fabry–Perot cavity causes a strong coupling 0740-3224/14/051087-09$15.00/0

between the oscillating mirror and the cavity mode resulting from increasing the intracavity photon number. Also, when the optomechanical cavity contains an optical Kerr medium with strong χ 3 nonlinearity, the photon–photon repulsion and the reduction of the cavity photon fluctuations provide a feasible route toward controlling the dynamics of the micromirror [20]. On the other hand, the interaction of a single-mode field with both a Kerr medium and a parametric nonlinearity is a well-known quantum optical model that has been proposed for the generation of nonclassical states of the cavity field [21,22]. Here, we consider the interaction of a single-mode field with both a weak Kerr medium and a parametric nonlinearity in an intrinsically nonlinear optomechanical system. In particular, we investigate the multistability, intensity, back-action cooling, and stationary optomechanical entanglement. There are different methods for the experimental implementation of a Kerr-down conversion nonlinear medium; it could be a twolayer nonlinear medium comprising Kerr and down conversion nonlinearities. Optical polymeric materials because of their processability, transparency, and large optical nonlinearity could be excellent candidates for this combination [23,24]. Also, recently, it has been experimentally verified [25] that the features of whispering gallery resonators enable both intrinsic second- and third-harmonic generation. Hence one of the most natural implementations of the model under consideration could be a whispering gallery resonator made out of an anisotropic material with χ 2 and the Kerr nonlinearities. © 2014 Optical Society of America

1088

J. Opt. Soc. Am. B / Vol. 31, No. 5 / May 2014

We show that in this type of hybrid optomechanical structure the mean intracavity photon number, in addition to bistability, exhibits tristability for a certain range of the parameters, which can be controlled by the intrinsic and external nonlinearities in the system. Then, we investigate the effect of Kerr-down conversion nonlinearity on the backaction ground-state cooling of the mirror based on the covariance matrix (CM) and identify the modified optimal regime for cooling. We will show that the lower bound of the resolved sideband regime and the minimum attainable phonon number can be less than those of a bare cavity by controlling the parametric nonlinearity and the phase of the field driving the OPA. Also, the weak Kerr nonlinearity can be used to extend the domain of the stability to the desired range of the effective detuning in which the effective temperature of the system minimizes. Then we show that for a fixed effective detuning, in spite of the bare cavity, one of the stability parameters (the counterpart of the bistability parameter) is a nonlinear function of the input power, which allows us to approach significant entanglement simultaneously with the ground-state cooling of the mirror. In the last part of the paper we shall focus on the generation of stationary entanglement in the presence of nonlinearity and show that not only is the entanglement not a monotonic function of the optomechanical coupling strength, but the two stability parameters of the system are also the key parameters for maximizing the degree of entanglement. Based on these results we show that in the tristable regime, for the first and second branches the degree of entanglement is maximum at the end of the branches, while for the third branch the phonon–photon entanglement is null. In brief, while in other relevant studies (i.e., Refs. [18–20]) the intrinsic optomechanical nonlinearity is ignored, considering self-action OPA (instead of cross-action OPA as investigated in Ref. [19]) and weak Kerr nonlinearity (instead of strong Kerr nonlinearity as considered in Ref. [20]) allows us to achieve strong optomechanical coupling and extend the domain of the stability, which makes the current scheme more useful for increasing the quantum effects (specifically the entanglement) in the optomechanical system.

2. PHYSICAL MODEL As shown in Fig. 1, we consider a Kerr-down conversion optomechanical system composed of a degenerate OPA and a nonlinear Kerr medium placed within a Fabry–Perot cavity formed by a fixed partially transmitting mirror and one movable perfectly reflecting mirror in equilibrium with a thermal bath at temperature T 0 . The movable mirror is free to move along the cavity axis and is treated as a quantum mechanical harmonic oscillator with effective mass m, frequency ωm , and energy decay rate γ m  ωm ∕Q, where Q is the mechanical quality factor. The cavity field is coherently driven by an input laser field with frequency ωL and amplitude ε through the fixed mirror. Furthermore, the system is pumped by a coupling field to produce parametric oscillation and induce the Kerr nonlinearity in the cavity. In our investigation, we restrict the model to the case of single-cavity and mechanical modes [26–28]. The single-cavity-mode assumption is justified in the adiabatic limit, i.e., ωm ≪ πc∕L, in which c is the speed of light in vacuum and L is the cavity length in the absence of the cavity field. We also assume that the induced resonance frequency shift of the cavity and the Kerr medium are much

Shahidani et al.

