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

Robust entanglement and steering in open Dicke models with individual atomic spontaneous emission and dephasing

Open Access Open Access

Abstract

In this paper, we study steady-state quantum entanglement and steering in an open Dicke model where cavity dissipation and individual atomic decoherence are taken into account. Specifically, we consider that each atom is coupled to independent dephasing and squeezed environments, which makes the widely-adopted Holstein-Primakoff approximation invalid. By discovering the features of quantum phase transition in the presence of the decohering environments, we mainly find that (i) in both normal and superradiant phases, the cavity dissipation and individual atomic decoherence can improve the entanglement and steering between the cavity field and atomic ensemble; (ii) the individual atomic spontaneous emission leads to the appearance of the steering between the cavity field and atomic ensemble but the steering in two directions cannot be simultaneously generated; (iii) the maximal achievable steering in normal phase is stronger than that in superradiant phase; (iv) the entanglement and steering between the cavity output field and the atomic ensemble are much stronger than that with the intracavity, and the steerings in two directions can be achieved even with the same parameters. Our findings reveal unique features of quantum correlations in the open Dicke model in the presence of individual atomic decoherence processes.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Dicke model was first proposed by Dicke in the 1950s to describe the coherent interaction between an ensemble of two-level atoms and a single-mode bosonic field [1]. It has been shown that under collective spontaneous emission of atoms, the atomic system in Dicke model undergoes a quantum phase transition from the normal to superradiant phase in the thermodynamic limit [25]. In recent years, Dicke model has been realized in various systems [611], such as atom-cavity setup [12], waveguide-emitter [13] superconducting circuits [14] setups. Besides quantum phase transition, new interesting features, like quantum critical behavior [15,16], quantum chaos [17,18] and quantum entanglement and correlations [1923], have also been discovered in the Dicke model. In paricular, the entanglement in the model has a critical point similar to the phase transition, showing interesting characteristics [8,12,2431]. However, in much work on the open Dicke model only the cavity dissipation has been considered to perform the Holstein-Primakoff approximation [32]. Recent studies has shown that the decay of atomic internal states due to spontaneous emission has nontrivial effects on the mean-field nonequilibrium steady states for spin and photon observables [3335] and quantum-fluctuation correlations of atoms and field [29].

In this paper, we intend to explore a new type of quantum correlations — quantum steering — in an open Dicke model in which not only the cavity dissipation but also the spontaneous emission and dehasing of individual atoms are taken into account. We consider the situation that each atom is immersed in an independent squeezed bath and undergoes dephasing process. Quantum steering originates from the well-known Einstein-Podolsky-Rosen (EPR) paradox, which describes the situation where two remote observers, say Alice and Bob, share a pair of perfect entangled particles and Alice can nonlocally manipulate the particle’s states in Bob via different types of local measurements on his own particle [36]. Such ability to nonlocally steer quantum states of remote particles was originally termed by Schrödinger as steering [37]. It has recently been revealed by Wiseman et al. [38] that EPR steering is a type of quantum correlation intermediate between entanglement (nonseparability) [39] and Bell nonlocality (the violation of Bell inequality) [40]. This unique feature makes steering a physical resource for one-sided device-independent quantum cryptography [41], subchannel discrimination [42], and secure quantum teleportation [43].

Specifically, here we found that in both normal and superradiant phases, the cavity dissipation and individual atomic decohernece can improve the entanglement and steering between the cavity field and atomic ensemble. The individual atomic spontaneous emission leads to the appearance of the steering between the cavity field and atomic ensemble but the steerings in two directions cannot be simultaneously generated. The maximal achievable steering in normal phase is stronger than that in superradiant phase. Furthermore, the entanglement and steering between the cavity output field and atomic ensemble is much stronger than that with the intracavity and the steerings in two directions can be achieved even with the same parameters.

This paper is organized as follows: In Sec. II, we introduce the model. In Sec. III and IV, the steady-state solution of the mean field and quantum fluctuations of the system are presented. In Sec. V, the quantum steering and entanglement with the intracavity field and output field are investigated. In the last section, some discussion and the conclusion are presented.

2. Model

As shown in Fig. 1, we consider a Dicke system in which an ensemble of two-level atoms is strongly coupled to a single-mode cavity field with the resonance frequency $\omega _0$ and denoted by the annihilation (creation) operator $\hat {a}_f$ ($\hat {a}^{\dagger}_f$). Each atom is described by the Pauli operator $\hat {\sigma }_j$ ($j=x,y,z$), where the operators $\hat {\sigma }_z=\frac {1}{2}(|\uparrow \rangle |\uparrow \rangle -|\downarrow \rangle |\downarrow \rangle )$ and $\hat {\sigma }_{x, y}=\frac {1}{2}(|\uparrow \rangle |\downarrow \rangle \pm |\downarrow \rangle |\uparrow \rangle )$, with $|\uparrow \rangle$ and $|\downarrow \rangle$ respectively denoting the spin-up and spin-down states of the frequency difference $\omega _z$. The interaction between the cavity field and the collective atoms can be described by the Hamiltonian ($\hbar =1$)

$$\hat{H}=\omega_0\hat{a}^{\dagger}_f \hat{a}_f+\omega_z \hat{S}_z +\frac{\lambda}{\sqrt{N_a}}(\hat{a}_f+\hat{a}^{\dagger}_f)(\hat{S}_+{+}\hat{S}_-),$$
where $\lambda$ is the coupling strength, the collective operators $\hat {S}_j=\sum _{k=1}^{N_a}\hat {\sigma }^{k}_j$, the raising and lowering operators $\hat {S}_\pm =\hat {S}_x\pm \hat {S}_y$, and $N_a$ is the number of the atoms.

 figure: Fig. 1.

Fig. 1. Schematic plot of the system. An ensemble of identical two-level systems is collectively coupled to a cavity with loss rate $\kappa$. Each atom undergoes independent spontaneous and dephasing processes with rates $\gamma _k$ and $\gamma _c$ respectively. The output field is filtered to define temporal modes to study the quantum correlations between the ensemble and output field.

Download Full Size | PDF

When considering that each atom is immersed in individual decoherencing environments, we can write the master equation for the density matrix $\hat \rho$ of the system as

$$\begin{aligned} \frac{d}{dt}\hat{\rho}(t)&={-}i[\hat{H},\hat{o}]+\kappa\mathcal{L}_{\kappa}[\hat a]\hat \rho+\sum_{k=1}^{N_a}\mathcal {\tilde{L}}_{\gamma_c}[\hat \sigma_z^k]\hat\rho\\ &~~+\sum_{k=1}^{N_a}\Big[\mathcal{L}_{\gamma_k(N_{\rm th}+1)}[\hat \sigma_-^k]\hat\rho+\mathcal{L}_{\gamma_kN_{\rm th}}[\hat \sigma_+^k]\hat\rho\\ &~~+\mathcal {\tilde{L}}_{\gamma_kM}[\hat \sigma_-^k]\hat \rho+\mathcal {\tilde{L}}_{\gamma_kM^*}[\hat \sigma_+^k]\hat\rho\Big], \end{aligned}$$
where $\mathcal {L}_{x}[\hat O]\hat \rho =x(\hat {O}^\dagger\hat {\rho }\hat {O}-\frac {1}{2}\hat {O}^\dagger\hat {O}\hat {\rho }-\frac {1}{2}\hat {\rho }\hat {O}^\dagger\hat {O})$ and $\mathcal {\tilde {L}}_{x}[\hat O]\hat \rho =x(\hat {O}\hat {\rho }\hat {O}-\frac {1}{2}\hat {O}^2\hat {\rho }-\frac {1}{2}\hat {\rho }\hat {O}^2)$, with arbitrary operator $\hat O$ and parameter $x$. The second term describes the cavity dissipation in vacuum with the rate $\kappa$, the third term corresponds to individual atomic dephasing at the rate $\gamma _c$, and the remaining terms characterizes the atoms spontaneously emit into individual squeezed environments at the rate $\gamma$, with identical thermal excitation number $N_{\rm th}$ and two-photon correlation $M$ satisfying $|M|\le \sqrt {N_{\rm th}(N_{\rm th}+1)}$. When $M=0$, the squeezed environments reduces to generic thermal ones with $N_{\rm th}\neq 0$. For a squeezed vacuum environment, $N_{\rm th}=\sinh ^{2}r$ and $M=\sinh r\cosh r$, with squeezing parameter $r$. When $r=0$, it means that the atoms are damped by vacuum fluctuations. For example, in realistic situations [12,34] the parameters $\omega _z/2\pi \approx 200$ kHz, $\gamma /2\pi \approx \kappa /2\pi \approx 100$ kHz, $\lambda /2\pi \approx 120$ kHz, and $\gamma _c\approx \gamma$ [35].

