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

$\mathcal {PT}$ symmetry in a superconducting hybrid quantum system with longitudinal coupling

Open Access Open Access

Abstract

We propose a scheme consisting of coupled nanomechanical cantilever resonators and superconducting flux qubits to engineer a parity-time- ($\mathcal {PT}$-) symmetric phononic system formed by active and passive modes. The effective gain (loss) of the phonon mode is achieved by the longitudinal coupling of the resonator and the fast dissipative superconducting qubit with a blue-sideband driving (red-sideband driving). A $\mathcal {PT}$-symmetric to broken-$\mathcal {PT}$-symmetric phase transition can be observed in both balanced gain-to-loss and unbalanced gain-to-loss cases. Applying a resonant weak probe field to the dissipative resonator, we find that (i) for balanced gain and loss, the acoustic signal absorption to amplification can be tuned by changing the coupling strength between resonators; (ii) for unbalanced gain and loss, both acoustically induced transparency and anomalous dispersion can be observed around Δ = 0, where the maximum group delay is also located at this point. Our work provides an experimentally feasible scheme to design $\mathcal {PT}$-symmetric phononic systems and a powerful platform for controllable acoustic signal transmission in a hybrid quantum system.

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

1. Introduction

$\mathcal {PT}$-symmetric quantum mechanics initially proposed by Bender in 1998, which proved that the eigenvalues of non-Hermitian Hamiltonians with $\mathcal {PT}$ symmetry can be real, and attracted extensive attention [112]. These non-Hermitian Hamiltonians satisfy the $\mathcal {PT}$-symmetric condition via an appropriate balance of gain and loss [1319]. A real-to-complex spectrum ($\mathcal {PT}$-symmetric to broken $\mathcal {PT}$-symmetric phase) transition can be observed in these non-Hermitian Hamiltonians, when the $\mathcal {PT}$-symmetry condition is broken [2022]. The phase transition point is called the exceptional point (EP) [2326], where both the eigenvalues and the corresponding eigenstates of the system coalesce.

Despite the successful development of $\mathcal {PT}$-symmetric quantum mechanical theory, the experimental realization of $\mathcal {PT}$ symmetry quantum systems remains challenging. Optical systems provide fertile ground for theoretical and experimental studies of $\mathcal {PT}$ symmetry since the paraxial optical diffraction equation is equivalent to the quantum-mechanical Schrödinger equation [27]. The $\mathcal {PT}$ symmetry has been experimentally demonstrated in various optical systems, such as resonators, waveguides, and lattices [2833]. Based on the similarity of the optical and acoustic field, $\mathcal {PT}$ symmetry has also been studied theoretically and experimentally in acoustic or phonon regimes [3436]. However, experimental realizations of acoustic or phonon $\mathcal {PT}$ symmetry have mainly focused on optomechanical systems and are rarely performed in other systems [3739].

Superconducting quantum circuits (SQC) are ideal platforms for constructing various quantum systems by coupling with other systems [4043]. Superconducting qubits [44] with Josephson junction are easily connected with other quantum systems, such as microwave resonators, nanomechanical resonators, and spin ensembles [4553]. The coupling form of superconducting qubits with other quantum systems is flexible, including transverse couplings and longitudinal couplings [5465]. Taking advantage of these properties of SQC, some quantum systems with gain and loss are constructed by SQC to study $\mathcal {PT}$ symmetry and related non-Hermitian properties [6668]. However, the phononic $\mathcal {PT}$-symmetric quantum systems constructed by SQC longitudinal coupling are rarely studied.

In this work, we propose a hybrid quantum system composed of two superconducting qubits and two coupled nanomechanical resonators for studying $\mathcal {PT}$ symmetry and non-Hermitian properties in the phononic system. The active (passive) subsystem is achieved by the longitudinal coupling of the resonator and the fast dissipative qubit with a blue-sideband driving (red-sideband driving). By adjusting the coupling strength between two nanomechanical resonators, the transition from $\mathcal {PT}$-symmetric to broken $\mathcal {PT}$-symmetric phase can be observed. The acoustic signal absorption, amplification, and group delay can be observed in our $\mathcal {PT}$-symmetric hybrid quantum system. These findings enable new applications for controlling the transmission of acoustic signals in quantum information processes.

This paper is organized as follows. In section 2, we introduce the hybrid quantum system that we propose for the realization of $\mathcal {PT}$ symmetry. In section 3, we construct the gain and loss subsystem via different sideband drivings of qubits and derive the effective $\mathcal {PT}$-symmetric Hamiltonian for the system. In section 4, we study the phase transitions in the $\mathcal {PT}$-symmetric hybrid quantum system by controlling the coupling strength between two resonators. We discuss the acoustic signal transmission in section 5. The unique control of output fields, especially for acoustic signal amplification, absorption, and delay, is demonstrated. The summary is presented in section 6.

2. Setup

We consider a superconducting hybrid quantum system for studying $\mathcal {PT}$ symmetry as shown in Fig. 1(a). This system is composed of two coupled nanomechanical cantilever resonators and two superconducting flux qubits that are magnetically coupled to two cantilevers via an external magnetic tip. The magnetic tip is attached to the cantilever and produces a time-varying magnetic field. The motion of the cantilever can be described by the Hamiltonian $H_{ri} = \hbar \omega _{ri}b_{i}^{\dagger }b_{i}$ $(i=1,2)$, where $b_{i}^{\dagger }(b_{i})$ is the creation (annihilation) operator for the fundamental vibrational mode with frequency $\omega _{ri}$. The qubit is a gap-tunable flux qubit that is composed of a qubit loop and a superconducting quantum interference device (SQUID) [6971], for which the Hamiltonian is [72,73]

$$H_{qi} = \frac{1}{2}\epsilon_i(\Phi_{zi})\bar{\sigma}_{zi} + \Omega_i(\Phi_{xi})\bar{\sigma}_{xi},$$
where $\bar {\sigma }_{zi} = |+\rangle \langle +| - |-\rangle \langle -|$ and $\bar {\sigma }_{xi} = |+\rangle \langle -| + |-\rangle \langle +|$ are the Pauli operators in the basis of clockwise $|+\rangle$ and anti clock-wise $|-\rangle$ of the persistent currents states. Here, $\epsilon _i$ and $\Omega _i$ can be controlled independently by adjusting the external magnetic flux $\Phi _{zi}$ and $\Phi _{xi}$.

 figure: Fig. 1.

Fig. 1. (a) Schematic of a superconducting hybrid quantum system for studying $\mathcal {PT}$ symmetry. Each silicon nanomechanical cantilever resonator with dimensions $(l,w,t)$ is coupled with a flux qubit via an external magnetic tip attached to the cantilever. Two resonators separated by a distance $d$ are coupled indirectly through capacitive interactions with nearby wires. The inset shows the lumped-circuit layout of the flux qubit. (b) Circuit model that describes the coupling between two silicon nanomechanical cantilever resonators.

Download Full Size | PDF

The first term of $H_{qi}$ is the energy bias, which satisfies $\epsilon _i(\Phi _{zi}) = 2I_{pi}(\Phi _{zi} - \Phi _0/2)$. Here, $I_{pi}$ is the persistent current in the qubit loop, $\Phi _0=\hbar /2\pi$ is the flux quantum, and $\Phi _{zi}$ is the external flux of the qubit loop that can be produced by the flux driving. In our system, the qubit loop is inductively coupled to two identical microwave control lines symmetrically distributed on both sides of the qubit via a mutual kinetic inductance $M_i$ [74,75]. We add a magnetic flux threading through the qubit loop $\Phi _{\text {ext}}=\Phi _0$ to minimize the flux noise. Then we apply a time-dependent current $I_i(t)=\varepsilon _{pi} \text {cos}(\omega _{pi} t)$ in the microwave control line, where $\varepsilon _{pi}$ is the microwave current amplitude and $\omega _{pi}$ is the frequency of the microwave current. The first term of $H_{qi}$ can be re-expressed as $H_{qi} = \eta _i I_i(t)\bar {\sigma }_{z}$, where $\eta _i = M_iI_{pi}$ is the coupling strength between qubit loop and microwave control line [72].

The second term of $H_{qi}$ is the qubit gap $\Omega _i(\Phi _{xi})$ that dependents on the external flux $\Phi _{xi}$ threading the SQUID loop, which can be expressed as $\Phi _{xi} = \int _{Si}[B_{\text {total,i}}(x,y,z) + B_{\text {ext,i}}]dxdy \simeq \int _{Si}[\frac {\partial B_{\text {total},i,z}(x,y,z)}{\partial z}|_{z=z_{0}}(z-z_{0})+B_{\text {total},i,z}(x,y,z_{0}) + B_{\text {ext,i}}]dxdy$ [46,52]. Here, $B_{\text {total},i,z}(x,y,z)$ is the $z$ component of the total fields created by the magnetic tip of the $i$ th resonator, $B_{\text {ext,i}}$ is an applied external magnetic field, $S_i$ is the coupling area of the SQUID loop, $z_{0}$ is the average position of the SQUID loop, and $\frac {\partial B_{\text {total,i,z}}(x,y,z)}{\partial z}|_{z=z_{0}}$ is the magnetic field gradient at the position of the qubit [46,76]. The size of the magnetic tip we use here is $\sim 100 \ \text {nm}$, as reported in Ref. [77]. The magnetic gradient does not change very much in the area under the tip, which can be considered as a constant $G_{mi}$ in the effective coupling area. Assume the static part of $\Omega _i(\Phi _{xi})$ is $\hbar \omega _{qi}/2$, which is controlled by an applied external magnetic field. The qubit gap can be expressed as $\Omega _i(\Phi _{xi}) = \hbar \omega _{qi}/2 + I_{pi}G_{mi} S_{\text {eff,i}}b_{0i}(b_i^{\dagger } + b_i)\bar {\sigma }_{xi}$, where $b_{0i}(b_i^{\dagger } + b_i) = z -z_{0}$, and $b_{0i} = \sqrt {\hbar /2m\omega _{ri}}$ is the zero-point fluctuation amplitude of the mechanical resonator with mass $m_i$. We can find that the second term of $H_{qi}$ includes the coupling between the qubit and the nanomechanical cantilever, that is, $H_{\text {int}}=g_{i}(b_i^{\dagger } + b_i)\bar {\sigma }_{x}$, where the coupling strength $g_i =I_{pi}G_{mi} S_{\text {eff,i}}b_{0i}/\hbar$. In the new basis of the eigenstates of the qubit ${|e\rangle = (|+\rangle + |-\rangle )/\sqrt {2}}$ and ${|g\rangle = (|+\rangle - |-\rangle )/\sqrt {2}}$, the qubit Hamiltonian can be written as

$$H_{qi} = \frac{1}{2}\hbar \omega_{qi} \sigma_{zi} +\hbar g_i\sigma_{zi}(b_i^{\dagger} + b_i) +\hbar \eta_i \varepsilon_{pi} \text{cos}(\omega_{pi} t)\sigma_{xi},$$
where $\sigma _{zi} = |e\rangle \langle e| - |g\rangle \langle g|$ and $\sigma _{xi} = |e\rangle \langle g| + |g\rangle \langle e|$.