Fig. 1. Schematic picture of the setup studied in the text. The cavity contains a Kerr-down conversion system that is pumped by a coupling field to produce parametric oscillation and induce Kerr nonlinearity in the cavity.

smaller than the longitudinal-mode spacing in the cavity. Furthermore, one can restrict to a single mechanical mode when the detection bandwidth is chosen such that it includes only a single, isolated, mechanical resonance and mode–mode coupling is negligible. Under these conditions, the total Hamiltonian of the system in a frame rotating at the laser frequency ωL can be written as H  H0  H 1;

(1)

where H 0  ℏω0 − ωL a† a 

ℏωm 2 q  p2  − ℏg0 a† aq 2

 iℏεa† − a; H 1  iℏGeiθ a†2 − e−iθ a2   ℏχa†2 a2 :

(2a)

(2b)

The first two terms in H 0 are, respectively, the free Hamiltonians of the cavity field with annihilation (creation) operator aa†  and the movable mirror with resonance frequency ωm and dimensionless position and momentum operators q and p. The third term describes the optomechanical coupling between the cavity field and the mechanical oscillator due to the radiation pressure force with coupling constant p g0  ω0 ∕L h∕mωm , and the last term in H 0 describes the driving of the intracavity mode with the input laser. Also, the two terms in H 1 describe, respectively, the coupling of the intracavity field with the OPA and the Kerr medium; G is the nonlinear gain of the OPA, which is proportional to the pump power driving amplitude, θ is the phase of the field driving the OPA, and χ is the anharmonicity parameter proportional to the third-order nonlinear susceptibility χ 3 of the Kerr medium. The input laser field populates the intracavity mode through the partially transmitting mirror; then the photons in the cavity will exert a radiation pressure force on the movable mirror. In a realistic treatment of the dynamics of the system, the cavity-field damping due to the photon leakage through the incomplete mirror and the Brownian noise associated with the coupling of the oscillating mirror to its thermal

Shahidani et al.

Vol. 31, No. 5 / May 2014 / J. Opt. Soc. Am. B

bath should be considered. Using the input–output formalism of quantum optics [29], we can consider the effects of these sources of noise and dissipation in the quantum Langevin equations of motion. For the given Hamiltonian (1), we obtain the following nonlinear equations of motion: q_  ωm p;

(3a)

p_  −ωm q  g0 a† a − γ m p  ξ;

(3b)

a_  −iω0 − ωL a  ig0 qa  ε − 2iχa† a2 p 2Ga† eiθ − κa  2κ ain ;

(3c)

where κ is the cavity decay rate through the input mirror and ain is the input vacuum noise operator characterized by the following correlation functions [29]: hain ta†in t0 i  δt − t0 ;

(4a)

hain tain t0 i  ha†in ta†in t0 i  0:

(4b)

1089

The multistability of the p solution fails if χ  g20 ∕2ωm and  requires j2G sin θ − Δ0 j > 3κ− . To study the effect of the presence of Kerr-down conversion nonlinearity on the steady-state response of the optomechanical system, we consider a cavity with length L  1 mm and decay rate κ  0.9ωm , which is driven by a laser with λ  810 nm. The mass, the mechanical frequency, and the damping rate of the oscillating end mirror are m  5 ng, ωm ∕2π  10 MHz, and γ m  100 Hz, and the environment temperature is T 0  400 mK. This set of parameters is close to several optomechanical experiments [7,32–34]. In Fig. 2 we plot the mean intracavity photon number as a function of the bare detuning Δ0 for input power P  15 mW for various values of χ [Fig. 2(a)], G [Fig. 2(b)], and θ [Fig. 2(c)]. Figure 2(a) shows that for χ < g20 ∕2ωm (χ  0.01) the presence of both

The Brownian noise operator ξ describes the heating of the mirror by the thermal bath at temperature T 0 and is characterized by the following correlation function [29–31]: hξtξt0 i 

γm πωm

Z

    ℏω 0 ωe−iωt−t  coth  1 dω; 2kB T 0

(5)

where kB is the Boltzmann constant. We are interested in the steady-state regime and small fluctuations with respect to the steady state. Thus we obtain the steady-state mean values of p, q, and a as ps  0;

as 

qs 

g0 2 a ; ωm s

κ − iΔ  2Geiθ ε; Δ2  κ2 − 4G2

(6)

(7)

where qs denotes the new equilibrium position of the movable mirror, κ−  κ − 2G cos θ, and Δ  ω0 − ωL − g0 qs  2χjas j2  Δ0  2χ − g20 ∕ωm jas j2 is the effective detuning of the cavity, which includes both the radiation pressure and the Kerr medium effects. We choose the phase of ε so that as can be taken real and positive. It is obvious that the optical path and hence the cavity detuning are modified in an intensitydependent way. The first modification, which is a mechanical nonlinearity, arises from the radiation pressure-induced coupling between the movable mirror and the cavity field, and the second modification comes from the presence of the nonlinear Kerr medium in the optomechanical system. Since the mean intracavity photon number in the steady state I a  a2s  satisfies a third-order equation, it can have three real solutions and hence the system may exhibit multistability for a certain range of parameters. The multisolution region exists for intracavity intensity values between I − and I  with p 4G sin θ − 2Δ0   2G sin θ − Δ0 2 − 3κ 2− : I  32χ − g20 ∕ωm 

(8)

Fig. 2. Mean intracavity photon number as a function of normalized bare detuning Δ0 ∕ωm : (a) for different values of the anharmonicity parameter χ with G  0.6κ and θ  π∕2, (b) for different values of the parametric nonlinearity G with χ  0.1 Hz and θ  π∕2, and (c) for different values of θ with G  1.1κ and χ  0.04 Hz.