3. Equations of mean fields

We at first study quantum phase transition of the open Dicke model in the thermodynamic limit ($N_a\rightarrow \infty$). By defining the following scaled operators

$$\hat{m}_j=\frac{\hat{S}_j}{N_a},~\hat{X}_f=\frac{\hat{a}_f+\hat{a}^\dagger_f}{\sqrt{2N_a}},~\hat{Y}_f=\frac{i(\hat{a}_f-\hat{a}^\dagger_f)}{\sqrt{2N_a}},$$
neglecting the correlations between the field and atoms in the thermodynamic limit, from Eq. (2) the mean-field equations of the scaled operators can be derived as
$$\frac{d}{dt}\langle \hat{m}_x\rangle={-}\omega_z \langle \hat{m}_y\rangle-\frac{\Gamma_x}{2}\langle \hat{m}_x\rangle,$$
$$\frac{d}{dt}\langle \hat{m}_y\rangle=\omega_z \langle \hat{m}_x\rangle-2\sqrt{2}\lambda \langle \hat{X}_f\rangle \langle \hat{m}_z\rangle-\frac{\Gamma_y}{2}\langle \hat{m}_y\rangle,$$
$$\frac{d}{dt}\langle \hat{m}_z\rangle=2\sqrt{2}\lambda \langle \hat{X}_f\rangle \langle \hat{m}_y\rangle-\Gamma_z \langle \hat{m}_z\rangle-\frac{\gamma}{2},$$
$$\frac{d}{dt}\langle \hat{Y}_f\rangle=\omega_0 \langle \hat{m}_X\rangle+2\sqrt{2}\lambda \langle \hat{m}_x\rangle-\frac{\kappa}{2} \langle \hat{Y}_f\rangle,$$
$$\frac{d}{dt}\langle \hat{X}_f\rangle=\omega_0 \langle \hat{Y}_f\rangle-\frac{\kappa}{2} \langle \hat{X}_f\rangle,$$
where $\Gamma _x=\gamma (2N_{\rm th}+1-2M)+\gamma _c$, $\Gamma _y=\gamma (2N_{\rm th}+1+2M)+\gamma _c$, $\Gamma _z=\gamma (2N_{\rm th}+1)$, and $\gamma =\gamma _k$.

The stationary solutions of the above equations can be found. When the coupling $\lambda <\lambda _c$, where the critical coupling strength

$$\lambda_c =\sqrt{\frac{(2N_{\rm th}+1)(\kappa^2/4+\omega_0^{2})(\Gamma_x\Gamma_y/4+\omega_z^{2})}{4\omega_0\omega_z}},$$
we have
$$\begin{aligned} &\langle\hat{m}_x\rangle_s=0,\\ &\langle\hat{m}_y\rangle_s=0,\\ &\langle\hat{m}_z\rangle_s={-}\frac{1}{2(2N_{\rm th}+1)},\\ &\langle\hat{Y}_f\rangle_s=0,\\ &\langle\hat{X}_f\rangle_s=0, \end{aligned}$$
the system is in the normal phase, while when the coupling $\lambda >\lambda _c$, the stationary solutions become into
$$\begin{aligned} &\langle\hat{m}_x\rangle_s={\pm}\sqrt{\frac{\gamma}{2\Gamma_xg}}\frac{\omega_z\lambda_c}{\lambda}\sqrt{1-\frac{\lambda^{2}_c}{\lambda^{2}}},\\ &\langle\hat{m}_y\rangle_s={\mp}\sqrt{\frac{\Gamma_x\gamma}{8g}}\frac{\lambda_c}{\lambda}\sqrt{1-\frac{\lambda^{2}_c}{\lambda^{2}}},\\ &\langle\hat{m}_z\rangle_s={-}\frac{\lambda^{2}_c}{2(2N_{\rm th}+1)\lambda^{2}},\\ &\langle\hat{Y}_f\rangle_s={\pm}\frac{2\kappa\omega_z\lambda_c}{(\kappa^2/4+\omega_0^{2})}\sqrt{\frac{\gamma}{\Gamma_xg}}\sqrt{1-\frac{\lambda^{2}_c}{\lambda^{2}}},\\ &\langle\hat{X}_f\rangle_s={\mp}\frac{2\omega_z\omega_0\lambda_c}{(\kappa^2/4+\omega_0^{2})}\sqrt{\frac{\gamma}{\Gamma_xg}}\sqrt{1-\frac{\lambda^{2}_c}{\lambda^{2}}}, \end{aligned}$$
with $g=(2N_{\rm th}+1)(\Gamma _x\Gamma _y/4+\omega _z^{2})$, meaning that the system is in superradiant phase and it undergoes a quantum phase transition. From Eq. (5) we see that the individual atomic dephasing ($\gamma _c$) and spontaneous emission in thermal environments ($\gamma$, $N_{\rm th}$) leads to the increasing of the critical coupling $\lambda _c$, while the two-photon correlations ($M$) decreases $\lambda _c$. In addition, for the normal phase, thermal noise ($N_{\rm th}$) excites the atoms without the change of the mean values of the other variables.

4. Quantum-fluctuation matrix correlations

We next consider quantum fluctuations of the system. Since the total spin number is no longer conservative due to individual atomic decoherence [44,45], the conventional approach of the Holstein-Primakoff approximation [28,32,46] to bosonify atomic variables is invalid. Instead, we consider the rotation of the reference frame for the atomic variables by rotating $\hat m_x$ and $\hat m_y$ an angle $-\varphi$ around $z$-axis and then rotating $\hat m_z$ and $\hat m_x$ an angle $-\theta$ around $y$-axis. In the new frame, the rotated spin operator $\hat {\tilde {m}}_z$ is aligned to the spin vector $\hat {\vec {m}}$, and therefore the rotated spin variables $\hat {\tilde {m}}_{x,y}$ are pure fluctuations and for $N_a\rightarrow \infty$ we have $\hat {\tilde {{m}}}_z\simeq \lvert \hat {\vec {m}}\rvert$. From the commutation relation $[\hat {\tilde {{m}}}_x, \hat {\tilde {{m}}}_y]=i\lvert \hat {\vec {m}}\rvert$ and by defining

$$\hat{q}_a=\frac{\hat{\tilde{{m}}}_x}{\sqrt{\lvert\hat{\vec{m}}\rvert}},~~ \hat{p}_a=\frac{\hat{\tilde{{m}}}_y}{\sqrt{\lvert\hat{\vec{m}}\rvert}},$$
we have
$$[\hat{q}_a, \hat{p}_a]=i.$$

This means that in the rotated frame the atomic ensemble can be treated as a harmonic oscillator and we thus can study continuous variable quantum correlations between the cavity field and the atomic ensemble [22,29].