To achieve the interaction between two resonators, the nanomechanical cantilevers are charged and interact capacitively with nearby wires interconnecting them [78,79]. The circuit model describes the coupling between two resonators as shown in Fig. 1(b). The effective interaction between two resonators is introduced by the variations of the resonator-wire capacitance $C_i(z_i)$ with the position of the tip $z_i = b_{0}(b_i + b_i^{\dagger })$. The phonon-phonon coupling is due to the electrostatic energy $W_{\text {el}}$ of the underlying circuit, which can be calculated by $\hbar J_{ij} = b_{0}^2(\partial ^2 W_{\text {el}}/\partial z_i \partial z_j)|_{{z_i}=0}$ [78]. We assume two resonators separated by a distance $d$ and connected by a wire of self-capacitance $C_w \approx \epsilon _0 d$, where $\epsilon _0$ is the vacuum permittivity. Then we calculate the electrostatic energy $W_{\text {el}} = -(U^2/2)(C_{\sum }C_w/C_{\sum } + C_w)$, where $C_{\sum } = C_1(z_1) + C_2(z_2)$ and $U$ is the applied voltage. The coupling Hamiltonian between two resonators is

$$H_c = \hbar J(b_1 + b_1^{\dagger})(b_2 + b_2^{\dagger}),$$
where $J$ denotes the coupling strength and can be calculated by [78]
$$J = \frac{1}{\hbar} \frac{C^2 C_w^2 U^2}{(2C + C_w)^3}\frac{b_i^2}{h^2}.$$

Here, $J \equiv J_{12}\equiv J_{21}$ and $C_i(z_i) \simeq C(1 - z_i/h)$, where $h$ is the mean electrode spacing. The coupling strength $J$ between two resonators can be controlled by the applied voltage $U$.

After combing all parts, the Hamiltonian of the system can be written as (setting $\hbar =1$)

$$H_{\text{total}}= \sum_{i=1,2}H_{ri}+\sum_{i=1,2}H_{qi}+H_c.$$

Applying a frame rotating at a frequency $\omega _{pi}$ and adopting the rotating wave approximation, the Hamiltonian of the total system becomes [80]

$$\begin{aligned} \begin{aligned}H_{\text{total}}= & \sum_{i=1,2}\omega_{ri}b_i^{\dagger}b_i + \sum_{i=1,2}\frac{1}{2}\Delta_i\sigma_{zi}+\sum_{i=1,2}g_i\sigma_{zi}(b_i^{\dagger} + b_i)\\ & +\sum_{i=1,2}\eta_i\varepsilon_i(\sigma_{{+}i}+\sigma_{{-}i}) + J(b_1^{\dagger}b_2+b_2^{\dagger}b_1), \end{aligned} \end{aligned}$$
where $\Delta _i = \omega _{qi} - \omega _{pi}$ are the drive-excitation detunings.

Here, we estimate the parameters of this Hamiltonian under experimental conditions. We consider the cantilever resonator with dimensions $(l,w,t) = (1,0.05, 0.05) \mu \text {m}$, which corresponds a fundamental frequency is approximately $\omega = 2\pi \times 50 \text {MHz}$, and the zero-point fluctuation of the phonon mode is $b_0 = 1.4 \times 10^{-13} \text {m}$ [81]. For superconducting persistent current, the range of superconducting flux qubits is $I_{p} \approx 500 \ \text {nA}$ [56]. A magnetic tip with a magnetic field gradient $G_{m} \approx 8\times 10^6 \ \text {T/m}$ is experimentally realized, which corresponds to the effective coupling area $S_{\text {eff}} \approx 80 \ \text {nm} \times 80 \ \text {nm} \sim 0.64 \times 10^{-14} \ \text {m}^2$ [77]. The coupling strength $g \approx 2\pi \times 5 \ \text {MHz} \approx 0.1 \omega$, which reaches the strong coupling regimes. The coupling of superconducting qubits and mechanical cantilevers has been experimentally achieved, and the strength of the coupling can be tuned by changing the relative positions of the magnetic tip and superconducting loop [52]. Moreover, for two resonators separated by $d = 100$ $\mu m$, $h = 100$ $\text {nm}$, and a typical electrode capacitance $C =0.1$ $\text {fF}$, we obtain the coupling strength $J = 1$ $\text {MHz}$ between two resonators with an applied voltage $U = 1$ $\text {V}$ [78]. In this way, a superconducting hybrid system with adjustable properties can be realized.

3. Realization of $\mathcal {PT}$ symmetric system by different sideband drives

Parity-time symmetric systems consist of non-Hermitian subsystems with balanced gain and loss configurations. In this section, we show how to construct the gain and loss subsystem based on the above setup. Performing the polariton transformation [82] to the total Hamiltonian in Eq. (6)

$$H_1 = e^SH_{\text{total}}e^{{-}S},$$
with $S = \sum _{i=1,2}\beta _i\sigma _{zi}(b_i^{\dagger } - b_i)$ and $\beta _i = g_i/\omega _{r}$, we obtain
$$\begin{aligned}\begin{aligned} H_1 = & \sum_{i=1,2}(\frac{1}{2}\Delta_i\sigma_{zi}+\omega_{r}b_i^{\dagger}b_i )+\sum_{i=1,2}[\eta_i\varepsilon_i\sigma_{{+}i}e^{2\beta_i(b_i^{\dagger} - b_i)}+\text{H.c.}] \\ & -\beta_1\sigma_{z1}(b_1^{\dagger}+b_1)-\beta_2\sigma_{z2}(b_2^{\dagger}+b_2) +J(b_1^{\dagger}b_2+b_2^{\dagger}b_1), \end{aligned} \end{aligned}$$
where $\text {H.c.}$ denotes the Hermitian conjugate of the third term in $H_1$. Applying a unitary transformation $U = \text {exp}(-i H_0 t)$ with $H_0 =\sum _{i=1,2}(\frac {1}{2}\Delta _i\sigma _{zi}+\omega _{r}b_i^{\dagger }b_i)$, the system Hamiltonian can be expressed as
$$\begin{aligned} \begin{aligned} H_2 = & \sum_{i=1,2}2\eta_i \epsilon_i\beta_i[\sigma_{{+}i}b_i^{\dagger}e^{i(\Delta_i+\omega_r)t} + \sigma_{{-}i} b_ie^{{-}i(\Delta_i+\omega_r)t} +\sigma_{{+}i}b_ie^{i(\Delta_i-\omega_r)t} + \sigma_{{-}i} b_i^{\dagger}e^{{-}i(\Delta_i-\omega_r)t}] \\ & + \beta_1\sigma_{z1}(b_1^{\dagger}e^{i\omega_{r} t}+b_1e^{{-}i\omega_r t})+\beta_2\sigma_{z2}(b_2^{\dagger}e^{i\omega_{r} t}+b_2e^{{-}i\omega_r t})+J(b_1^{\dagger}b_2+b_2^{\dagger}b_1). \end{aligned} \end{aligned}$$

To build the gain and loss subsystems, we need different sideband drives for the qubits. The energy diagrams of two qubits with different sideband drives are shown in Fig. 2. For the subsystem that realizes the gain effect, the qubit drive satisfies the blue-sideband drive condition $\omega _{p1} =\omega _{q1} +\omega _r$. For the subsystem that implements the loss effect, the qubit drive satisfies the red-sideband drive condition $\omega _{p2} =\omega _{q2} -\omega _r$. After two qubits are driven with different sidebands driving, the Eq. (9) becomes

$$\begin{aligned}\begin{aligned} H_3 = & G_1(\sigma_{{+}1}e^{{-}i\omega_rt}+\sigma_{{-}1}e^{i\omega_rt})(b_1^{\dagger}e^{i\omega_r t}+b_1e^{{-}i\omega_r t})\\ & +G_2(\sigma_{{+}2}e^{i\omega_rt}+\sigma_{{-}2}e^{{-}i\omega_rt})(b_2^{\dagger}e^{i\omega_r t}+b_2e^{{-}i\omega_r t}) \\ & +\beta_1\sigma_{z1}(b_1^{\dagger}e^{i\omega_{r} t}+b_1e^{{-}i\omega_r t})+\beta_2\sigma_{z2}(b_2^{\dagger}e^{i\omega_{r} t}+b_2e^{{-}i\omega_r t}) \\ & +J(b_1^{\dagger}b_2+b_2^{\dagger}b_1). \end{aligned} \end{aligned}$$

For convenience, we denote the coupling strength of qubit and resonator as $G_1 = 2\eta _1 \epsilon _1 \beta _1$ and $G_2 = 2\eta _2 \epsilon _2 \beta _2$, respectively. In order to focus more clearly on the gain and loss effects of the two resonators, the qubit’s freedom needs to be removed.

 figure: Fig. 2.

Fig. 2. Energy diagram of blue-sideband driving qubit $1$ with frequency $\omega _{p1} =\omega _{q1} +\omega _r$ (a) and red-sideband driving qubit $2$ with frequency $\omega _{p2} =\omega _{q2} -\omega _r$(b).