J. Opt. Soc. Am. B / Vol. 31, No. 5 / May 2014

Shahidani et al.

intrinsically optomechanical and Kerr nonlinearities shifts the center of the resonance while the curve is nearly Lorentzian. However, for χ > g20 ∕2ωm , the resonance frequency of the cavity shifts to the lower values and the third-order polynomial equation for I a has three real roots for Δ0 < 0. The frequency shift of the cavity mode and multi solutions due to the Kerr nonlinearity can be reduced or compensated by the term 2G sin θ, which acts to shift the cavity resonant frequency to the right for θ ≥ π∕2 [Figs. 2(b) and 2(c)].

3. DYNAMICS OF SMALL FLUCTUATIONS In order to investigate the dynamics of the system, we need to study the dynamics of small fluctuations near the steady state. We assume that the nonlinearity in the system is weak and decompose each operator in Eq. (3) as the sum of its steady-state value and a small fluctuation with zero mean value, a  as  δa;

q  qs  δq;

p  ps  δp:

(9)

Inserting the above linearized forms of the system operators into Eq. (3), pand defining the cavity-field p quadratures δx  δa  δa† ∕ 2 and δy  iδa† − p δa∕  2 and the input noise quadratures δxin  δain  δa†in ∕ 2 and δyin  iδa†in − δain ∕ p 2, the linearized quantum Langevin equations for the fluctuation operators can be written in the compact matrix form u_  Mut  nt;

(10)

T where ut  δq; δp;p δx; p δy Tis the vector of fluctuations, nt  0; ξ; 2κδxin ; 2κ δyin  is the vector of noise sources, and the matrix M is given by

0

0 B −ωm M B @ 0 g1

ωm −γ m 0 0

0 g1 −κ  Γr −Δ1  Γi

1 0 C 0 C; Δ1  Γ i A −κ − Γr

(11)

p where g1  2g0 as is the enhanced optomechanical coupling rate, Γr (Γi ) is the real (imaginary) part of Γ  2Geiθ − 2iχa2s , and Δ1  Δ  2χa2s . A. Stability Analysis of the Solutions Here, we concentrate on the stationary properties of the system. For this purpose, we should consider the steady-state condition governed by Eq. (10). The system is stable only if the real part of all eigenvalues of the matrix M is negative, which is also the requirement of the validity of the linearized method. The parameter region in which the system is stable can be obtained from the Routh–Hurwitz criterion [35,36], which gives the following three independent conditions:

The violation of the third condition, s3 < 0, indicates instability in the region Δ1  Γi < 0. For the bare cavity (G  χ  0) this condition reduces to instability in the domain of the bluedetuned laser. Within this frequency range, the effective mechanical damping rate becomes negative and self-sustained oscillations set in [37,38]. The violation of the second condition, s2 < 0, indicates instability in the region Δ1  Γi > 0. For a bare cavity this condition causes a bistability of the system [15]. There is an additional stability condition given by s1 , which is always satisfied for the bare cavity, and gives the condition for the threshold for parametric oscillation. Accordingly, for Δ1  Γi > 0, we can define the following stability parameters: η1  1 −

g21 Δ1  Γi  ; ωm κ 2  Δ21 − jΓj2 

η2  1 −

(13)

jΓj2 : κ  Δ21

(14)

2

For G  χ  0, η1 reduces to the well-known “bistability parameter” [39]. One of the main features arising from Kerr-down conversion nonlinearity is the appearance of three stable states for the mirror. Figure 3 shows the hystersis loop for the intracavity mean photon number when Δ0 < 0 (blue-detuned laser). In this figure the unstable solutions are represented by dashed lines. It shows that depending on the value of θ the steady-state response of the mirror can be monostable, bistable, and tristable. From an experimental point of view, controllable triple-state switching is possible practically by adding a pulse sequence onto the input field [40]. Such tristability can be used for all-optical switching purposes functioning as memory devices for optical computing and quantum information processing. B. Correlation Matrix of the Quantum Fluctuations of the System Due to the linearization method and the Gaussian nature of the noise operators the asymptotic steady state of the quantum fluctuations is a zero mean Gaussian state. As a consequence, the steady state can be fully characterized by the

Π 6

4

0

3

Ia 109

1090

0.57 Π 2

1

s1  κ2  Δ21 − jΓj2 > 0;

(12a) 0

2

s2  ωm κ 

Δ21

2

− jΓj  −

g20 Δ1

 Γi  > 0;

(12b)

s3  2κγ m fs1 − ω2m 2  γ m  2κγ m s1  2κω2m g  g20 Δ1  Γi ωm 2κ  γ m 2 > 0:

(12c)

0

20

40

60

80

100

120

P mW

Fig. 3. Mean photon number I a versus the input power P at Δ0  −2.5ωm . The solid and dashed lines correspond to the stable and unstable branches, respectively. The parameters are G  κ, χ  0.05 Hz, and κ  0.9ωm . Other parameters are the same as those in Fig. 2.

Shahidani et al.