The above rotation corresponds to a local unitary operation on the spin, which can be described by the matrix

$$\begin{aligned} R_{\theta\varphi}=&\left( \begin{array}{ccccc} \cos\theta\cos\varphi & \cos\theta\sin\varphi & -\sin\theta & 0 & 0\\ -\sin\varphi & \cos\varphi & 0 & 0 & 0\\ \sin\theta\cos\varphi & \sin\theta\sin\varphi & \cos\theta & 0 & 0\\ 0 & 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 0 & 1\\ \end{array} \right), \end{aligned}$$
where the angles $\theta$ and $\varphi$ are determined by
$$\theta=\arccos(\frac{m_z}{|\overrightarrow{m}|}), $$
$$\varphi=\left\{ \begin{array}{ll} 2\pi-\arccos(\frac{m_x}{|\overrightarrow{m}|\sin\theta}), & m_y<0, \\ 0, & m_y=0,~ m_x=0,\\ \arccos(\frac{m_x}{|\overrightarrow{m}|\sin\theta}), & otherwise. \end{array} \right. $$

It shows that if $m_z\rightarrow |\overrightarrow {m}|$, i.e., the spins are almost aligned to the $z$ component, the rotation matrix is unnecessary since $R_{\theta \varphi }\simeq 1$.

We proceed to determine the correlation matrix between the atomic and field fluctuations characterized by $\delta \hat o=\sqrt {N_a}(\hat o-\langle \hat o\rangle _s)$ ($\hat o$ being $\hat {m}_j$, $\hat {X}_f$ or $\hat {Y}_f$), which is rescaled by a factor of $\sqrt {N_a}$. For the fluctuation operators $u(t)=(\delta \hat {m}_x, \delta \hat {m}_y, \delta \hat {m}_z, \delta \hat {Y}_f, \delta \hat {X}_f)^{T}$, from Eq. (2), the equation of motion for $u(t)$ can be derived as

$$\frac{d}{dt}u(t)=Au(t)+n^{\rm in}(t),$$
where the drift matrix
$$\begin{aligned} A=&\left( \begin{array}{ccccc} -\frac{\Gamma_x}{2} & -\omega_z & 0 & 0 & 0\\ \omega_z & -\frac{\Gamma_y}{2} & -2\sqrt{2}\lambda\langle\hat{X}_f\rangle & 0 & -2\sqrt{2}\lambda\langle\hat{m}_z\rangle\\ 0 & 2\sqrt{2}\lambda\langle\hat{X}_f\rangle & -\Gamma_z & 0 & 2\sqrt{2}\lambda\langle\hat{m}_y\rangle\\ 2\sqrt{2}\lambda & 0 & 0 & -\frac{\kappa}{2} & \omega_0\\ 0 & 0 & 0 & -\omega_0 & -\frac{\kappa}{2}\\ \end{array} \right), \end{aligned}$$
and $n^{\rm in}(t)=(\hat {m}_x^{\rm in}, \hat {m}_y^{\rm in}, \hat {m}_z^{\rm in}, \sqrt {\kappa }\hat {Y}^{\rm in}_f, \sqrt {\kappa }\hat {X}^{\rm in}_f)^{T}$ are input noises of the atoms and cavity field. The correlation matrix $V_{ii'}=\frac {1}{2}\langle u_i u_{i'}+u_{i'}u_i\rangle$ can be obtained as
$$\frac{d}{dt}V=AV+VA^{T}+D,$$
where the noise matrix $D$ reads
$$\begin{aligned} D=&\left( \begin{array}{ccccc} \frac{\Gamma_x}{4} & 0 & \frac{\langle\hat{m}_x\rangle}{2}\gamma & 0 & 0\\ 0 & \frac{\Gamma_y}{4} & \frac{\langle\hat{m}_y\rangle}{2}\gamma & 0 & 0\\ \frac{\langle\hat{m}_x\rangle}{2}\gamma & \frac{\langle\hat{m}_y\rangle}{2}\gamma & \frac{\Gamma_z}{2}+\langle\hat{m}_z\rangle\gamma & 0 & 0\\ 0 & 0 & 0 & \frac{\kappa}{2} & 0\\ 0 & 0 & 0 & 0 & \frac{\kappa}{2}\\ \end{array} \right), \end{aligned}$$
of which the submatrix of atomic variables can be derived with the generalized Einstein relation [47].

For the fluctuation operators $\tilde {u}(t)=(\delta \hat {q}_a, \delta \hat {p}_a, \delta \hat {\tilde {m}}_z, \delta \hat {Y}_f, \delta \hat {X}_f)^{T}$, with Eq. (8) and Eq. (10) we have $\tilde {u}(t)=WR_{\theta \varphi }u(t)$ and thus the corresponding correlation matrix $\tilde {V}_{ii'}\equiv \frac {1}{2}\langle \tilde {u}_i \tilde {u}_{i'}+\tilde {u}_{i'}\tilde {u}_i\rangle$ can be given by

$$\tilde{V}=WR_{\theta\varphi}VR^{T}_{\theta\varphi}W,$$
where
$$\begin{aligned} W=&\left( \begin{array}{ccccc} \frac{1}{\sqrt{\overrightarrow{m}}} & 0 & 0 & 0 & 0\\ 0 & \frac{1}{\sqrt{\overrightarrow{m}}} & 0 & 0 & 0\\ 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 0 & 1\\ \end{array} \right). \end{aligned}$$

We consider the stationary solution $\tilde {V}(t\rightarrow \infty )$ of the Eq. (16), omit the third row and the third column and a $4\times 4$ reduced correlation matrix which carries the information we need for the quantum correlations can be finally obtained as

$$\begin{aligned} \tilde{\sigma}=&\left( \begin{array}{cc} \tilde{V}_a & \tilde{V}_c\\ \tilde{V}_c ^{T} & \tilde{V}_f\\ \end{array} \right), \end{aligned}$$
where $\tilde {V}_a$, $\tilde {V}_f$, $\tilde {V}_c$ are $2\times 2$ sub-block matrices of $\tilde {\sigma }$, they represent the atomic ensemble, the cavity mode and the second-order quantities that link them, respectively.