Download Full Size | PDF

It is well known that mechanical modes decay to the thermal environment with finite temperature $T$, and the thermal phonon number at frequency $\omega _r$ can be calculated by $n_{th}=(e^{\hbar \omega _r/k_B T} -1)^{-1} \simeq k_B T/\hbar \omega _r$. Defining the Lindblad operator $D[A,\Omega ]\hat {\rho }=\frac {\Omega }{2}(A\hat {\rho }A^{\dagger }-A^{\dagger }A\hat {\rho })+\text {H.c}$, the system density matrix $\hat {\rho }(t)$ evolution follows the master equation

$$\begin{aligned} \frac{d \rho(t)}{dt}= & -i[H_3,\rho(t)]+\sum_{i=1,2} \Gamma_i D[\sigma_{{-}i}] + n_{th}\xi_i D[b_i^{\dagger}]\rho(t) + (n_{th}+1)\xi_i D[b_i]\rho(t). \end{aligned}$$

In the above equation, $\Gamma _i$ represents the energy relaxing rate of the qubit, and $\xi _i =\omega _r/Q_i$ is the dissipative rate for the mechanical resonator. As demonstrated in Appendix A, the excited states of the qubit for the gain subsystem can be adiabatically eliminated under the condition $2G_1^{2}/\Gamma _1 \ll \Gamma _1$, and the excited states of the qubit of the loss subsystem can be adiabatically eliminated under the condition $2G_2^{2}/\Gamma _2 \ll \Gamma _2$. Therefore, we obtain the evolution equation for the reduced density matrix $\mu _i$ of the two mechanical cantilevers as follows [46,83]

$$\begin{aligned}\begin{aligned} \frac{d\mu_1 (t)}{dt}= & D[b_1^{\dagger},\frac{2G_1^{2}}{\Gamma_1} + n_{th}\xi_1]\mu_1 +D[b_1,\frac{2G_1^{2}}{\Gamma_1 +2i\omega_r} + (n_{th}+1)\xi_1)]\mu_1 \\ \frac{d\mu_1 (t)}{dt}= & D[b_2^{\dagger},\frac{2G_2^{2}}{\Gamma_2+2i\omega_r} + n_{th}\xi_2]\mu_2 +D[b_2,\frac{2G_2^{2}}{\Gamma_2} + (n_{th}+1)\xi_2)]\mu_2. \end{aligned} \end{aligned}$$

Let $\bar {n}_i(t)= \text {Tr}\{\mu _i(t)b_i^{\dagger }b_i\}$ denotes the mean phonon number of each mechanical cantilever resonator, the classical equation which describes the evolution for $\bar {n}_i$ can be obtained [84]

$$\begin{aligned} \begin{aligned} & \frac{d\bar{n}_1}{dt}=(\frac{2G_1^2}{\Gamma_1}-\xi_1)\bar{n}_1 + (\frac{2G_1^2}{\Gamma_1}+n_{th}\xi_1),\\ & \frac{d\bar{n}_2}{dt}={-} (\frac{2G_2^2}{\Gamma_2}+\xi_2)\bar{n}_2+n_{th}\xi_2. \end{aligned} \end{aligned}$$

It is easy to find that the photon number of resonator 1 will increase with $2G_1^2/\Gamma _1-\xi _1$, so there exists a gain effect. The phonon number of resonator 2 will decrease with $2G_2^2/\Gamma _2+\xi _2$, so there exists a loss effect. The equation of phonon number in the mechanical cantilevers versus time is calculated as:

$$\begin{aligned} \begin{aligned} & \langle\bar{n}_1(t)\rangle = \langle\bar{n}_1(0)+1\rangle e^{2G_1^2 t/\Gamma_1} - 1,\\ & \langle\bar{n}_2(t)\rangle = \langle\bar{n}_2(0)\rangle e^{{-}2G_2^2 t/\Gamma_2}. \end{aligned} \end{aligned}$$

Here we assume $2G_i^2/\Gamma _i \gg \xi _i$, that is, the coupling rate between the qubit and the resonator is much greater than the loss of the resonator. It should be noted that there exists a gain limit for resonator 1 and a loss limit for resonator 2. When the whole system reaches its steady state, which requires $d\bar {n}_1/{dt}=0$ and $d\bar {n}_2/{dt} =0$, so we have $\bar {n}_1 = -(n_{th}\xi _1\Gamma _1)/2G_1^2 -1$ and $\bar {n}_2=(n_{th}\xi _2\Gamma _2)/2G_2^2$. In the regime, $\Gamma _i \gg 2 G_i^2/\Gamma _i$, with the very large dissipative rate of qubits, the coherent transfer of phonons between the cantilever and the qubit will be destroyed due to the qubit energy decay process with the environment. The gain rate $2G_1^2/\Gamma _1$ and the loss rate $2G_2^2/\Gamma _2$ will decrease with $\Gamma _1$ and $\Gamma _2$. This means that the very strong energy dissipative rate of the qubit will destroy the gaining or losing process.

We numerically simulate the phonon number of two mechanical cantilevers by using the original Hamiltonian in Eq. (6) and compare it with the analytical results in Eq. (14). The mean phonon number of two nanomechanical reasoners versus time $gt/2\pi$ is plotted in Fig. 3(a) or 3(b), respectively. As the evolution time increases, the average phonon number of resonator 1 increases gradually, and the average phonon number of resonator 2 decreases gradually. It also can be found that as the evolution time increases, the numerical solution and the analytical solution do not fit well. This is because many high-frequency oscillatory terms are removed in solving the effective Hamiltonian. We then give the analytical and numerical solution error analysis of the gain system and the dissipative system, as shown in Fig. (3c) or (3d), respectively. When the coupling strength $g_i$ and the driving field strength $\epsilon _i$ are increased simultaneously, the numerical and analytical solutions’ errors become more considerable. The red dashed line indicates that the error between the analytical and numerical solutions of the gain or loss subsystem is $0.2$. To keep the error below $0.2$, the values of $g_i$ and $\epsilon _i$ should be in the region of the left half of the red dashed line in Fig. (3c) or (3d). In our following analysis, we all keep the error of our hybrid system below $0.2$. This error control is relatively easy to achieve in our hybrid quantum system due to the longitudinal coupling. The longitudinal coupling between the qubit and the resonant cavity will not produce a Lamb shift to the qubit, and there is no dispersive coupling between them, so the lifetime of the qubit does not depend on the resonator. In our system, we only need to consider the dissipation rate of the qubit, not the effect of the resonator on the qubit. After adiabatically eliminating the qubit’s degree of freedom, the system can be treated as two coupled harmonic resonators with the same frequency $\omega _r$. The effective non-Hermitian Hamiltonian of this hybrid quantum system can be written as

$$H = (\omega_r + i\frac{\gamma}{2})b_1^{\dagger}b_1 + (\omega_r - i\frac{\kappa}{2})b_1^{\dagger}b_1 + J(b_1^{\dagger}b_2 + b_2^{\dagger}b_1),$$
where $\gamma = 2G_1^2/\Gamma _1$ and $\kappa = 2G_2^2/\Gamma _2$ represent the gain and dissipation of the resonator, respectively. The passive mode and active mode are coupled, forming a $\mathcal {PT}$-symmetric physical system. The effective coupling strength $J$ between passive mode and active mode can be adjusted by tuning the applied voltage $U$. Such controllable coupling parameter $J$ can be used to generate phase transitions in our $\mathcal {PT}$-symmetric hybrid quantum system, which will be discussed in detail in the following section.

 figure: Fig. 3.

Fig. 3. The mean phonon number of the nanomechanical cantilever resonator as a function of time $gt/2\pi$ in the gain subsystem (a) and loss subsystem (b). The analytical and numerical solution error analysis of the gain (c) and dissipation subsystem (d). Here, we choose $\omega _r = 10$ MHz, $\Gamma _i = 4$ MHz, and $G_i= 0.05$ MHz (corresponding to $\eta _i = 2$, $\epsilon _i = 0.5$ MHz, and $g_i = 0.5$ MHz) (i=1,2).

Download Full Size | PDF

4. Controllable phase transition

It is well known that a $\mathcal {PT}$-symmetric Hamiltonian may exhibit a real spectrum for some certain parameter regimes. The (phase) transition from a real spectrum ($\mathcal {PT}$-symmetric phase) to a complex-valued spectrum (broken $\mathcal {PT}$-symmetric phase) occurs at the phase transition point. The eigenvalues and corresponding eigenvectors of the Hamiltonian merge simultaneously at the phase transition point, and many interesting and counterintuitive phenomena can be observed. In the following, we discuss how phase transition can take place in our hybrid quantum system.

According to Eq. (15), the coupling between two resonators leads two supermodes $B_+ = (b_1+b_2)/\sqrt {2}$ and $B_{-} = (b_1-b_2)/\sqrt {2}$, with the eigenfrequencies

$$\omega_{{\pm}} = \omega_r + \frac{i}{4}(\gamma - \kappa) \pm \sqrt{J^2 - \frac{(\gamma + \kappa)}{16}}.$$

The real and imaginary parts of $\omega _{\pm }$ quantify the frequencies and linewidth of these two supermodes, and the square root in the expression measures the competition between the coupling strength and the loss and gain of the resonators.

We first consider that the gain and loss are strictly balanced with $\gamma = \kappa$, as shown in the black-solid and red-dashed curves in Figs. 4(a) and 4(b). When $J > \kappa /2$, the coupling of two mechanical resonators produces two separate supermodes with different frequencies $\omega _r \pm \sqrt {J^2 - \kappa /8}$). However, these two supermodes have the same linewidth $(\gamma -\kappa )/4 = 0$. This regime corresponds to the $\mathcal {PT}$-symmetric regime, which is the blue area in Figs. 4(a) and 4(b). In the $\mathcal {PT}$-symmetric regime, since the coupling strength is large enough, the energy in the gain resonator can be transferred to the dissipated resonator in time, and the system can be in equilibrium. When $J < \kappa /2$, these two supermodes are degenerate with the same frequency $\omega _r$ but with different linewidths $\pm \sqrt {\kappa ^2/4-J^2}$. This regime corresponds to the broken $\mathcal {PT}$-symmetric regime, which is the pink area in Figs. 4(a) and 4(b). In broken $\mathcal {PT}$-symmetric regime, the energy in the gain resonator increases exponentially, while the energy in the dissipative resonator decreases exponentially. However, when the coupling strength is small, the gain is not sufficient to quickly compensate for the dissipation. At the critical point $J = \kappa /2$, the resonance frequencies and the linewidths of these two supermodes become degenerate. This critical point is the EP of our hybrid system, also known as the $\mathcal {PT}$ phase transition point. As shown in Figs. 4(a) and 4(b), by increasing the effective coupling strength $J$, the transition from the broken to the unbroken $\mathcal {PT}$-symmetric phase can be achieved. Unlike the $\mathcal {PT}$-symmetric system with gain and loss, when two mechanical modes have the same loss $\kappa$, the phase transition point does not exist, as shown in the blue-solid and the green-dashed curves in Figs. 4(a) and 4(b). The frequencies of these two supermodes $\omega _r \pm \sqrt {J^2}$ are different and the frequency separation between these two supermodes is proportional to the coupling strength $J$. These two supermodes always have identical linewidth $\kappa /2$.

 figure: Fig. 4.