Vol. 31, No. 5 / May 2014 / J. Opt. Soc. Am. B

MV  V M T  −D;

0.40

0.30 G  0.6 Κ

(15)

0.25

where the elements of the correlation matrix V are defined as V ij  hui uj  uj ui i∕2, and D  Diag0; γ m 2n¯  1; κ; κ is the diagonal diffusion matrix, in which we used the following approximation:

0.20

 ω coth

 ℏω 2k T ≃ ωm B 0 ≃ ωm 2n¯  1; 2kB T 0 ℏωm

(16)

where n¯  eℏωm ∕kB T 0 − 1−1 . The Lyapunov Eq. (15) for the steady-state CM can be straightforwardly solved. However, the general exact expression is too cumbersome and will not be reported here.

4. EFFECTS OF KERR-DOWN CONVERSION NONLINEARITY ON GROUND-STATE COOLING AND STATIONARY ENTANGLEMENT It has been shown [42,44–48] that within each optomechanical cavity the radiation pressure coupling can lead the system into a stationary state with genuine quantum features, including cooling of the resonator toward the ground state and photon–phonon entanglement. It is therefore interesting to investigate these quantum features in the presence of Kerr-down conversion nonlinearity. The CM, V , contains all the information about the steady state. In particular, the mean energy of the mechanical oscillator is given by   ℏωm 1 2 2 U hδqi  hδpi   ℏωm neff  ; 2 2

(17)

where neff is the mean effective excitation number of the mechanical mode corresponding to an effective mode temperature T  ℏωm ∕kB ln1  1∕neff . The ground-state cooling is approached if neff ≃ 0, or U ≃ ℏωm ∕2. Moreover, the photon–phonon entanglement can be quantified by the logarithmic negativity E N [49] E N  max0; − ln 2η− ; (18) p 1∕2 where η− ≡ 2−1∕2 ΣV  − ΣV 2 − 4 det V  is the lowest symplectic eigenvalue of the partial transpose of the CM with ΣV   det V A  det V B − 2 det V C , and we have used the 2 × 2 block form of the CM 

VA V TC

 VC ; VB

(19)

with V A associated to the oscillating mirror, V B to the cavity mode, and V C describing the optomechanical correlations. We now focus on numerical examples to see how the Kerrdown conversion nonlinearity affects the ground-state cooling of the movable mirror. Figure 4 shows the variation of the

G 0

0.35

TmK

CM, V . This formalism provides a unified framework for exploring both cooling of the mechanical oscillator and phonon– photon entanglement. When the stability conditions of Eq. (12) are fulfilled, we can solve Eq. (10) for the 4 × 4 stationary correlation matrix V [41–43]

1091

G  0.8 Κ

0.15 0.2

0.4

0.6

0.8

Ωm

1.0

1.2

1.4

1.6

Fig. 4. Plot of the effective temperature T versus Δ∕ωm for different values of G. The red dashed curve corresponds to the bare cavity. The parameters are : χ  0.05 Hz, θ  0.81π for G  0.6κ and θ  0.71π for G  0.8κ, T 0  400 mK, P  5 mW, and κ  0.3ωm . Other parameters are the same as those in Fig. 3.

effective temperature T with Δ∕ωm and G for the red-detuned laser (Δ0 > 0) and in the monostable regime. In this figure we have assumed the good-cavity limit, κ  0.3ωm , and the reservoir temperature to be T 0  0.4K. The effective mode temperature of the bare cavity (red dashed line) is plotted for comparison. It shows that the parametric process inside the cavity leads to lower temperature. To illustrate this explicitly, we can solve Eq. (15) in the limit of a large mechanical quality factor and low-temperature environment (i.e., ωm ≫ ¯ m ) and obtain the position (V 11 ) and momentum γ m and κ ≫ nγ (V 22 ) variances and neff by using neff  12 V 11  V 22 − 1. The results are as follows: V 11 

η1 ω2m  Δ1  Γi 2  κ  Γr 2 ; 4η1 ωm Δ1  Γi 

(20)

ω2m  Δ1  Γi 2  κ  Γr 2 ; 4ωm Δ1  Γi 

(21)

Δ1  Γi − ωm 2  κ 2 ; 4ωm Δ1  Γi 

(22)

V 22  and for η1 ≃ 1,

neff 

where κ   κ  Γr . For a bare cavity neff reduces to Eq. (6) in [45]. The minimum attainable phonon number can be q achieved for Δ  −2G sin θ  ω2m  κ2 ,

nmin

0q 1 2 2  κ ω m  1B C  @ − 1A; ωm 2

(23)

in agreement with Eq. (7) in [45] for G  0. As can be seen, when π∕2 < θ < 3π∕2 the lower bound of the resolved sideband regime (ωm ≫ κ ) and nmin ≃κ2 ∕4ωm  can be less than that of a bare cavity. For the data used in Fig. 4, and for G  0.6κ, θ  0.81π, we have κ  ≃ 10−6 κ, and the minimum effective temperature is about 0.14 mK at Δ ≃ 0.8ωm . Also, for G  0.8κ and θ  0.71π we have κ  ≃ 8 × 10−6 κ, and the minimum effective temperature is about 0.16 mK at Δ ≃ 0.6ωm .