With the steady-state correlation matrix we can investigate quantum correlations between the cavity field and the atomic ensemble. The atom-field bipartite entanglement can be quantified by the logarithmic negativity [4850]

$$E_n=\max[0,-\log(2\tilde{\nu}_-)],$$
with $\tilde {\nu }_-=\frac {\sqrt {\tilde {\Delta }-\sqrt {\tilde {\Delta }^{2}-4\det \tilde \sigma }}}{\sqrt {2}}$ being the smallest of the two symplectic eigenvalues of the partial transpose correlation matrix and $\tilde {\Delta }=\det \tilde {V}_a+\det \tilde {V}_f-2\det \tilde \sigma$. The quantum steering from cavity mode to the atomic mode (cavity-to-atomic steering) and the reverse steering (atomic-to-cavity steering) can be verified by [51]
$$S_{a|f}=\max[0,\frac{1}{\mathfrak{S}_{a|f}}-1],$$
$$S_{f|a}=\max[0,\frac{1}{\mathfrak{S}_{f|a}}-1],$$
where $\mathfrak {S}_{a|f}=4V_{inf}(\hat {q}_{a}^{\theta _{a}})V_{inf}(\hat {p}_{a}^{\theta _{a}})$, $\mathfrak {S}_{f|a}=4V_{inf}(\hat {Y}_{f}^{\theta _{f}})V_{inf}(\hat {X}_{f}^{\theta _{f}})$, with $V_{inf}(\hat {X}_{f}^{\theta _{f}})=V(\hat {X}_{f}^{\theta _{f}}+h_{fx}\hat {X}_{f}^{\theta _{f}})$, $V_{inf}(\hat {Y}_{f}^{\theta _{f}})=V(\hat {Y}_{f}^{\theta _{f}}+h_{fy}\hat {Y}_{f}^{\theta _{f}})$ represent the inferred variances of cavity mode, $V_{inf}(\hat {p}_{a}^{\theta _{a}})=V(\hat {p}_{a}^{\theta _{a}}+h_{ax}\hat {p}_{a}^{\theta _{a}})$, $V_{inf}(\hat {q}_{a}^{\theta _{a}})=V(\hat {q}_{a}^{\theta _{a}}+h_{ay}\hat {p}_{a}^{\theta _{a}})$ represent the inferred variances of atomic ensemble. The gain factors $h_{fx(y)}$, $h_{ax(y)}$ and the angles $\theta ^{f, a}$ are chosen such that the inferred variances minimize. Actually, the state is steerable if and only if $\mathfrak {S}_{a|f}<1$ or $\mathfrak {S}_{f|a}<1$.

5. Results

5.1 Intracavity case

We first investigate the entanglement and steering between the intracavity field and atomic ensemble. Fig. 2 plots the dependencies of the steering $S_{f|a}$, $S_{a|f}$ and entanglement $E_n$ on the cavity dissipation $\kappa$ and individual atomic spontaneous emission rate $\gamma$ and dephasing rate $\gamma _c$. Since the critical point increases with the increase of $\gamma$ and $\kappa$, we consider temporally the regime in normal phase for which the coupling $\lambda$ is fixed at $0.1$. It is shown that in the absence of individual atomic spontaneous emission and dephasing, the steering in either direction is unachievable. However, when the spontaneous emission is present ($\gamma \neq 0$ and $\gamma _c=0$), we have $S_{f|a}\neq 0$ and $S_{a|f}\neq 0$, meaning that the steering in two directions can occur, as shown in Fig. 2 ($a_1$) and ($a_2$). This is because that the individual atomic decoherence processes decrease the fluctuations of the atomic ensemble. Further, we see from Fig. 2 ($a_{1,2}$) and ($b_{1,2}$) that the steering in two directions cannot be simultaneously achieved for the parameters $\kappa$, $\gamma$ and $\gamma _c$. This means that one can only achieve one-way steering between the intracavity field and atomic ensemble. The atom-to-field steering $S_{f|a}$ only exists for $\{\gamma,~\gamma _c\}\lesssim 1$, whereas the inverse steering $S_{a|f}$ is just absent for $\{\gamma, \gamma _c\}\lesssim 0.3$. Likewise, the steering $S_{f|a}$ occurs for $\kappa \gtrsim 1.5\omega _z$ and the steering $S_{a|f}$ can exist for much less values of $\kappa$. Note that for the same rates $\gamma =\kappa$, only the steering $S_{a|f}$ is present, which means that the local states of the atomic ensemble and intracavity field is asymmetric even for the balanced dissipation rates. Different from the steering, the entanglement exists even when $\gamma =0$ and $\gamma _c=0$, as already studied in Ref. [28]. More importantly, we can see that the steering and entanglement can be increased by the individual atomic decoherence and cavity decoherence. It is clearly shown that the steering and entanglement increases at first and then decreases as $\gamma$, $\gamma _c$ and $\kappa$ arise continuously. Thus, for the open Dicke model, individual atomic decoherence and cavity dissipation may be constructive to the quantum correlations between the field and atomic ensemble. In addition, the maximal achievable entanglement and steering are achieved for different parameters.

 figure: Fig. 2.

Fig. 2. Density plot of the steering $S_{f|a}$, $S_{a|f}$ and entanglement $E_n$ between intracavity field and spin field versus $\gamma$ and $\kappa /\omega _z$ in ($a_1$)-($a_3$), $\gamma _c$ and $\kappa /\omega _z$ in ($b_1$)-($b_3$), $r$ and $\kappa /\omega _z$ in ($c_1$)-($c_3$). Related parameters in plots: $\omega _0=\omega _z=0.2$, $\lambda =0.1$. In (a1)-(a3) $\gamma _c=r=0$; in ($b_1$)-($b_3$) $\gamma =0.1$, $r=0$; in ($c_1$) and ($c_3$) $\gamma =\gamma _c=0.1$; in ($c_2$) $\gamma =0.1$, $\gamma _c=1.5$.

Download Full Size | PDF

In order to observe the characteristics of quantum steerable correlations in the superradiant phase, we plot the dependence of the steering and entanglement on the coupling $\lambda$ for different values of $\gamma$ and $\gamma _c$ in Fig. 3, where the critical coupling point $\lambda _c$ is marked with asterisks. The value of $\kappa$ is fixed at $6\omega _z$. Note that the values of $\gamma$ are different in Fig. 3 ($a_1$) and ($a_2$), since the steerings in two directions cannot be achieved simultaneously for the same value. It is found that the steering $S_{a|f}$ only occurs when $\gamma >6\omega _z$ and always be zero in the superradiant phase, while $S_{a|f}$ only occurs when $\gamma <6\omega _z$ and has nonzero values in the superradiant phase. The maximal achievable steering is much weaker than that in the normal phase. It is also shown that the steering is almost absent near the critical coupling points. This is because that near the critical coupling point, the quantum fluctuations of quadratures of the cavity field and atomic ensemble (described by their variances) become much larger, which makes the Gaussian quantum states of the two subsystems are difficult to be locally steered (by Gaussian measurement) via the finite quantum correlations. Further, when the system is in the superradiant phase, the spontaneous emission process can also improve the steering, as shown in Fig. 3 ($a_1$). For the same value of $\gamma$, the optimal entanglement in the superradiant phase can be larger than that in the normal phase, as shown in Fig. 3 ($a_3$), different from the behavior of the steering. When $\gamma =0$, the entanglement peaks around the critical coupling point $\lambda _c$, but it may become local minimal around $\lambda _c$ for nonzero $\gamma$. In both normal and superradiant phases, the steerings in two directions and entanglement do not increase monotonically with the coupling $\lambda$ and they exhibit local peaks with respect to $\lambda$. But for the entanglement, it has local dips or peaks near the critical coupling $\lambda _c$. Figures 3 ($b_1$)-($b_3$) depicts the effect of the dephasing processes on the steering and entanglement. Since quantum phase transition disappears in the open Dicke model only with dephasing processes, we set $\gamma =0.5\omega _z=0.1$ in Figs. 3 ($b_1$) and ($b_3$) and $\gamma =7.5\omega _z=1.5$ in Fig. 3 ($b_2$). We see that when $\gamma _c$ increases, the steering $S_{f|a}$ increases at first and then decreases and $S_{a|f}=0$. As shown in Figs. 2 and 3 ($c_1$)-($c_3$), the local squeezed environment with small squeezing can also lead to the enhancement of the entanglement and steering. This is due to that stronger squeezing causes the amplification of the noise and the decrease of the entanglement and steering.

 figure: Fig. 3.