Fig. 4. $\mathcal {PT}$ phase transition in our hybrid quantum system. The real (a,c) and imaginary (b,d) parts of the eigenfrequencies $\omega _{\pm }$ plotted as a function of coupling strength $J$. The results for the case of balanced gain and loss are shown in (a) and (b), and for the case of the unbalanced gain-to-loss are plotted in (c) and (d). The black-solid and red-dashed curves in (a) and (b) represent the system gain and loss balance, while the blue-solid and green-dashed curves represent the system without mechanical gain. The black-solid and red-dashed curves in (c) and (d) are calculated with $\gamma = \kappa /2$, and the blue-solid and green-dashed are calculated with $\gamma = 2\kappa$. Broken and unbroken $\mathcal {PT}$-symmetric regimes are distinguished by different background colors.

Download Full Size | PDF

We next consider more general cases in that gain and loss are not strictly balanced. Figures 4(c) and 4(d) show two examples with $\gamma = \kappa /2$ and $\gamma = 2\kappa$. It can be found that the gain and loss ratios are different, and the critical point of the $\mathcal {PT}$ phase transition of the system is different. The smaller the ratio of gain to loss, the greater coupling strength $J$ that needs to induce the $\mathcal {PT}$ phase change. This can be understood as when the dissipation is large, the small coupling strength cannot transfer the energy in the gain system to the dissipation system in time, so the system cannot be in a balanced state. Therefore, the greater the dissipation $\kappa$, the greater coupling strength $J$ is required to reach a balanced state.

5. Controllable acoustic output field in $\mathcal {PT}$-symmetric hybrid quantum system

Controllable acoustic signal transmission and detection play an important role in quantum information processing [8,85,86]. The $\mathcal {PT}$-symmetric system consisting non-Hermitian subsystems with balanced gain and loss configurations, can be used for signal detection and manipulation. In this section, we will study the output field of our $\mathcal {PT}$-symmetric hybrid quantum system. In particular, we will show that the $\mathcal {PT}$ symmetry parameters affect the transmission properties for both balanced gain-to-loss and unbalanced gain-to-loss cases.

Here, we assume that the detection field is applied to the dissipative subsystem, the system Hamiltonian can be expressed as

$$\begin{aligned} H = & (\omega_r + i\frac{\gamma}{2})b_1^{\dagger}b_1 + (\omega_r - i\frac{\kappa}{2})b_2^{\dagger}b_2 + J(b_1^{\dagger}b_2 + b_2^{\dagger}b_1) + \varepsilon_p(b_2 e^{i\omega_p t} + b_2^{\dagger}e^{{-}i\omega_p t}), \end{aligned}$$
where $\omega _p = \omega _r$ is the frequency of the probe field and $\varepsilon _p$ is the amplitude of the detection field. Defining the input field of the dissipative resonator as $b_{\text {in2}}$, we can write the following dynamic equations
$$\begin{aligned} \begin{aligned} & \frac{d b_1}{d t}={-}i\omega_r b_1 + \frac{\gamma}{2}b_1 -i J b_2\\ & \frac{d b_2}{d t}={-}i\omega_r b_2 - \frac{\kappa}{2}b_2-i J b_1-i\varepsilon_p e^{{-}i\omega_r t} \sqrt{\kappa^{\prime}}b_{\text{in2}}, \end{aligned} \end{aligned}$$
together with the input-output relations $b_{\text {out2}} = b_{\text {in2}} + \sqrt {\kappa ^{\prime }}b_2$. Here $\kappa ^{\prime }$ denotes the losses due to the coupling of the resonators with the driving field. Substituting $b_{k} = B_k e^{-i\omega t}$ and $db_k/d t=-i\omega B_k e^{-i\omega t}+dB_k/d t e^{-i\omega t}$ in Eq. (18), we get
$$\begin{aligned} \begin{aligned} & \frac{d B_1}{d t}={-}i(\omega-\omega_r)B_1 + \frac{\gamma}{2}B_1-i J B_2\\ & \frac{dB_2}{d t}={-}i(\omega-\omega_r)B_2 - \frac{\kappa}{2}B_2-i J B_1-i\varepsilon_p e^{{-}i\omega_r t}-\sqrt{\kappa^{\prime}}B_{\text{in2}}. \end{aligned} \end{aligned}$$

Solving Eq. (19) in the steady-state, we obtain the intracavity fields $B_1$ and $B_2$ as

$$\begin{aligned}\begin{aligned} & B_1 = \frac{i J\sqrt{\kappa^{\prime}}}{J^2+(i\Delta-\kappa/2)(i\Delta+\gamma/2)}B_{\text{in2}} \\ & B_2 = \frac{\sqrt{\kappa^{\prime}}(i\Delta+\gamma/2)}{J^2+(i\Delta-\kappa/2)(i\Delta+\gamma/2)}B_{\text{in2}}, \end{aligned} \end{aligned}$$
where $\Delta = \omega -\omega _r$ is the detuning between the probe frequency and the field frequency. According to the input-output relations $B_{\text {out2}} = B_{\text {in2}} + \sqrt {\kappa ^{\prime }}B_2$, one can easily get the transmittance $T$ of the system
$$\begin{aligned} T = \vert\frac{B_{\text{out2}}}{B_{\text{in2}}}\vert^2 =|1+\frac{\kappa^{\prime}(i\Delta+ \gamma/2)}{J^2+(i\Delta-\kappa/2)(i\Delta+\gamma/2)}|^2. \end{aligned}$$

Using this transmission coefficient, we discuss below how to achieve controllable acoustic signal amplification, absorption, and group delay in the $\mathcal {PT}$-symmetric hybrid quantum system.

5.1 Balanced gain-to-loss case

We first consider the case where the gain and loss of the system are balanced. The relationship between the transmission coefficient $T$ and the detuning $\Delta$ with the same coupling strength $J = 0.425\gamma$ is calculated and shown in Fig. 5(a). For the loss-loss system (represented by the black-dotted line), an Autler-Townes-splitting-like spectrum [8794]is observed. However, for our $\mathcal {PT}$-symmetric system with gain and loss (represented by the red-solid line), signal amplification can be realized between two Autler-Townes absorption dips, where the peak is located at $\Delta = 0$. The dispersion behavior for both cases is calculated and shown in Fig. 5(b). Around the parameter regime $\Delta = 0$, both the gain-loss and loss-loss system exhibit a normal dispersion. Our $\mathcal {PT}$-symmetric gain-loss system has a larger dispersion change at $\Delta = 0$.

 figure: Fig. 5.

Fig. 5. Transmission coefficient $T$ and the dispersion Im($t_p$) versus detuning $\Delta$ with $J/\gamma = 0.425$ are shown in (a) and (b), respectively. The transmission coefficient $T$ as a function of detuning $\Delta$ for different coupling strengths $J$ of the gain-loss system is plotted in (c) and (d), and the loss-loss system is plotted in (e) and (f). The other system parameters are $\gamma = 0.4 \text {MHz}$, $\kappa = 0.4 \text {MHz}$, and $\kappa ^{\prime } = 0.2 \text {MHz}$.

Download Full Size | PDF

We next investigate how the transmission coefficient $T$ is affected by the coupling strength $J$ in the gain-loss system and the loss-loss system, respectively. The transmission coefficient $T$ versus detuning $\Delta$ for different coupling strength $J$ is calculated and shown in Figs. 5(c) and 5(d). In the broken $\mathcal {PT}$-symmetry scheme, as shown in Fig. 5(c), the transmission spectrum varies with the coupling strength $J$. When $J$ is small, the transmission spectrum appears as an absorption valley at $\Delta =0$. With the increase of the coupling strength $J$, the transmission coefficient $T$ is greater than $1$ at $\Delta =0$. When the coupling strength $J$ is close to the phase transition point, the transmission coefficient $T$ is much larger than $1$, and the amplification of the acoustic signal is realized. When the coupling strength $J$ exceeds the phase transition point, the transmission coefficient $T$ appears as two amplified peaks, as shown in Fig. 5(d). This is because beyond the phase transition point, the $\mathcal {PT}$-symmetric hybrid quantum system behaves as two supermodes, and there are two amplified peaks in the transmission spectrum.

In contrast to the $\mathcal {PT}$-symmetric system, the transmission coefficient $T$ is always small than $1$ in the loss-loss system without gain, as shown in Figs. 5(e) and 5(f). It can also be found that the transmission spectrum appears as an absorption valley when the coupling strength is weak. As the coupling strength $J$ increases, the transmission spectrum appears as two absorption valleys. According to the previous discussion, there is no phase transition point in a system without gain. Therefore, by changing the coupling strength $J$, the transmission spectrum should maintain the mode splitting. However, when the coupling strength is small, the energy exchange between the two dissipative systems is small, so the transmission spectrum appears as an absorption valley. Mode splitting becomes apparent when the coupling strength is increased. To explore the influence of coupling strength $J$ on signal absorption and amplification, we take the logarithm of the transmission coefficient $T$ at $\Delta =0$. According to Eq. (21), the transmission coefficient $T$ at $\Delta =0$ is given by

$$\begin{aligned} T =|1+\frac{\kappa^{\prime}\gamma/2}{J^2-\kappa\gamma/4}|^2, \end{aligned}$$
which indicates a strong amplification near the phase transition point $J = \sqrt {\kappa \gamma /4}$. Figure 6 shows the logarithm of the transmission coefficient $T$ versus $J$. For the $\mathcal {PT}$-symmetric system with gain and loss, as $J$ increases, the signal transmission gradually changes from absorption to amplification. Strong signal amplification can be achieved near the phase transition point. Additionally, a perfect signal absorption can be realized in the broken $\mathcal {PT}$-symmetric regime. In contrast, there is no signal amplification occurs in a system without gain. When the coupling strength is small, an obvious signal absorption can be observed. As $J$ increases, the logarithm of the transmission coefficient $T$ is approximately a straight line near zero. Therefore, in our $\mathcal {PT}$-symmetric hybrid quantum system, the absorption, and amplification of the acoustic signal can be achieved by changing the coupling strength $J$.

 figure: Fig. 6.

Fig. 6. The logarithm of the transmission coefficient $T$ versus coupling strength $J$ for detuning $\Delta = 0$. The black solid line represents the case in which gain and loss are balanced. The red solid line indicates the case where there is no gain in the system.