1092

J. Opt. Soc. Am. B / Vol. 31, No. 5 / May 2014

According to Eq. (23) it seems that nmin does not depend on the anharmonicity parameter χ, but it should be noted that even if the effect of Kerr nonlinearity on η1 can be eliminated, the above analysis is carried out for the limit of low¯ m ). In a realistic case the temperature environment (κ ≫ nγ thermal noise will also be present, which in turn modifies the minimum attainable temperature. Figures 5(a) and 5(b) show the variation of effective temperature with Δ for different values of χ and for two different environment temperatures, T 0  400 and 25 mK. For G  0.8κ and θ  3π∕4 the minimum temperature is achieved for the optimal detuning Δ ≃ 0.66ωm . As can be seen, the system is unstable in this range for χ  0 and χ  0.1. For χ  0.03 and χ  0.06 the stability domain is extended to the desired range, and the effective temperature is increased with increasing χ. As shown in Fig. 5(b) this heating effect is smaller for lower environment temperature T 0  25 mK. Also, in Fig. 5 the dependence of effective temperature on the Kerr nonlinearity has been illustrated for fixed values of stability parameter η1 . It should be emphasized, however, that the controllable ground-state cooling of the vibrational mode is possible only if the limit η1 ≃ 1 is reachable in the presence of the nonlinear medium. Figure 6 shows the variation of the parameter η1 versus the input power P and the normalized effective detuning Δ∕ωm for the data of Fig. 5(a) and for χ  0.03. As can be seen, 0.8 < η1 < 1, and the required condition holds actually very well. Also, Fig. 6 shows that for a fixed effective detuning, in contrast to the case of bare cavity

Shahidani et al.

1.00 0.95 Η1 0.90 0.85 0.80

4

0.5

2

Ωm

PmW

1.0

1.5

0

Fig. 6. Plot of the stability parameter η1 versus Δ∕ωm and input power P for χ  0.03. Other parameters are the same as those in Fig. 5(a).

[17], η1 is not a linear function of the input power P. This feature arises from the Kerr nonlinearity [the term 2χa2s in Δ1 and Γ in Eq. (13)] and allows us to approach significant entanglement simultaneously with the ground-state cooling of the mirror (since one can enhance optomechanical coupling by increasing the input power while η1 remains approximately unaffected). As an example, Fig. 7 represents the optomechanical entanglement and the effective temperature for G  1.3κ, χ  0.05, and Δ  0.5ωm as functions of the input power. It shows that the minimum value of the effective

0.26

a 0.25

T mK

0.24 0.23 0.22 0.21 0.20 0.19 2

4

6

8

10

12

P mW 0.22

b 0.20

EN

0.18

0.16

0.14

Fig. 5. Plot of the effective temperature T versus Δ for environment temperature (a) T 0  400 mK and (b) T 0  25 mK for different values of χ. The red dashed curve corresponds to the cavity containing only the gain nonlinearity. The parameters are G  0.8κ, θ  3π∕4, and P  5 mW. Other parameters are the same as those in Fig. 3, and the stability parameter is fixed to be η1  0.99.

0.12 2

4

6

8

10

12

P mW

Fig. 7. Plot of (a) effective temperature T and (b) logarithmic negativity versus the input power P for Δ  0.5ωm , G  1.3κ, θ  0.67π, and χ  0.05. Other parameters are the same as those in Fig. 4.

Shahidani et al.

Vol. 31, No. 5 / May 2014 / J. Opt. Soc. Am. B

0.25

6

0.09

0.12

0 a

0.11

0.57 Π

0.20

1093

0.16 0.05

5

0.11

0.13

0.15

0.03

EN

0.67 Π

0.13 4

0.05

P mW

0.10

0.75 Π

0.16

0.19 0.12

3

0.17

0.02 0.14

0.18

0.00 0.2

0.4

0.6

0.8

1.0

1.2

0.06

0.2 0.07

1.4

Ωm

2

Fig. 8. Plot of the logarithmic negativity versus the normalized effective detuning Δ∕ωm for different values of θ. The parameters are P  3 mW, G  1.3κ, and χ  0.05. Other parameters are the same as those in Fig. 4.