Fig. 3. The dependencies of the steering $S_{f|a}$ and $S_{a|f}$, entanglement $E_n$ between intracavity field and spin field on $\lambda$ for different $\gamma$ in ($a_1$)-($a_3$), $\gamma _c$ in ($b_1$)-($b_3$), $r$ in ($c_1$)-($c_3$). Related parameters in plots: $\kappa =1.2$. In ($a_1$)-($a_3$) $\gamma _c=0$, $r=0$; in ($b_1$) and ($b_3$) $\gamma =0.1$, $r=0$; in ($b_2$) $\gamma =1.5$, $r=0$; in ($c_1$) and ($c_3$) $\gamma =\gamma _c=0.1$; in ($c_2$) $\gamma =\gamma _c=1.5$; the other parameters are the same as Fig. 2.

Download Full Size | PDF

5.2 Output case

Since the intracavity field in the open system cannot be observed directly, we then study the quantum steering and entanglement between the output field emitted from the cavity and the atomic ensemble. The output field is related to the intracavity field via the input-output relation [52]

$$\hat{a}^{\rm out}_f(t)=\sqrt{\kappa}\delta\hat{a}_f(t)-\hat{a}^{\rm in}_f(t).$$

Since the output field is spectral continuous, we define temporal mode by

$$\hat{a}^{\rm out}_f(t)=\int^{t}_{-\infty}f(t-s)\hat{a}^{\rm out}_f(s)ds,$$
where the filter function
$$f(t)=\sqrt{\frac{2}{\tau}}\theta(t)e^{-(\tau^{{-}1}+i\Omega_f)},$$
defining a specified spectral component which possesses central frequency $\Omega _f$ and bandwidth $\tau ^{-1}$. For $\tilde {u}^{out}(t)=(\delta \hat {q}_a, \delta \hat {p}_a, \delta \hat {\tilde {m}}_z, \delta \hat {Y}_f^{\rm out}, \delta \hat {X}_f^{\rm out})^{T}$, its Fourier transform $\tilde {u}^{out}(\omega )=(\delta \hat {q}_a(\omega ), \delta \hat {p}_a(\omega ),$ $\delta \tilde {\hat {m}}_z(\omega ), \delta \hat {Y}^{out}_f(\omega ), \delta \hat {X}^{out}_f(\omega ))^{T}$ can be found from Eq. (12) and Eq. (21), given by
$$\tilde{u}^{out}(\omega)=T(\omega)[\tilde{M}(\omega)+\frac{P^{out}}{\kappa}]WR_{\theta\varphi} n^{\rm in}(\omega).$$

Here $\tilde {M}(\omega )=(WR_{\theta \varphi }AR^{T}_{\theta \varphi }W^{-1}+i\omega )^{-1}$, $P^{out}=Diag[0,0,0,1,1]$,

$$\begin{aligned} T(\omega)=&\left( \begin{array}{ccccc} 1 & 0 & 0 & 0 & 0\\ 0 & 1 & 0 & 0 & 0\\ 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & \sqrt{\kappa}t_1 & \sqrt{\kappa}t_2\\ 0 & 0 & 0 & -\sqrt{\kappa}t_2 & \sqrt{\kappa}t_1\\ \end{array} \right), \end{aligned}$$
with $t_1=\frac {1}{2}[f(\omega )+f^{*}(-\omega )]$, $t_2=\frac {i}{2}[f(\omega )-f^{*}(-\omega )]$, and $f(\omega )$ being the Fourier transform of $f(t)$. The correlation matrix $\tilde {V}^{out}$ of the filtered cavity output mode and the atomic ensemble can be derived as [53,54]
$$\begin{aligned} \tilde{V}^{out}=&\int d\omega T(\omega)[\tilde{M}(\omega)+\frac{P^{out}}{\kappa}]\\ &\tilde{D}(\omega)[\tilde{M}^{\dagger}(\omega)+\frac{P^{out}}{\kappa}]T^{\dagger}(\omega). \end{aligned}$$

In Fig. 4, we plot the dependence of the steering and entanglement between the filtered output field and the atomic ensemble on the spectral frequency $\Omega _f$ and bandwidth $\tau ^{-1}$, by considering the coupling strengthes respectively in the normal phase and superradiant phase. Different from the intracavity case, one can see that the steerings in both directions can be achieved when $\gamma$ and $\gamma _c$ deviate from zero. By comparing Fig. 3 and Fig. 4, we can see that the steering and the entanglement with the output field are much stronger than that with the intracavity field for the same parameters. This is because that due to the fact that the latter can be considered as the sum of the steering and entanglement over all the spectral components. Therefore, robust atom-field steering and entanglement are built up by filtering the cavity output, which also means an effective entanglement and steering distillation. It is also shown that the steering and entanglement become optimal around $\Omega _f=-\omega _z$ and $\tau ^{-1}=\omega _z$ in the normal phase, while in the superradiant phase the optimal point of $\Omega _f$ shifts to the left on $\Omega _f=-2.5\omega _z$ and $\tau ^{-1}$ still keeps the value of $\omega _z$.

 figure: Fig. 4.

Fig. 4. The dependencies of the steering $S_{f|a}$ and $S_{a|f}$, entanglement $E_n$ between output field and spin field on $\Omega _f/\omega _z$ in ($a_1$), ($b_1$) and $\omega _z\tau$ in ($a_2$), ($b_2$). In ($a_1$)-($a_2$) the system is in normal phase and $\lambda =\omega _z=0.2$, in ($b_1$)-($b_2$) the system is in superradiant phase and $\lambda =2\omega _z=0.4$. Related parameters in plots: $\gamma =0$, $\gamma _c=0$, $r=0$, in ($a_1$)($b_1$) $\tau ^{-1}=\omega _z=0.2$; in ($a_2$) $\Omega _f=-\omega _z=-0.2$; in ($b_2$) $\Omega _f=-2.5\omega _z=-0.5$; the other parameters are the same as Fig. 3.

Download Full Size | PDF

In Fig. 5, we plot the dependence of the output steering and entanglement with the filtered output of frequency $\Omega _f=-\omega _z$ on the coupling $\lambda$. Similar features are shown to the intracavity case. For the same rates $\gamma$ and $\gamma _c$ and coupling $\lambda$, the steering from the atomic ensemble to the output field $S_{f|a}$ is larger than the steering reverse $S_{a|f}$. The steering $S_{f|a}$ is more sensitive to $\gamma$ and $\gamma _c$ than the steering $S_{a|f}$. When $\gamma$ and $\gamma _c$ increase from $0$ to $\omega _z$, $S_{f|a}$ decreases considerably and approaches to zero, but the steering $S_{a|f}$ drops lightly. This result implies that the direction of the steering can be controlled by the dissipation of atoms.

 figure: Fig. 5.

Fig. 5. The dependencies of the steering $S_{f|a}$ and $S_{a|f}$, entanglement $E_n$ between output field and spin field on $\lambda$ for different $\gamma$ in ($a_1$)-($a_3$), $\gamma _c$ in ($b_1$)-($b_3$), $r$ in ($c_1$)-($c_3$). Related parameters in plots: $\kappa =1.2$, $\Omega _f=-\omega _z$, $\tau ^{-1}=2\omega _z=0.4$. In ($a_1$)-($a_3$) $\gamma _c=0$, $r=0$; in ($b_1$)-($b_3$) $\gamma =0.1$, $r=0$; in ($c_1$)-($c_3$) $\gamma =\gamma _c=0.1$; the other parameters are the same as Fig. 2.