Download Full Size | PDF

5.2 Unbalanced gain-to-loss case

Here we focus on the case where the gain and loss are not equal. The transmission coefficient $T$ versus detuning $\Delta$ is calculated and plotted in Fig. 7(a), where the coupling strength $J = 11\gamma$. For our $\mathcal {PT}$-symmetric hybrid quantum system (the blue-solid-curve), a remarkable acoustic signal amplification is observed around $\Delta = 0$. However, for the system without gain (red-dot-dashed curve), one can obtain an acoustically induced transparency spectrum [49,9597]. The inset of Fig. 7(a) shows the center peak. The dispersion behavior for a $\mathcal {PT}$-symmetric system with gain and loss and a system without gain are shown in Fig. 7(b), where we find that both of them exhibit anomalous dispersion. However, we noticed an abrupt variation of the dispersion in $\mathcal {PT}$-symmetric system with gain and loss (see the blue solid curve in Fig. 7(b)), which is due to the enhancement of the acoustic transmission. The center dispersion variation is plotted in the inset of Fig. 7(b).

 figure: Fig. 7.

Fig. 7. The transmission coefficient $T$ and dispersion Im($t_p$) versus the detuning $\Delta$ with $J = 11\gamma$ are plotted in (a) and (b), respectively. The blue-solid curve (red-dot-solid curve) represents the gain-loss system (loss-loss system). Here, we set $\gamma = 0.1\text {MHz}$, $\kappa = 400 \gamma$, and $\kappa ^{\prime } = 1/2 \kappa$.

Download Full Size | PDF

It is well known that the group delay $\tau$ is caused by the extremely abrupt change of the refractive index. The group delay $\tau$ of our transmission field is defined as [98]

$$\tau = \frac{d\theta}{d\Delta},$$
where $\theta = \text {arg}[t_p(\Delta )]$ is the phase of the output field at the detuning $\Delta$. The phase $\theta$ and group delay $\tau$ for the $\mathcal {PT}$-symmetric system with gain and loss and the system without gain are plotted in Figs. 8(a) and 8(b), where the coupling strength is assumed as $J = 11\gamma$. It can be found that a tiny gain leads to a faster variation of the phase (see the green-solid curve in Fig. 8(a)). For the system without gain, the maximum group delay is $0.3$ $\mu s$ (see the orange-dashed curve in Fig. 8(b)), which can be prolonged to $8$ $\mu s$ with gain (see the green-solid curve in Fig. 8(b)). This offers a new way to control acoustic signal transmission beyond what is possible in conventional quantum systems.

 figure: Fig. 8.

Fig. 8. Phase $\theta$ and group delay $\tau$ for the $\mathcal {PT}$-symmetric system with gain and loss (see the green-solid curve) and the system without gain (see the orange-dashed curve). The parameters are the same as Fig. 7.

Download Full Size | PDF

In addition, the acoustic signal absorption and amplification can also be achieved by changing the coupling strength $J$ in our unbalanced gain-to-loss case. The logarithm of the transmission coefficient $T$ varies with the coupling strength $J$ in the unbalanced gain-to-loss case is the same as in the balanced gain-to-loss case (shown in Fig. 6). When the gain and loss are unbalanced, the signal is also amplified most at the phase transition point in the PT-symmetric system with gain and loss, and no signal is amplified in the system without gain. It means that in our $\mathcal {PT}$ symmetric hybrid quantum system, even if gain and loss are not equal, the absorption and amplification of the acoustic signal can be realized by changing the coupling strength $J$.

6. Conclusion

In conclusion, we have proposed a hybrid quantum system consisting of two superconducting qubits and two coupled nanomechanical cantilever resonators. We show that a $\mathcal {PT}$-symmetric phononic system can be constructed by active and passive modes. Effective gain or loss of the phonon mode is introduced by longitudinal coupling of the resonator and the fast dissipative qubit with a blue-sideband or red-sideband driving, respectively. A $\mathcal {PT}$-symmetric to broken $\mathcal {PT}$-symmetric phase transition can be achieved by changing the coupling strength $J$ between resonators for both balanced and unbalanced gain-to-loss cases. When a resonant weak probe field is added to the dissipative resonator, the transition of the acoustic signal from absorption to amplification can be achieved in the balanced gain-loss case. For the case of unbalanced gain and loss, both acoustically induced transparency and anomalous dispersion can be observed around $\Delta = 0$, which results in an ultralong group delay. This hybrid quantum system for constructing a $\mathcal {PT}$-symmetric phononic system can be experimentally realized, which provides a good candidate for the controllable transmission of acoustic signals.

Appendix

Here we use a single resonator to characterize the adiabatic elimination process [46]. Taking the gain subsystem as an example, the subsystem Hamiltonian is

$$\begin{aligned} H_{\text{gain}} = G_1(\sigma_{{+}1}e^{{-}i\omega_rt}+\sigma_{{-}1}e^{i\omega_rt})(b_1^{\dagger}e^{i\omega_r t}+b_1e^{{-}i\omega_r t}). \end{aligned}$$

We suppose that the subsystem density matrix is $\rho (t)= \mu (t)\otimes \rho _q(t)\otimes \rho _B$, where $\mu (t)$, $\rho _q(t)$, $\rho _B$ represent the reduced density matrix of the mechanical modes, qubit, and thermal environment, respectively. Here we do not consider the interaction between the mechanical modes and the heat reservoir. The interaction Hamiltonian between the qubit and the thermal environment can be expressed as $V(t) =\sum _{k}\Gamma _k \sigma _+a_k e^{-i(\omega _k -\omega _q)t} +\text {H.c.}$, where $a_k$ is the annihilation operator for the mode $k$ with frequency $\omega _k$ of the environment. Thus the Hamiltonian for the gain subsystem is given by

$$H_{A1} = H_{\text{gain}} + V(t).$$

We assume that the coupling between the qubit and the vibration mode with a weak coupling strength $G_1$ while the qubit interacts with the thermal environment with a strong coupling rate $\Gamma _1$, then the qubit is dominated by the dissipative terms and approximately in its ground state at all times. Performing a transformation $\hat {U}(t) = \text {exp}[-i\int _0^{t}V(t_1)dt_1]$ to Eq. (25), the Hamiltonian becomes

$$\begin{aligned} H_{A2} = G_1 \hat{U}(t)^{\dagger}(\sigma_{{+}1}e^{{-}i\omega_rt}+\sigma_{{-}1}e^{i\omega_rt})\hat{U}(t)(b_1^{\dagger}e^{i\omega_r t}+b_1e^{{-}i\omega_r t}), \end{aligned}$$
and the master equation for the resonator reduced density matrix $\mu _1 (t)$ is given by
$$\begin{aligned} \frac{d\mu_1 (t)}{dt}={-}\int_0^{t}dt^{\prime}\text{tr}_q \text{tr}_B[H_{A2}(t),[H_{A2}(t^{\prime}),\mu (t^{\prime})\otimes \rho_q(t^{\prime})\otimes \rho_B]]. \end{aligned}$$

Here we have assumed that

$$\begin{aligned} \text{tr}_q \text{tr}_B[H_{A2}(t),\hat{\rho}(0)]=0, \end{aligned}$$
where $\text {tr}_q$ and $\text {tr}_B$ mean tracing over the qubit and the thermal environment. The Markov approximation means that $\mu (t^{\prime })$ does not change greatly, so the evolution time $\tau _{\mu } \gg t$. Thus we can replace $\mu (t^{\prime })$ with $\mu (t)$,
$$\begin{aligned} \begin{aligned} \frac{d\mu_1 (t)}{dt} & ={-}\int_0^{t} dt^{\prime}\text{tr}_q \text{tr}_B[H_{A2}(t),[H_{A2}(t^{\prime}),\mu (t)\otimes \rho_q(t^{\prime})\otimes \rho_B]]\\ & =\int_0^{t} dt^{\prime}\text{tr}_q \text{tr}_B[H_{A2}(t^{\prime})\mu(t) \otimes \rho_q(t^{\prime}) \otimes \rho_B H_{A2}(t) \\ & - H_{A2}(t)H_{A2}(t^{\prime})\mu(t) \otimes \rho_q(t^{\prime})\otimes \rho_B]+\text{H.c.}, \end{aligned} \end{aligned}$$

We reexpress $H_{A2}$ as $H_{A2}= C(t)(b_1^{\dagger }e^{i\omega _r t}+b_1e^{-i\omega _r t})$, where $C(t) = G_1 (\sigma _{+1}(t)e^{-i\omega _rt}+\sigma _{-1}(t)e^{i\omega _rt})$, and $\sigma _{+1}(t)=\hat {U}(t)^{\dagger }\sigma _{+1}\hat {U}(t)$, $\sigma _{-1}(t)=\hat {U}(t)^{\dagger }\sigma _{-1}\hat {U}(t)$. Neglecting the oscillating terms in Eq. (29) and defining $t^{\prime }= t-s$, we obtain

$$\begin{aligned} \begin{aligned} \frac{d\mu_1 (t)}{dt}= & \int_0^{t} ds \text{tr}_q \text{tr}_B [C(t-s)\rho_q(t-s)\otimes \rho_B C(t)] e^{i\omega_r s} b_1^{\dagger} \mu(t)b_1 \\ & + \int_0^{t} ds \text{tr}_q \text{tr}_B [C(t-s)\rho_q(t-s)\otimes \rho_B C(t)]e^{{-}i\omega_r s}b_1\mu(t)b_1^{\dagger} \\ & -\int_0^{t} ds \text{tr}_q \text{tr}_B [C(t)C(t-s)\rho_q(t-s)\otimes \rho_B]e^{{-}i\omega_r s}b_1^{\dagger}b_1 \mu(t)\\ & -\int_0^{t} ds \text{tr}_q \text{tr}_B [C(t)C(t-s)\rho_q(t-s)\otimes \rho_B]e^{i\omega_r s}b_1b_1^{\dagger} \mu(t) +\text{H.c}. \end{aligned} \end{aligned}$$