0.04 0.17 1

0.08 0.1 0.01

0 0.5

1.0

1.5

2.0

Ωm 6

0.18

0.1 0.01 0

0.17 5

b

0.06

0.23

0.11 0.21

0.03 0.12

4

0.19

0.07 0.14

P mW

temperature T and the maximum value of entanglement is achieved in the same range of the input power. Nonetheless, we will show that as in the case of bare cavity, entanglement and cooling are different phenomena and generally are optimized in different regimes. We study the effects of θ, G, and χ on EN separately to find the regime of maximal phonon–photon entanglement. We first study the behavior of EN when the phase of the field driving the OPA, i.e., θ, is varied. Figure 8 shows that entanglement increases when decreasing the phase of the field driving the OPA. This entanglement increment is related to the increasing of the photon number and the optomechanical coupling rate g1 [see Fig. 2(c)]. Figure 9 shows the entanglement as a function of the normalized effective detuning Δ∕ωm and laser power P for (a) G  0.6κ and (b) G  κ. It shows that the maximum possible values of the logarithmic negativity occur at low values of the input power and increase when increasing the gain nonlinearity. It is notable that here the entanglement increment is not just related to the increasing of the optomechanical coupling rate g1 . In Table 1 we summarize the maximum values of the logarithmic negativity E N and other parameters associated with it for input powers of P  2.5 mW and P  5 mW related to Fig. 9. In this table one can see that although the optomechanical coupling strength is larger for G  0.6κ for both values of the input power, EN is smaller. This feature suggests to examine other key parameters that determine the entanglement behavior. According to the data given in Table 1 for both G  0.6κ and G  κ the maximum values of the entanglement are obtained for η2 very close to unity, but the effective detuning and stability parameter η1 are smaller for G  κ. This result is in agreement with that obtained in Refs. [17,42] for a bare cavity, showing that the entanglement becomes maximal for η1  0. It is interesting to note that in contrast to the bare cavity for P  2.5 mW and G  0.6κ the maximum value of the logarithmic negativity is obtained for the minimum value of η1 , since at the end of the stability region Δ1  Γi  0 and η1  1. Figure 10 illustrates the behavior of EN when the anharmonicity parameter, χ, is varied. It shows that for the cavity with only the gain nonlinearity the amount of achievable optomechanical entanglement at the steady state is seriously limited by the stability condition. The Kerr nonlinearity makes the

0.15

0.02

3

0.15

0.22

0.08

0.2 0.16

2

0.04

0.13 0.09 1

0.05

0 0.5

1.0

1.5

2.0

Ωm

Fig. 9. Contour plot of the logarithmic negativity versus the normalized effective detuning Δ∕ωm and input power P for (a) G  0.6κ and (b) G  κ. The parameters are θ  0.67π, χ  0.05. Other parameters are the same as those in Fig. 4.

Table 1. Calculated Logarithmic Negativities, Normalized Effective Detunings, Normalized Optomechanical Couplings, and the Two Stability Parameters for the Input Powers P  2.5 mW and P  5 mW in Fig. 9 G∕κ

PmW 

EN

Δ∕ωm

η1

η2

g1 ∕ωm

0.6 1 0.6 1

2.5 2.5 5 5

0.20 0.21 0.18 0.23

0.53 0.43 0.68 0.30

0.77 0.67 0.77 0.66

0.93 0.87 0.87 0.88

0.56 0.47 0.69 0.66

1094

J. Opt. Soc. Am. B / Vol. 31, No. 5 / May 2014

0.30

Shahidani et al.

0.03

0.25 0.05

EN

0.20

0

0.15 0.10 0.05 0.00 0.0

0.5

1.0

1.5

2.0

Ωm

Fig. 10. Logarithmic negativity versus the normalized effective detuning Δ∕ωm for χ  0, χ  0.03, and χ  0.05. The parameters are θ  0.67π, G  κ, and P  6 mW. Other parameters are the same as those in Fig. 4.

stability region larger, and allows overcoming this limitation. As before, the reduction of the entanglement for χ  0.05 (in comparison to χ  0.03) is related to the increment of the stability parameter η1 . Until now we have studied the entanglement in the goodcavity limit (κ < ωm ) and have chosen the values of the input power and the detuning far from the multistability region. It is also interesting to examine the entanglement in the bad cavity limit and in the tristable regime. Figure 11 shows the entanglement and the stability parameter η1 as functions of the laser input power P in the tristable regime for the data of Fig. 3. It 0.30

(a) 0.25

I

EN

0.20 0.15 0.10

II

0.05 III

0.00 60

70

80

90

100

110

120

P mW 1.1 (b)

1.0 0.9

shows that for the first and second branches the entanglement is maximum at the end of the branches, while for the third branch the phonon–photon entanglement is null. This result might be interpreted as arising from the Kerr-induced shift of the resonance frequency of the cavity (for the third branch this shift is more than 1.7ωm ). Also, it can be seen that at the end of each stable branch η1 ≠ 0 and the entanglement is found only in the region with η1 < 0.65. It should be noted that in order to stay within the range of validity of the linearization approximation, we have been assured that the condition hδa† δaiss ≪ a2s is satisfied in all the results given above.

5. CONCLUSIONS In conclusion, we have studied the interaction of a singlemode field with both a weak Kerr medium and a parametric nonlinearity in an intrinsically nonlinear optomechanical system. We have investigated the stability behavior of the intracavity mean photon number, intensity, cooling, and stationary entanglement. We have found that the combination of the nonlinearities leads to a shift of the resonance frequency of the cavity toward lower values and the appearance of three real roots for the steady-state response of the system. We have derived the general condition for tristability in the system. Furthermore, we have studied the cooling and stationary entanglement by using the CM formalism. In particular, we have shown that the lower bound of the resolved sideband regime and the minimum attainable phonon number can be less than those of a bare cavity by controlling the parametric nonlinearity and the phase of the field driving the OPA. The weak Kerr nonlinearity can be used to extend the domain of the stability to the desired range of the effective detuning in which the effective temperature of the system is minimized. Also, the Kerr nonlinearity modifies the behavior of the stability parameter and allows us to approach significant entanglement simultaneously with the ground-state cooling of the mirror. In the investigation of the degree of stationary entanglement between the cavity and the mechanical modes we have identified four key parameters that affect the behavior of the entanglement. They are the effective detuning of the cavity, the optomechanical coupling, and the two stability parameters of the system. Also, as shown in this paper, the present scheme allows us to have significant entanglement in the tristable regime for the lower and middle branches, which makes the current scheme distinct from the bare optomechanical system.