Download Full Size | PDF

6. Conclusion

In conclusion, we study steady-state quantum entanglement and steering in an open Dicke model where cavity dissipation and individual atomic decoherence are taken into account. Specifically, we consider that each atom is coupled to independent dephasing and squeezed environments, which makes the conventional Holstein-Primakoff approximation invalid. We find that in both of normal and superradiant phases, the cavity dissipation and individual atomic decoherence can improve the entanglement and steering between the cavity field and atomic ensemble. The individual atomic spontaneous emission leads to the appearance of the steering between the cavity field and atomic ensemble but the steerings in two directions cannot be simultaneously generated. In addition, the maximal achievable steering in normal phase is stronger than that in superradiant phase. The entanglement and steering between the cavity output field and atomic ensemble is much stronger than that with the intracavity. Our findings reveal unique features of quantum correlations in the open Dicke model in the presence of individual atomic decoherence processes. The nonclassical atom-field states can be used for quantum information processing, e.g., quantum teleportation channel for quantum-state transfer, and quantum metrology, such as the sensing of weak magnetic field.

Funding

National Natural Science Foundation of China (12174140, 62275075).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. R. H. Dicke, “Coherence in Spontaneous Radiation Processes,” Phys. Rev. 93(1), 99–110 (1954). [CrossRef]  

2. K. Hepp and E. H. Lieb, “On the superradiant phase transition for molecules in a quantized radiation field: the Dicke maser model,” Ann. Phys. 76(2), 360–404 (1973). [CrossRef]  

3. F. T. Hioe, “Phase Transitions in Some Generalized Dicke Models of Superradiance,” Phys. Rev. A 8(3), 1440–1445 (1973). [CrossRef]  

4. Y. K. Wang and F. T. Hioe, “Phase Transition in the Dicke Model of Superradiance,” Phys. Rev. A 7(3), 831–836 (1973). [CrossRef]  

5. P. Kirton, M. M. Roses, J. Keeling, and E. G. D. Torre, “Introduction to the Dicke Model: From Equilibrium to Nonequilibrium, and Vice Versa,” Adv. Quantum Technol. 2(1-2), 1800043 (2019). [CrossRef]  

6. K. Baumann, C. Guerlin, F. Brennecke, and T. Esslinger, “Dicke quantum phase transition with a superfluid gas in an optical cavity,” Nature 464(7293), 1301–1306 (2010). [CrossRef]  

7. Z. Zhang, C. H. Lee, R. Kumar, K. J. Arnold, S. J. Masson, A. S. Parkins, and M. D. Barrett, “Nonequilibrium phase transition in a spin-1 Dicke model,” Optica 4(4), 424 (2017). [CrossRef]  

8. Z. Zhang, C. H. Lee, R. Kumar, K. J. Arnold, S. J. Masson, A. L. Grimsmo, A. S. Parkins, and M. D. Barrett, “Dicke-model simulation via cavity-assisted Raman transitions,” Phys. Rev. A 97(4), 043858 (2018). [CrossRef]  

9. D. Schneble, Y. Torii, M. Boyd, E. W. Streed, D. E. Pritchard, and W. Ketterle, “The Onset of Matter-Wave Amplification in a Superradiant Bose-Einstein Condensate,” Science 300(5618), 475–478 (2003). [CrossRef]  

10. K. Baumann, R. Mottl, F. Brennecke, and T. Esslinger, “Exploring Symmetry Breaking at the Dicke Quantum Phase Transition,” Phys. Rev. Lett. 107(14), 140402 (2011). [CrossRef]  

11. D. Nagy, G. Kónya, G. Szirmai, and P. Domokos, “Dicke-Model Phase Transition in the Quantum Motion of a Bose-Einstein Condensate in an Optical Cavity,” Phys. Rev. Lett. 104(13), 130401 (2010). [CrossRef]  

12. M. P. Baden, K. J. Arnold, A. L. Grimsmo, S. Parkins, and M. D. Barrett, “Realization of the Dicke Model Using Cavity-Assisted Raman Transitions,” Phys. Rev. Lett. 113(2), 020408 (2014). [CrossRef]  

13. Y. Meng, C. Liedl, S. Pucher, A. Rauschenbeutel, and P. Schneeweiss, “Imaging and Localizing Individual Atoms Interfaced with a Nanophotonic Waveguide,” Phys. Rev. Lett. 125(5), 053603 (2020); T. Ramos, H. Pichler, A. J. Daley, and P. Zoller, “Quantum Spin Dimers from Chiral Dissipation in Cold-Atom Chains,” Phys. Rev. Lett. 113, 237203 (2014). [CrossRef]  

14. A. A. Houck, H. E. Türeci, and J. Koch, “On-chip quantum simulation with superconducting circuits,” Nat. Phys. 8(4), 292–299 (2012). [CrossRef]  

15. M. Feng, Y. P. Zhong, T. Liu, L. L. Yan, W. L. Yang, J. Twamley, and H. Wang, “Exploring the quantum critical behaviour in a driven Tavis–Cummings circuit,” Nat. Commun. 6(1), 7111 (2015). [CrossRef]  

16. H. Zhu, G. Zhang, and H. Fan, “Quantum Criticality in the Biased Dicke Model,” Sci. Rep. 6(1), 19751 (2016). [CrossRef]  

17. C. Emary and T. Brandes, “Chaos and the quantum phase transition in the Dicke model,” Phys. Rev. E 67(6), 066203 (2003). [CrossRef]  

18. A. Leviatan and M. Macek, “Evolution of order and chaos across a first-order quantum phase transition,” Phys. Lett. B 714(1), 110–114 (2012). [CrossRef]  

19. O. Castaños, R. López-Peña, E. Nahmad-Achar, and J. G. Hirsch, “Quantum information approach to the description of quantum phase transitions,” J. Phys.: Conf. Ser. 403, 012003 (2012). [CrossRef]  

20. R. R. Soldati, M. T. Mitchison, and G. T. Landi, “Multipartite quantum correlations in a two-mode Dicke model,” Phys. Rev. Lett. 104(5), 052423 (2021). [CrossRef]  

21. D. Z. Rossatto, D. P. Pires, F. M. de Paula, and O. P. de Sá Neto, “Quantum coherence and speed limit in the mean-field Dicke model of superradiance,” Phys. Rev. A 102(5), 053716 (2020). [CrossRef]  

22. G. Buonaiuto, F. Carollo, B. Olmos, and I. Lesanovsky, “Dynamical Phases and Quantum Correlations in an Emitter-Waveguide System with Feedback,” Phys. Rev. Lett. 127(13), 133601 (2021). [CrossRef]  

23. S. Jia, B. Li, Y. Jiao, H. Jing, and L. Kuang, “Einstein-Podolsky-Rosen steering of quantum phases in a cavity Bose-Einstein condensate with a single impurity,” arXiv, arXiv:2205.13938 (2022). [CrossRef]  

24. N. Lambert, C. Emary, and T. Brandes, “Entanglement and the Phase Transition in Single-Mode Superradiance,” Phys. Rev. Lett. 92(7), 073602 (2004). [CrossRef]  

25. J. Reslen, L. Quiroga, and N. F. Johnson, “Direct equivalence between quantum phase transition phenomena in radiation-matter and magnetic systems: Scaling of entanglement,” Europhys. Lett. 69(1), 8–14 (2005). [CrossRef]  