We define

$$\begin{aligned} & \int_0^{t} ds \text{tr}_q \text{tr}_B [C(t-s)\rho_q(t-s)\otimes \rho_B C(t)] e^{i\omega_r s}\nonumber\\ & = \int_0^{t} ds \text{tr}_q \text{tr}_B [C(t)C(t-s)\rho_q(t-s)\otimes\rho_B]e^{i\omega_r s}=A_{{+}1}(t), \\ & \int_0^{t} ds \text{tr}_q \text{tr}_B [C(t-s)\rho_q(t-s)\otimes \rho_B C(t)]e^{{-}i\omega_r s}\\ & =\int_0^{t} ds \text{tr}_q \text{tr}_B[C(t)C(t-s)\rho_q(t-s)\otimes \rho_B]e^{{-}i\omega_r s}=A_{{-}1}(t), \end{aligned} $$
the Eq. (30) can be rewritten as
$$\begin{aligned} \frac{d\mu_1 (t)}{dt}= & A_{{+}1}(t)[b_1^{\dagger}\mu_1(t)b_1 -b_1 b_1^{\dagger}\mu_1 (t)]+A_{{-}1}(t)[b_1\mu_1(t)b_1^{\dagger} -b_1^{\dagger} b_1\mu_1 (t)]+\text{H.c}. \end{aligned}$$

Now we use some new signs, such as $\rho _{q,ee}(t)=tr_q[|e\rangle \langle e|\rho _q(t)]$, $\rho _{q,gg}(t)=tr_q[|g\rangle \langle g|\rho _q(t)]$ and $\rho _{q,ge}(t)=tr_q[|g\rangle \langle e|\rho _q(t)]$. The dynamics of $\rho _{q,ge}(t)$ and $\rho _{q,eg}(t)$ can be approximately expressed as $\sigma _{-1}(t)=\rho _{q,ge}(t) \simeq \rho _{q,ge}(t-s)e^{-(\Gamma _1/2) s}$, and $\sigma _{+1}(t)=\rho _{q,eg}(t) \simeq \rho _{q,eg}(t-s)e^{-(\Gamma _1/2) s}$, due to the fast dissipates of qubit. The qubit is approximately in the ground states all the time, which means $\rho _{q,ee}(t-s) \simeq 0$ and $\rho _{q,gg}(t-s) \simeq 1$. Under the condition $t \gg \Gamma _1^{-1}$, $A_{+1}$ and $A_{-1}$ will reduce to the simple forms,

$$\begin{aligned} A_{{+}1}={}&G_1^{2}\int_0^{t} \rho_{q,ee}(t-s)e^{2i\omega_r s}e^{-(\Gamma_1/2) s}ds +G_1^{2}\int_0^{t} \rho_{q,gg}(t-s)e^{-(\Gamma_1/2) s}ds \simeq \frac{2}{\Gamma_1},\nonumber\\ A_{{-}1}={}&G_1^{2}\int_0^{t} \rho_{q,ee}(t-s)e^{-(\Gamma_1/2) s}ds +G_1^{2}\int_0^{t} \rho_{q,gg}(t-s)e^{{-}2i\omega_r s}e^{-(\Gamma_1/2) s}ds \simeq \frac{G_1^{2}}{\Gamma_1+2i\omega_r }, \end{aligned}$$
and the equation $\mu _1(t)$ has the form
$$\begin{aligned} \frac{d\mu_1 (t)}{dt}= & \frac{G_1^{2}}{\Gamma_1}[2b_1^{\dagger}\mu_1(t)b_1 -b_1b_1^{\dagger}\mu_1(t)-\mu_1(t)b_1b_1^{\dagger}]\\ & +\frac{G_1^{2}}{\Gamma_1+2i\omega_r}[b_1\mu_1(t)b_1^{\dagger} -b_1^{\dagger}b_1\mu_1(t)]\\ & +\frac{G_1^{2}}{\Gamma_1-2i\omega_r}[b_1\mu_1(t)b_1^{\dagger} -\mu_1(t)b_1^{\dagger}b_1] \end{aligned}$$

We can find from Eq. (32) that $\tau _{\mu } \sim 2\Gamma _1/G_1^{2}$, since the equation was obtained under the condition $\tau _{\mu } \gg t \gg \Gamma _1^{-1}$. Thus, the condition for adiabatically eliminating the qubit’s excited states of the gain subsystem is $\Gamma _1 \gg \frac {2G_1^{2}}{\Gamma _1}$. The qubit of the loss subsystem can also be eliminated adiabatically in the same way.

Funding

National Natural Science Foundation of China (No.11804270, No.12174303, No. 11774284); China Postdoctoral Science Foundation (No. 2018M631136); Natural Science Foundation of Henan Province (No. 222300420233).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

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. C. M. Bender and S. Boettcher, “Real spectra in non-Hermitian Hamiltonians having PT symmetry,” Phys. Rev. Lett. 80(24), 5243–5246 (1998). [CrossRef]  

2. C. M. Bender, “Introduction to $\mathcal {PT}$-symmetric quantum theory,” Contemp. Phys. 46(4), 277–292 (2005). [CrossRef]  

3. C. M. Bender, B. K. Berntson, D. Parker, and E. Samuel, “Observation of $\mathcal {PT}$ phase transition in a simple mechanical system,” Am. J. Phys. 81(3), 173–179 (2013). [CrossRef]  

4. H. Ramezani, T. Kottos, R. El-Ganainy, and D. N. Christodoulides, “Unidirectional nonlinear $\mathcal {PT}$-symmetric optical structures,” Phys. Rev. A 82(4), 043803 (2010). [CrossRef]  

5. A. Guo, G. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. Siviloglou, and D. Christodoulides, “Observation of $\mathcal {PT}$-Symmetry Breaking in Complex Optical Potentials,” Phys. Rev. Lett. 103(9), 093902 (2009). [CrossRef]  

6. B. Peng, Ş. Özdemir, S. Rotter, H. Yilmaz, M. Liertzer, F. Monifi, C. Bender, F. Nori, and L. Yang, “Loss-induced suppression and revival of lasing,” Science 346(6207), 328–332 (2014). [CrossRef]  

7. L. Fan and J. Mei, “Acoustic metagrating circulators: Nonreciprocal, robust, and tunable manipulation with unitary efficiency,” Phys. Rev. Appl. 15(6), 064002 (2021). [CrossRef]  

8. L. Feng, Y.-L. Xu, W. S. Fegadolli, M.-H. Lu, J. E. Oliveira, V. R. Almeida, Y.-F. Chen, and A. Scherer, “Experimental demonstration of a unidirectional reflectionless parity-time metamaterial at optical frequencies,” Nat. Mater. 12(2), 108–113 (2013). [CrossRef]  

9. H. Hodaei, M.-A. Miri, M. Heinrich, D. N. Christodoulides, and M. Khajavikhan, “Parity-time–symmetric microring lasers,” Science 346(6212), 975–978 (2014). [CrossRef]  

10. D.-X. Chen, Y. Zhang, J.-L. Zhao, Q.-C. Wu, Y.-L. Fang, C.-P. Yang, and F. Nori, “Quantum state discrimination in a $\mathcal {PT}$-symmetric system,” Phys. Rev. A 106(2), 022438 (2022). [CrossRef]  

11. Y.-L. Fang, J.-L. Zhao, Y. Zhang, D.-X. Chen, Q.-C. Wu, Y.-H. Zhou, C.-P. Yang, and F. Nori, “Experimental demonstration of coherence flow in PT-and anti-PT-symmetric systems,” Commun. Phys. 4(1), 223 (2021). [CrossRef]  

12. S. Liu, S. Ma, C. Yang, L. Zhang, W. Gao, Y. J. Xiang, T. J. Cui, and S. Zhang, “Gain-and loss-induced topological insulating phase in a non-Hermitian electrical circuit,” Phys. Rev. Appl. 13(1), 014047 (2020). [CrossRef]  

13. R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, “Non-Hermitian physics and PT symmetry,” Nat. Phys. 14(1), 11–19 (2018). [CrossRef]  

14. C. E. Rüter, K. G. Makris, R. El-Ganainy, D. N. Christodoulides, M. Segev, and D. Kip, “Observation of parity-time symmetry in optics,” Nat. Phys. 6(3), 192–195 (2010). [CrossRef]  

15. B. Peng, Ş. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity–time-symmetric whispering-gallery microcavities,” Nat. Phys. 10(5), 394–398 (2014). [CrossRef]  

16. H. Xu, D.-G. Lai, Y.-B. Qian, B.-P. Hou, A. Miranowicz, and F. Nori, “Optomechanical dynamics in the $\mathcal {PT}$-and broken-$\mathcal {PT}$-symmetric regimes,” Phys. Rev. A 104(5), 053518 (2021). [CrossRef]  

17. H. Jing, S. Özdemir, X.-Y. Lü, J. Zhang, L. Yang, and F. Nori, “$\mathcal {PT}$-symmetric Phonon Laser,” Phys. Rev. Lett. 113(5), 053604 (2014). [CrossRef]  

18. X. Y. Lü, H. Jing, J. Y. Ma, and Y. Wu, “$\mathcal {PT}$-Symmetry-Breaking Chaos in Optomechanics,” Phys. Rev. Lett. 114(25), 253601 (2015). [CrossRef]  

19. Y. Jiao, H. Lü, J. Qian, Y. Li, and H. Jing, “Nonlinear optomechanics with gain and loss: amplifying higher-order sideband and group delay,” New J. Phys. 18(8), 083034 (2016). [CrossRef]  

20. R. Fleury, D. Sounas, and A. Alù, “An invisible acoustic sensor based on parity-time symmetry,” Nat. Commun. 6(1), 5905 (2015). [CrossRef]  

21. Z.-P. Liu, J. Zhang, Ş. K. Özdemir, B. Peng, H. Jing, X.-Y. Lü, C.-W. Li, L. Yang, F. Nori, and Y.-X. Liu, “Metrology with $\mathcal {PT}$-Symmetric Cavities: Enhanced Sensitivity near the $\mathcal {PT}$-Phase Transition,” Phys. Rev. Lett. 117(11), 110802 (2016). [CrossRef]  

22. T. Chen, W. Gou, D. Xie, T. Xiao, W. Yi, J. Jing, and B. Yan, “Quantum Zeno effects across a parity-time symmetry breaking transition in atomic momentum space,” npj Quantum Inf. 7(1), 78 (2021). [CrossRef]  

23. V. Domínguez-Rocha, R. Thevamaran, F. Ellis, and T. Kottos, “Environmentally induced exceptional points in elastodynamics,” Phys. Rev. Appl. 13(1), 014060 (2020). [CrossRef]  

24. K. Ding, G. Ma, Z. Zhang, and C. T. Chan, “Experimental demonstration of an anisotropic exceptional point,” Phys. Rev. Lett. 121(8), 085702 (2018). [CrossRef]  