ACKNOWLEDGMENTS

I

S. S. is grateful to M. Xiao and R. Ghobadi for useful discussions. The authors thank The Office of Graduate Studies of The University of Isfahan for their support.

Η1

0.8 0.7 III

0.6

II

REFERENCES

0.5 60

70

80

90

100

110

120

P mW

Fig. 11. (a) Logarithmic negativity and (b) stability parameter η1 as functions of the input power P corresponding to the three stable branches for θ  0.57π in Fig. 3. The dashed line corresponds to the unstable region.

1. V. Braginsky, S. E. Strigin, and S. P. Vyatchanin, “Analysis of parametric oscillatory instability in power recycled LIGO interferometer,” Phys. Lett. A 305, 111–124 (2002). 2. D. Rugar, R. Budakian, H. J. Mamin, and B. W. Chui, “Single spin detection by magnetic resonance force microscopy,” Nature 430, 329–332 (2004). 3. 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

Shahidani et al.

4.

5.

6. 7.

8. 9.

10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26.

state and single-phonon control of a mechanical resonator,” Nature 464, 697–703 (2010). J. D. Teufel, T. Donner, D. Li, J. W. Harlow, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W. Simmonds, “Sideband cooling of micromechanical motion to the quantum ground state,” Nature 475, 359–363 (2011). J. Chan, T. P. M. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Groblacher, M. Aspelmeyer, and O. Painter, “Laser cooling of a nanomechanical oscillator into its quantum ground state,” Nature 478, 89–92 (2011). O. Arcizet, P. F. Cohadon, T. Briant, M. Pinard, and A. Heidmann, “Radiation-pressure cooling and optomechanical instability of a micromirror,” Nature 444, 71–74 (2006). S. Gigan, H. R. Bohm, M. Paternostro, F. Blaser, G. Langer, J. B. Hertzberg, K. C. Schwab, D. Bauerle, M. Aspelmeyer, and A. Zeilinger, “Self-cooling of a micromirror by radiation pressure,” Nature 444, 67–70 (2006). C. H. Metzger and K. Karrai, “Cavity cooling of a microlever,” Nature 432, 1002–1005 (2004). S. Groblacher, K. Hammerer, M. R. Vanner, and M. Aspelmeyer, “Observation of strong coupling between a micromechanical resonator and an optical cavity field,” Nature 460, 724–727 (2009). Z. R. Gong, H. Ian, Yu.-x. Liu, C. P. Sun, and F. Nori, “Effective Hamiltonian approach to the Kerr nonlinearity in an optomechanical system,” Phys. Rev. A 80, 065801 (2009). C. Fabre, M. Pinard, S. Bourzeix, A. Heidmann, E. Giacobino, and S. Reynaud, “Quantum-noise reduction using a cavity with a movable mirror,” Phys. Rev. A 49, 1337–1343 (1994). S. Mancini and P. Tombesi, “Quantum noise reduction by radiation pressure,” Phys. Rev. A 49, 4055–4065 (1994). P. Rabl, “Photon blockade effect in optomechanical systems,” Phys. Rev. Lett. 107, 063601 (2011). A. Nunnenkamp, K. Borkje, and S. M. Girvin, “Single-photon optomechanics,” Phys. Rev. Lett. 107, 063602 (2011). A. Dorsel, J. D. McCullen, P. Meystre, E. Vignes, and H. Walther, “Optical bistability and mirror confinement induced by radiation pressure,” Phys. Rev. Lett. 51, 1550–1553 (1983). A. Gozzini, F. Maccarone, F. Mango, I. Longo, and S. Barbarino, “Light-pressure bistability at microwave frequencies,” J. Opt. Soc. Am. B 2, 1841–1845 (1985). R. Ghobadi, A. R. Bahrampour, and C. Simon, “Quantum optomechanics in the bistable regime,” Phys. Rev. A 84, 033846 (2011). S. Huang and G. S. Agarwal, “Enhancement of cavity cooling of a micromechanical mirror using parametric interactions,” Phys. Rev. A 79, 013821 (2009). A. Xuereb, M. Barbier, and M. Paternostro, “Multipartite optomechanical entanglement from competing nonlinearities,” Phys. Rev. A 86, 013809 (2012). T. Kumar, A. Bhattacherjee, and M. Mohan, “Dynamics of a movable micromirror in a nonlinear optical cavity,” Phys. Rev. A 81, 013835 (2010). B. Wielinga and G. J. Milburn, “Quantum tunneling in a Kerr medium with parametric pumping,” Phys. Rev. A 48, 2494–2496 (1993). W. Leon’ski, “Fock states in a Kerr medium with parametric pumping,” Phys. Rev. A 54, 3369–3372 (1996). T. Kaino and S. Tomaru, “Organic material for nonlinear optics,” Adv. Mater. 5, 172–178 (1993). P. D. Townsend, G. L. Baker, J. L. Jackel, J. A. Shelburne, and S. Etemad, “Nonlinear optical properties of organic materials II,” Proc. SPIE 1147, 256 (1989). J. S. Levy, M. A. Foster, A. L. Gaeta, and M. Lipson, “Harmonic generation in silicon nitride ring Resonators,” Opt. Express 19, 11415–11421 (2011). C. K. Law, “Interaction between a moving mirror and radiation pressurer: a Hamiltonian formulation,” Phys. Rev. A 51, 2537–2541 (1995).