26. S. J. Masson and S. Parkins, “Extreme spin squeezing in the steady state of a generalized Dicke model,” Phys. Rev. A 99(2), 023822 (2019). [CrossRef]  

27. S. Morrison and A. S. Parkins, “Collective spin systems in dispersive optical cavity QED: Quantum phase transitions and entanglement,” Phys. Rev. A 77(4), 043810 (2008). [CrossRef]  

28. F. Dimer, B. Estienne, A. S. Parkins, and H. J. Carmichael, “Proposed realization of the Dicke-model quantum phase transition in an optical cavity QED system,” Phys. Rev. A 75(1), 013804 (2007). [CrossRef]  

29. M. Boneberg, I. Lesanovsky, and F. Carollo, “Quantum fluctuations and correlations in open quantum Dicke models,” Phys. Rev. A 106(1), 012212 (2022). [CrossRef]  

30. N. Pierre and C. Ciuti, “No-go theorem for superradiant quantum phase transitions in cavity QED and counter-example in circuit QED,” Nat. Commun. 1(1), 72 (2010). [CrossRef]  

31. D. A. Ivanov, T. Y. Ivanova, S. F. Caballero-Benitez, and I. B. Mekhov, “Feedback-Induced Quantum Phase Transitions Using Weak Measurements,” Phys. Rev. Lett. 124(1), 010603 (2020). [CrossRef]  

32. T. Holstein and H. Primakoff, “Field Dependence of the Intrinsic Domain Magnetization of a Ferromagnet,” Phys. Rev. 58(12), 1098–1113 (1940). [CrossRef]  

33. J. Gelhausen, M. Buchhold, and P. Strack, “Many-body quantum optics with decaying atomic spin states: (γ, κ) Dicke model,” Phys. Rev. A 95(6), 063824 (2017). [CrossRef]  

34. D. A. Paz and M. F. Maghrebi, “Driven-dissipative Ising model: An exact field-theoretical analysis,” Phys. Rev. A 104(2), 023713 (2021). [CrossRef]  

35. P. Kirton and J. Keeling, “Suppressing and Restoring the Dicke Superradiance Transition by Dephasing and Decay,” Phys. Rev. Lett. 118(12), 123602 (2017). [CrossRef]  

36. A. Einstein, B. Podolsky, and N. Rosen, “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?” Phys. Rev. 47(10), 777–780 (1935). [CrossRef]  

37. E. Schrödinger, “Discussion of Probability Relations between Separated Systems,” Math. Proc. Cambridge Philos. Soc. 31(4), 555–563 (1935). [CrossRef]  

38. H. M. Wiseman, S. J. Jones, and A. C. Doherty, “Steering, Entanglement, Nonlocality, and the Einstein-Podolsky-Rosen Paradox,” Phys. Rev. Lett. 98(14), 140402 (2007); S. J. Jones, H. M. Wiseman, and A. C. Doherty, “Entanglement, Einstein-Podolsky-Rosen correlations, Bell nonlocality, and steering,” Phys. Rev. A 76, 052116 (2007). [CrossRef]  

39. R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, “Quantum entanglement,” Rev. Mod. Phys. 81(2), 865–942 (2009). [CrossRef]  

40. J. S. Bell, “On The Einstein Podolsky Rosen Paradox,” Physics 1(3), 195–200 (1964). [CrossRef]  

41. C. Branciard, E. G. Cavalcanti, S. P. Walborn, V. Scarani, and H. M. Wiseman, “One-sided device-independent quantum key distribution: Security, feasibility, and the connection with steering,” Phys. Rev. A 85(1), 010301 (2012); N. Walk, S. Hosseini, J. Geng, O. Thearle, J. Y. Haw, S. Armstrong, S. M. Assad, J. Janousek, T. C. Ralph, T. Symul, H. M. Wiseman, and P. K. Lam, “Experimental demonstration of Gaussian protocols for one-sided device-independent quantum key distribution,” Optica 3, 634 (2016). [CrossRef]  

42. M. Piani and J. Watrous, “Necessary and Sufficient Quantum Information Characterization of Einstein-Podolsky-Rosen Steering,” Phys. Rev. Lett. 114(6), 060404 (2015). [CrossRef]  

43. Q. He, L. Rosales-Zárate, G. Adesso, and M. D. Reid, “Secure Continuous Variable Teleportation and Einstein-Podolsky-Rosen Steering,” Phys. Rev. Lett. 115(18), 180502 (2015). [CrossRef]  

44. F. Benatti, F. Carollo, R. Floreanini, and H. Narnhofer, “Quantum spin chain dissipative mean-field dynamics,” J. Phys. A: Math. Theor. 51(32), 325001 (2018). [CrossRef]  

45. F. Benatti, F. Carollo, R. Floreanini, and H. Narnhofer, “Non-markovian mesoscopic dissipative dynamics of open quantum spin chains,” Phys. Lett. A 380(3), 381–389 (2016). [CrossRef]  

46. A. Klein and E. R. Marshalek, “Boson realizations of Lie algebras with applications to nuclear physics,” Rev. Mod. Phys. 63(2), 375–558 (1991). [CrossRef]  

47. M. O. Scully and M. S. Zubairy, Quantum optics (Cambridge University Press, 1997).

48. G. Vidal and R. F. Werner, “Computable measure of entanglement,” Phys. Rev. A 65(3), 032314 (2002). [CrossRef]  

49. G. Adesso, A. Serafini, and F. Illuminati, “Extremal entanglement and mixedness in continuous variable systems,” Phys. Rev. A 70(2), 022318 (2004). [CrossRef]  

50. R. Simon, “Peres-Horodecki Separability Criterion for Continuous Variable Systems,” Phys. Rev. Lett. 84(12), 2726–2729 (2000). [CrossRef]  

51. I. Kogias, A. R. Lee, S. Ragy, and G. Adesso, “Quantification of Gaussian Quantum Steering,” Phys. Rev. Lett. 114(6), 060403 (2015). [CrossRef]  

52. C. W. Gardiner and P. Zoller, Quantum Noise (Springer, 2000).

53. H. Tan, W. Deng, Q. Wu, and G. Li, “Steady-state light-mechanical quantum steerable correlations in cavity optomechanics,” Phys. Rev. A 95(5), 053842 (2017). [CrossRef]  

54. C. Genes, A. Mari, P. Tombesi, and D. Vitali, “Robust entanglement of a micromechanical resonator with output optical fields,” Phys. Rev. A 78(3), 032316 (2008). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (5)