25. W. Xiong, Z. Li, Y. Song, J. Chen, G. Q. Zhang, and M. Wang, “Higher-order exceptional point in a pseudo-Hermitian cavity optomechanical system,” Phys. Rev. A 104(6), 063508 (2021). [CrossRef]  

26. W. Zhu, X. Fang, D. Li, Y. Sun, Y. Li, Y. Jing, and H. Chen, “Simultaneous observation of a topological edge state and exceptional point in an open and non-Hermitian acoustic system,” Phys. Rev. Lett. 121(12), 124501 (2018). [CrossRef]  

27. R. El-Ganainy, K. Makris, D. Christodoulides, and Z. H. Musslimani, “Theory of coupled optical PT-symmetric structures,” Opt. Lett. 32(17), 2632–2634 (2007). [CrossRef]  

28. F. Zhang, Y. Feng, X. Chen, L. Ge, and W. Wan, “Synthetic anti-PT symmetry in a single microcavity,” Phys. Rev. Lett. 124(5), 053901 (2020). [CrossRef]  

29. X.-L. Zhang, S. Wang, B. Hou, and C. T. Chan, “Dynamically Encircling Exceptional Points: In situ Control of Encircling Loops and the Role of the Starting Point,” Phys. Rev. X 8(2), 021066 (2018). [CrossRef]  

30. W. Song, W. Sun, C. Chen, Q. Song, S. Xiao, S. Zhu, and T. Li, “Breakup and recovery of topological zero modes in finite non-Hermitian optical lattices,” Phys. Rev. Lett. 123(16), 165701 (2019). [CrossRef]  

31. Z. Xiao, H. Li, T. Kottos, and A. Alù, “Enhanced Sensing and Nondegraded Rhermal Noise Performance Based on $\mathcal {PT}$-Symmetric Electronic Circuits with a Sixth-Order Exceptional Point,” Phys. Rev. Lett. 123(21), 213901 (2019). [CrossRef]  

32. G.-Q. Zhang, Z. Chen, W. Xiong, C.-H. Lam, and J. Q. You, “Parity-symmetry-breaking quantum phase transition via parametric drive in a cavity magnonic system,” Phys. Rev. B 104(6), 064423 (2021). [CrossRef]  

33. J. Schindler, A. Li, M. C. Zheng, F. M. Ellis, and T. Kottos, “Experimental study of active LRC circuits with $\mathcal {PT}$ symmetries,” Phys. Rev. A 84(4), 040101 (2011). [CrossRef]  

34. Y.-L. Liu, R. Wu, J. Zhang, Ş. K. Özdemir, L. Yang, F. Nori, and Y.-x. Liu, “Controllable optical response by modifying the gain and loss of a mechanical resonator and cavity mode in an optomechanical system,” Phys. Rev. A 95(1), 013843 (2017). [CrossRef]  

35. J. Lan, L. Wang, X. Zhang, M. Lu, and X. Liu, “Acoustic multifunctional logic gates and amplifier based on passive parity-time symmetry,” Phys. Rev. Appl. 13(3), 034047 (2020). [CrossRef]  

36. S.-N. Huai, Y.-L. Liu, J. Zhang, L. Yang, and Y.-X. Liu, “Enhanced sideband responses in a $\mathcal {PT}$-symmetric-like cavity magnomechanical system,” Phys. Rev. A 99(4), 043803 (2019). [CrossRef]  

37. Ş. K. Özdemir, S. Rotter, F. Nori, and L. Yang, “Parity–time symmetry and exceptional points in photonics,” Nat. Mater. 18(8), 783–798 (2019). [CrossRef]  

38. F. Klauck, L. Teuber, M. Ornigotti, M. Heinrich, S. Scheel, and A. Szameit, “Observation of PT-symmetric quantum interference,” Nat. Photonics 13(12), 883–887 (2019). [CrossRef]  

39. X.-W. Xu, Y.-X. Liu, C.-P. Sun, and Y. Li, “Mechanical $\mathcal {PT}$ symmetry in coupled optomechanical systems,” Phys. Rev. A 92(1), 013852 (2015). [CrossRef]  

40. R. Manenti, A. F. Kockum, A. Patterson, T. Behrle, J. Rahamim, G. Tancredi, F. Nori, and P. J. Leek, “Circuit quantum acoustodynamics with surface acoustic waves,” Nat. Commun. 8(1), 975 (2017). [CrossRef]  

41. Z. L. Xiang, S. Ashhab, J. Q. You, and F. Nori, “Hybrid quantum circuits: Superconducting circuits interacting with other quantum systems,” Rev. Mod. Phys. 85(2), 623–653 (2013). [CrossRef]  

42. A. Blais, A. L. Grimsmo, S. M. Girvin, and A. Wallraff, “Circuit quantum electrodynamics,” Rev. Mod. Phys. 93(2), 025005 (2021). [CrossRef]  

43. X. Gu, A. F. Kockum, A. Miranowicz, Y.-x. Liu, and F. Nori, “Microwave photonics with superconducting quantum circuits,” Phys. Rep. 718-719, 1–102 (2017). [CrossRef]  

44. A. F. Kockum and F. Nori, “Quantum bits with Josephson junctions,” in Fundamentals and Frontiers of the Josephson Effect, F. Tafuri, ed. (Springer, 2019).

45. H. Paik, D. I. Schuster, L. S. Bishop, G. Kirchmair, G. Catelani, A. P. Sears, B. Johnson, M. Reagor, L. Frunzio, L. I. Glazman, S. M. Girvin, M. H. Devoret, and R. J. Schoelkopf, “Observation of high coherence in josephson junction qubits measured in a three-dimensional circuit QED architecture,” Phys. Rev. Lett. 107(24), 240501 (2011). [CrossRef]  

46. X. Wang, H.-R. Li, P.-B. Li, C.-W. Jiang, H. Gao, and F.-l. Li, “Preparing ground states and squeezed states of nanomechanical cantilevers by fast dissipation,” Phys. Rev. A 90(1), 013838 (2014). [CrossRef]  

47. 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 464(7289), 697–703 (2010). [CrossRef]  

48. X. Wang, H.-R. Li, and F.-L. Li, “Generating synthetic magnetism via Floquet engineering auxiliary qubits in phonon-cavity-based lattice,” New J. Phys. 22(3), 033037 (2020). [CrossRef]  

49. J.-J. Xue, W.-Q. Zhu, Y.-N. He, X. Wang, and H.-R. Li, “Two-acoustic-cavity interaction mediated by superconducting artificial atoms,” Quantum Inf. Process. 19(9), 333 (2020). [CrossRef]  

50. Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, “Coherent coupling between a ferromagnetic magnon and a superconducting qubit,” Science 349(6246), 405–408 (2015). [CrossRef]  

51. X. Wang, A. Miranowicz, H.-R. Li, and F. Nori, “Hybrid quantum device with a carbon nanotube and a flux qubit for dissipative quantum engineering,” Phys. Rev. B 95(20), 205415 (2017). [CrossRef]  

52. D. Zoepfl, M. Juan, C. Schneider, and G. Kirchmair, “Single-photon cooling in microwave magnetomechanics,” Phys. Rev. Lett. 125(2), 023601 (2020). [CrossRef]  

53. W. Xiong, D.-Y. Jin, J. Jing, C.-H. Lam, and J. Q. You, “Controllable coupling between a nanomechanical resonator and a coplanar-waveguide resonator via a superconducting flux qubit,” Phys. Rev. A 92(3), 032318 (2015). [CrossRef]  

54. M. Mariantoni, F. Deppe, A. Marx, R. Gross, F. K. Wilhelm, and E. Solano, “Two-resonator circuit quantum electrodynamics: A superconducting quantum switch,” Phys. Rev. B 78(10), 104508 (2008). [CrossRef]  

55. G. M. Reuther, D. Zueco, F. Deppe, E. Hoffmann, E. P. Menzel, T. Weißl, M. Mariantoni, S. Kohler, A. Marx, E. Solano, R. Gross, and P. Hänggi, “Two-resonator circuit quantum electrodynamics: Dissipative theory,” Phys. Rev. B 81(14), 144510 (2010). [CrossRef]  

56. F. Xue, Y. Wang, C. Sun, H. Okamoto, H. Yamaguchi, and K. Semba, “Controllable coupling between flux qubit and nanomechanical resonator by magnetic field,” New J. Phys. 9(2), 35 (2007). [CrossRef]  

57. A. Baust, E. Hoffmann, M. Haeberlein, M. Schwarz, P. Eder, J. Goetz, F. Wulschner, E. Xie, L. Zhong, F. Quijandría, D. Zueco, J. García Ripoll, L. García-Álvarez, G. Romero, E. Solano, K. Fedorov, E. Menzel, F. Deppe, A. Marx, and R. Gross, “Ultrastrong coupling in two-resonator circuit QED,” Phys. Rev. B 93(21), 214501 (2016). [CrossRef]  

58. Y.-J. Zhao, Y.-L. Liu, Y.-x. Liu, and F. Nori, “Generating nonclassical photon states via longitudinal couplings between superconducting qubits and microwave fields,” Phys. Rev. A 91(5), 053820 (2015). [CrossRef]  

59. Y.-X. Liu, C.-X. Yang, H.-C. Sun, and X.-B. Wang, “Coexistence of single-and multi-photon processes due to longitudinal couplings between superconducting flux qubits and external fields,” New J. Phys. 16(1), 015031 (2014). [CrossRef]  

60. Y.-J. Zhao, C. Wang, X. Zhu, and Y.-x. Liu, “Engineering entangled microwave photon states through multiphoton interactions between two cavity fields and a superconducting qubit,” Sci. Rep. 6(1), 1–17 (2016). [CrossRef]  

61. Y.-X. Liu, H.-C. Sun, Z. Peng, A. Miranowicz, J. S. Tsai, and F. Nori, “Controllable microwave three-wave mixing via a single three-level superconducting quantum circuit,” Sci. Rep. 4(1), 7289 (2014). [CrossRef]  

62. X. L. He, J. Q. You, Y.-x. Liu, L. Wei, and F. Nori, “Switchable coupling between charge and flux qubits,” Phys. Rev. B 76(2), 024517 (2007). [CrossRef]  

63. X. Wang, A. Miranowicz, H.-R. Li, and F. Nori, “Multiple-output microwave single-photon source using superconducting circuits with longitudinal and transverse couplings,” Phys. Rev. A 94(5), 053858 (2016). [CrossRef]  