Vol. 31, No. 5 / May 2014 / J. Opt. Soc. Am. B

1095

27. S. Mancini, V. I. Manko, and P. Tombesi, “Ponderomotive control of quantum macroscopic coherence,” Phys. Rev. A 55, 3042 (1997). 28. S. Bose, K. Jacobs, and P. L. Knight, “Preparation of nonclassical states in cavities with a moving mirror,” Phys. Rev. A 56, 4175–4186 (1997). 29. C. W. Gardiner and P. Zoller, Quantum Noise (Springer-Verlag, 1991). 30. V. Giovannetti and D. Vitali, “Phase-noise measurement in a cavity with a movable mirror undergoing quantum Brownian motion,” Phys. Rev. A 63, 023812 (2001). 31. L. Landau and E. Lifshitz, Statistical Physics (Pergamon, 1958). 32. D. Kleckner and D. Bouwmeester, “Sub-kelvin optical cooling of a micromechanical resonator,” Nature 444, 75–78 (2006). 33. D. Kleckner, W. Marshall, M. J. A. de Dood, K. N. Dinyari, B.-J. Pors, W. T. M. Irvine, and D. Bouwmeester, “High finesse optomechanical cavity with a movable thirty-micron-size mirror,” Phys. Rev. Lett. 96, 173901 (2006). 34. T. Carmon, H. Rokhsari, L. Yang, T. J. Kippenberg, and K. J. Vahala, “Temporal behavior of radiation-pressure-induced vibrations of an optical microcavity phonon mode,” Phys. Rev. Lett. 94, 223902 (2005). 35. I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series and Products (Academic, 1980). 36. A. Hurwitz, Selected Papers on Mathematical Trends in Control Theory, R. Bellman and R. Kabala, eds. (Dover, 1964). 37. F. Marquardt, J. G. Harris, and S. M. Girvin, “Dynamical multistability induced by radiation pressure in high-finesse micromechanical optical cavities,” Phys. Rev. Lett. 96, 103901 (2006). 38. C. Metzger, M. Ludwig, C. Neuenhahn, A. Ortlieb, I. Favero, K. Karrai, and F. Marquardt, “Self-induced oscillations in an optomechanical system driven by bolometric backaction,” Phys. Rev. Lett. 101, 133903 (2008). 39. 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). 40. J. Sheng, U. Khadka, and M. Xiao, “Realization of all-optical multistate switching in an atomic coherent medium,” Phys. Rev. Lett. 109, 223906 (2012). 41. D. Vitali, S. Gigan, A. Ferreira, H. R. Bohm, P. Tombesi, A. Guerreiro, V. Vedral, A. Zeilinger, and M. Aspelmeyer, “Optomechanical entanglement between a movable mirror and a cavity field,” Phys. Rev. Lett. 98, 030405 (2007). 42. C. Genes, A. Mari, P. Tombesi, and D. Vitali, “Robust entanglement of a micromechanical resonator with output optical fields,” Phys. Rev. A 78, 032316 (2008). 43. A. Mari and J. Eisert, “Gently modulating optomechanical system,” Phys. Rev. Lett. 103, 213603 (2009). 44. F. Marquardt, J. P. Chen, A. A. Clerck, and S. M. Girvin, “Quantum theory of cavity-assisted sideband cooling of mechanical motion,” Phys. Rev. Lett. 99, 093902 (2007). 45. I. Wilson-Rae, N. Nooshi, W. Zwerger, and T. J. Kippenberg, “Theory of ground state cooling of a mechanical oscillator using dynamical backaction,” Phys. Rev. Lett. 99, 093901 (2007). 46. M. Paternostro, D. Vitali, S. Gigan, M. S. Kim, C. Brukner, J. Eisert, and M. Aspelmeyer, “Creating and probing multipartite macroscopic entanglement with light,” Phys. Rev. Lett. 99, 250401 (2007). 47. J. M. Dobrindt, I. Wilson-Rae, and T. J. Kippenberg, “Parametric normal-mode splitting in cavity optomechanics,” Phys. Rev. Lett. 101, 263602 (2008). 48. C. Genes, A. Mari, D. Vitali, and P. Tombesi, “Quantum effects in optomechanical systems,” Adv. At. Mol. Opt. Phys. 57, 33–86 (2009). 49. G. Adesso, A. Serafini, and F. Illuminati, “Extremal entanglement and mixedness in continuous variable systems,” Phys. Rev. A 70, 022318 (2004).