Fig. 1.
Fig. 1. Schematic plot of the system. An ensemble of identical two-level systems is collectively coupled to a cavity with loss rate $\kappa$. Each atom undergoes independent spontaneous and dephasing processes with rates $\gamma _k$ and $\gamma _c$ respectively. The output field is filtered to define temporal modes to study the quantum correlations between the ensemble and output field.
Fig. 2.
Fig. 2. Density plot of the steering $S_{f|a}$, $S_{a|f}$ and entanglement $E_n$ between intracavity field and spin field versus $\gamma$ and $\kappa /\omega _z$ in ($a_1$)-($a_3$), $\gamma _c$ and $\kappa /\omega _z$ in ($b_1$)-($b_3$), $r$ and $\kappa /\omega _z$ in ($c_1$)-($c_3$). Related parameters in plots: $\omega _0=\omega _z=0.2$, $\lambda =0.1$. In (a1)-(a3) $\gamma _c=r=0$; in ($b_1$)-($b_3$) $\gamma =0.1$, $r=0$; in ($c_1$) and ($c_3$) $\gamma =\gamma _c=0.1$; in ($c_2$) $\gamma =0.1$, $\gamma _c=1.5$.
Fig. 3.
Fig. 3. The dependencies of the steering $S_{f|a}$ and $S_{a|f}$, entanglement $E_n$ between intracavity field and spin field on $\lambda$ for different $\gamma$ in ($a_1$)-($a_3$), $\gamma _c$ in ($b_1$)-($b_3$), $r$ in ($c_1$)-($c_3$). Related parameters in plots: $\kappa =1.2$. In ($a_1$)-($a_3$) $\gamma _c=0$, $r=0$; in ($b_1$) and ($b_3$) $\gamma =0.1$, $r=0$; in ($b_2$) $\gamma =1.5$, $r=0$; in ($c_1$) and ($c_3$) $\gamma =\gamma _c=0.1$; in ($c_2$) $\gamma =\gamma _c=1.5$; the other parameters are the same as Fig. 2.
Fig. 4.
Fig. 4. The dependencies of the steering $S_{f|a}$ and $S_{a|f}$, entanglement $E_n$ between output field and spin field on $\Omega _f/\omega _z$ in ($a_1$), ($b_1$) and $\omega _z\tau$ in ($a_2$), ($b_2$). In ($a_1$)-($a_2$) the system is in normal phase and $\lambda =\omega _z=0.2$, in ($b_1$)-($b_2$) the system is in superradiant phase and $\lambda =2\omega _z=0.4$. Related parameters in plots: $\gamma =0$, $\gamma _c=0$, $r=0$, in ($a_1$)($b_1$) $\tau ^{-1}=\omega _z=0.2$; in ($a_2$) $\Omega _f=-\omega _z=-0.2$; in ($b_2$) $\Omega _f=-2.5\omega _z=-0.5$; the other parameters are the same as Fig. 3.
Fig. 5.
Fig. 5. The dependencies of the steering $S_{f|a}$ and $S_{a|f}$, entanglement $E_n$ between output field and spin field on $\lambda$ for different $\gamma$ in ($a_1$)-($a_3$), $\gamma _c$ in ($b_1$)-($b_3$), $r$ in ($c_1$)-($c_3$). Related parameters in plots: $\kappa =1.2$, $\Omega _f=-\omega _z$, $\tau ^{-1}=2\omega _z=0.4$. In ($a_1$)-($a_3$) $\gamma _c=0$, $r=0$; in ($b_1$)-($b_3$) $\gamma =0.1$, $r=0$; in ($c_1$)-($c_3$) $\gamma =\gamma _c=0.1$; the other parameters are the same as Fig. 2.

Equations (32)

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

H ^ = ω 0 a ^ f a ^ f + ω z S ^ z + λ N a ( a ^ f + a ^ f ) ( S ^ + + S ^ ) ,
d d t ρ ^ ( t ) = i [ H ^ , o ^ ] + κ L κ [ a ^ ] ρ ^ + k = 1 N a L ~ γ c [ σ ^ z k ] ρ ^     + k = 1 N a [ L γ k ( N t h + 1 ) [ σ ^ k ] ρ ^ + L γ k N t h [ σ ^ + k ] ρ ^     + L ~ γ k M [ σ ^ k ] ρ ^ + L ~ γ k M [ σ ^ + k ] ρ ^ ] ,
m ^ j = S ^ j N a ,   X ^ f = a ^ f + a ^ f 2 N a ,   Y ^ f = i ( a ^ f a ^ f ) 2 N a ,
d d t m ^ x = ω z m ^ y Γ x 2 m ^ x ,
d d t m ^ y = ω z m ^ x 2 2 λ X ^ f m ^ z Γ y 2 m ^ y ,
d d t m ^ z = 2 2 λ X ^ f m ^ y Γ z m ^ z γ 2 ,
d d t Y ^ f = ω 0 m ^ X + 2 2 λ m ^ x κ 2 Y ^ f ,
d d t X ^ f = ω 0 Y ^ f κ 2 X ^ f ,
λ c = ( 2 N t h + 1 ) ( κ 2 / 4 + ω 0 2 ) ( Γ x Γ y / 4 + ω z 2 ) 4 ω 0 ω z ,
m ^ x s = 0 , m ^ y s = 0 , m ^ z s = 1 2 ( 2 N t h + 1 ) , Y ^ f s = 0 , X ^ f s = 0 ,
m ^ x s = ± γ 2 Γ x g ω z λ c λ 1 λ c 2 λ 2 , m ^ y s = Γ x γ 8 g λ c λ 1 λ c 2 λ 2 , m ^ z s = λ c 2 2 ( 2 N t h + 1 ) λ 2 , Y ^ f s = ± 2 κ ω z λ c ( κ 2 / 4 + ω 0 2 ) γ Γ x g 1 λ c 2 λ 2 , X ^ f s = 2 ω z ω 0 λ c ( κ 2 / 4 + ω 0 2 ) γ Γ x g 1 λ c 2 λ 2 ,
q ^ a = m ~ ^ x | m ^ | ,     p ^ a = m ~ ^ y | m ^ | ,
[ q ^ a , p ^ a ] = i .
R θ φ = ( cos θ cos φ cos θ sin φ sin θ 0 0 sin φ cos φ 0 0 0 sin θ cos φ sin θ sin φ cos θ 0 0 0 0 0 1 0 0 0 0 0 1 ) ,
θ = arccos ( m z | m | ) ,
φ = { 2 π arccos ( m x | m | sin θ ) , m y < 0 , 0 , m y = 0 ,   m x = 0 , arccos ( m x | m | sin θ ) , o t h e r w i s e .
d d t u ( t ) = A u ( t ) + n i n ( t ) ,
A = ( Γ x 2 ω z 0 0 0 ω z Γ y 2 2 2 λ X ^ f 0 2 2 λ m ^ z 0 2 2 λ X ^ f Γ z 0 2 2 λ m ^ y 2 2 λ 0 0 κ 2 ω 0 0 0 0 ω 0 κ 2 ) ,
d d t V = A V + V A T + D ,
D = ( Γ x 4 0 m ^ x 2 γ 0 0 0 Γ y 4 m ^ y 2 γ 0 0 m ^ x 2 γ m ^ y 2 γ Γ z 2 + m ^ z γ 0 0 0 0 0 κ 2 0 0 0 0 0 κ 2 ) ,
V ~ = W R θ φ V R θ φ T W ,
W = ( 1 m 0 0 0 0 0 1 m 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 ) .
σ ~ = ( V ~ a V ~ c V ~ c T V ~ f ) ,
E n = max [ 0 , log ( 2 ν ~ ) ] ,
S a | f = max [ 0 , 1 S a | f 1 ] ,
S f | a = max [ 0 , 1 S f | a 1 ] ,
a ^ f o u t ( t ) = κ δ a ^ f ( t ) a ^ f i n ( t ) .
a ^ f o u t ( t ) = t f ( t s ) a ^ f o u t ( s ) d s ,
f ( t ) = 2 τ θ ( t ) e ( τ 1 + i Ω f ) ,
u ~ o u t ( ω ) = T ( ω ) [ M ~ ( ω ) + P o u t κ ] W R θ φ n i n ( ω ) .
T ( ω ) = ( 1 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 κ t 1 κ t 2 0 0 0 κ t 2 κ t 1 ) ,
V ~ o u t = d ω T ( ω ) [ M ~ ( ω ) + P o u t κ ] D ~ ( ω ) [ M ~ ( ω ) + P o u t κ ] T ( ω ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.