64. N. Lambert, M. Cirio, M. Delbecq, G. Allison, M. Marx, S. Tarucha, and F. Nori, “Amplified and tunable transverse and longitudinal spin-photon coupling in hybrid circuit-QED,” Phys. Rev. B 97(12), 125429 (2018). [CrossRef]  

65. R. Stassi and F. Nori, “Long-lasting quantum memories: Extending the coherence time of superconducting artificial atoms in the ultrastrong-coupling regime,” Phys. Rev. A 97(3), 033823 (2018). [CrossRef]  

66. L. Shao, W. Mao, S. Maity, N. Sinclair, Y. Hu, L. Yang, and M. Lončar, “Non-reciprocal transmission of microwave acoustic waves in nonlinear parity-time symmetric resonators,” Nat. Electron. 3(5), 267–272 (2020). [CrossRef]  

67. F. Quijandría, U. Naether, S. K. Özdemir, F. Nori, and D. Zueco, “$\mathcal {PT}$-symmetric circuit QED,” Phys. Rev. A 97(5), 053846 (2018). [CrossRef]  

68. A. Purkayastha, M. Kulkarni, and Y. N. Joglekar, “Emergent $\mathcal {PT}$ symmetry in a double-quantum-dot circuit QED setup,” Phys. Rev. Res. 2(4), 043075 (2020). [CrossRef]  

69. F. Paauw, A. Fedorov, C. M. Harmans, and J. Mooij, “Tuning the gap of a superconducting flux qubit,” Phys. Rev. Lett. 102(9), 090501 (2009). [CrossRef]  

70. M. Schwarz, J. Goetz, Z. Jiang, T. Niemczyk, F. Deppe, A. Marx, and R. Gross, “Gradiometric flux qubits with a tunable gap,” New J. Phys. 15(4), 045001 (2013). [CrossRef]  

71. X. Zhu, A. Kemp, S. Saito, and K. Semba, “Coherent operation of a gap-tunable flux qubit,” Appl. Phys. Lett. 97(10), 102503 (2010). [CrossRef]  

72. O. Astafiev, A. M. Zagoskin, A. Abdumalikov Jr, Y. A. Pashkin, T. Yamamoto, K. Inomata, Y. Nakamura, and J. S. Tsai, “Resonance fluorescence of a single artificial atom,” Science 327(5967), 840–843 (2010). [CrossRef]  

73. X. Zhu, S. Saito, A. Kemp, K. Kakuyanagi, S.-I. Karimoto, H. Nakano, W. J. Munro, Y. Tokura, M. S. Everitt, N. Kae, M. Kasu, N. Mizuochi, and S. Kouichi, “Coherent coupling of a superconducting flux qubit to an electron spin ensemble in diamond,” Nature 478(7368), 221–224 (2011). [CrossRef]  

74. L. Zhou, Z. Gong, Y.-X. Liu, C. Sun, and F. Nori, “Controllable scattering of a single photon inside a one-dimensional resonator waveguide,” Phys. Rev. Lett. 101(10), 100501 (2008). [CrossRef]  

75. X. Wang, H.-r. Li, D.-X. Chen, W.-X. Liu, and F.-l. Li, “Tunable electromagnetically induced transparency in a composite superconducting system,” Opt. Commun. 366, 321–327 (2016). [CrossRef]  

76. P.-B. Li, Y.-C. Liu, S.-Y. Gao, Z.-L. Xiang, P. Rabl, Y.-F. Xiao, and F.-L. Li, “Hybrid Quantum Device Based on NV Centers in Diamond Nanomechanical Resonators Plus Superconducting Waveguide Cavities,” Phys. Rev. Appl. 4(4), 044003 (2015). [CrossRef]  

77. L. Tian, M. Allman, and R. Simmonds, “Parametric coupling between macroscopic quantum resonators,” New J. Phys. 10(11), 115001 (2008). [CrossRef]  

78. P. Rabl, S. J. Kolkowitz, F. Koppens, J. Harris, P. Zoller, and M. D. Lukin, “A quantum spin transducer based on nanoelectromechanical resonator arrays,” Nat. Phys. 6(8), 602–608 (2010). [CrossRef]  

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

80. S.-B. Zheng and G.-C. Guo, “Efficient scheme for two-atom entanglement and quantum information processing in cavity QED,” Phys. Rev. Lett. 85(11), 2392–2395 (2000). [CrossRef]  

81. J. A. Sidles, J. L. Garbini, K. Bruland, D. Rugar, O. Züger, S. Hoen, and C. Yannoni, “Magnetic resonance force microscopy,” Rev. Mod. Phys. 67(1), 249–265 (1995). [CrossRef]  

82. P.-M. Billangeon, J. S. Tsai, and Y. Nakamura, “Circuit-QED-based scalable architectures for quantum information processing with superconducting qubits,” Phys. Rev. B 91(9), 094517 (2015). [CrossRef]  

83. T. J. Kippenberg and K. J. Vahala, “Cavity opto-mechanics,” Opt. Express 15(25), 17172–17205 (2007). [CrossRef]  

84. J. I. Cirac, R. Blatt, P. Zoller, and W. D. Phillips, “Laser cooling of trapped ions in a standing wave,” Phys. Rev. A 46(5), 2668–2681 (1992). [CrossRef]  

85. X. Zhu, L. Feng, P. Zhang, X. Yin, and X. Zhang, “One-way invisible cloak using parity-time symmetric transformation optics,” Opt. Lett. 38(15), 2821–2824 (2013). [CrossRef]  

86. X. Zhu, B. Liang, W. Kan, X. Zou, and J. Cheng, “Acoustic cloaking by a superlens with single-negative materials,” Phys. Rev. Lett. 106(1), 014301 (2011). [CrossRef]  

87. B. Peng, Ş. K. Özdemir, W. Chen, F. Nori, and L. Yang, “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nat. Commun. 5(1), 5082 (2014). [CrossRef]  

88. H.-C. Sun, Y.-X. Liu, H. Ian, J. Q. You, E. Il’Ichev, and F. Nori, “Electromagnetically induced transparency and Autler-Townes splitting in superconducting flux quantum circuits,” Phys. Rev. A 89(6), 063822 (2014). [CrossRef]  

89. H. Wang, X. Gu, Y.-X. Liu, A. Miranowicz, and F. Nori, “Optomechanical analog of two-color electromagnetically induced transparency: Photon transmission through an optomechanical device with a two-level system,” Phys. Rev. A 90(2), 023817 (2014). [CrossRef]  

90. Y.-X. Liu, X.-W. Xu, A. Miranowicz, and F. Nori, “From blockade to transparency: Controllable photon transmission through a circuit-QED system,” Phys. Rev. A 89(4), 043818 (2014). [CrossRef]  

91. H. Jing, Ş. K. Özdemir, Z. Geng, J. Zhang, X.-Y. Lü, B. Peng, L. Yang, and F. Nori, “Optomechanically-induced transparency in parity-time-symmetric microresonators,” Sci. Rep. 5(1), 9663 (2015). [CrossRef]  

92. Q.-C. Liu, T.-F. Li, X.-Q. Luo, H. Zhao, W. Xiong, Y.-S. Zhang, Z. Chen, J. Liu, W. Chen, F. Nori, J. S. Tsai, and J. Q. You, “Method for identifying electromagnetically induced transparency in a tunable circuit quantum electrodynamics system,” Phys. Rev. A 93(5), 053838 (2016). [CrossRef]  

93. D.-G. Lai, X. Wang, W. Qin, B.-P. Hou, F. Nori, and J.-Q. Liao, “Tunable optomechanically induced transparency by controlling the dark-mode effect,” Phys. Rev. A 102(2), 023707 (2020). [CrossRef]  

94. X. Wang, A. Miranowicz, H.-R. Li, F.-L. Li, and F. Nori, “Two-color electromagnetically induced transparency via modulated coupling between a mechanical resonator and a qubit,” Phys. Rev. A 98(2), 023821 (2018). [CrossRef]  

95. Y. Liu, A. Talbi, O. B. Matar, P. Pernod, and B. Djafari-Rouhani, “Autler-Townes splitting and acoustically induced transparency based on love waves interacting with a pillared metasurface,” Phys. Rev. Appl. 11(6), 064066 (2019). [CrossRef]  

96. A. Santillán and S. I. Bozhevolnyi, “Acoustic transparency and slow sound using detuned acoustic resonators,” Phys. Rev. B 84(6), 064304 (2011). [CrossRef]  

97. F. Liu, M. Ke, A. Zhang, W. Wen, J. Shi, Z. Liu, and P. Sheng, “Acoustic analog of electromagnetically induced transparency in periodic arrays of square rods,” Phys. Rev. E 82(2), 026601 (2010). [CrossRef]  

98. A. H. Safavi-Naeini, T. Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature 472(7341), 69–73 (2011). [CrossRef]  

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 (8)

Fig. 1.
Fig. 1. (a) Schematic of a superconducting hybrid quantum system for studying $\mathcal {PT}$ symmetry. Each silicon nanomechanical cantilever resonator with dimensions $(l,w,t)$ is coupled with a flux qubit via an external magnetic tip attached to the cantilever. Two resonators separated by a distance $d$ are coupled indirectly through capacitive interactions with nearby wires. The inset shows the lumped-circuit layout of the flux qubit. (b) Circuit model that describes the coupling between two silicon nanomechanical cantilever resonators.
Fig. 2.
Fig. 2. Energy diagram of blue-sideband driving qubit $1$ with frequency $\omega _{p1} =\omega _{q1} +\omega _r$ (a) and red-sideband driving qubit $2$ with frequency $\omega _{p2} =\omega _{q2} -\omega _r$(b).
Fig. 3.
Fig. 3. The mean phonon number of the nanomechanical cantilever resonator as a function of time $gt/2\pi$ in the gain subsystem (a) and loss subsystem (b). The analytical and numerical solution error analysis of the gain (c) and dissipation subsystem (d). Here, we choose $\omega _r = 10$ MHz, $\Gamma _i = 4$ MHz, and $G_i= 0.05$ MHz (corresponding to $\eta _i = 2$, $\epsilon _i = 0.5$ MHz, and $g_i = 0.5$ MHz) (i=1,2).
Fig. 4.
Fig. 4. $\mathcal {PT}$ phase transition in our hybrid quantum system. The real (a,c) and imaginary (b,d) parts of the eigenfrequencies $\omega _{\pm }$ plotted as a function of coupling strength $J$. The results for the case of balanced gain and loss are shown in (a) and (b), and for the case of the unbalanced gain-to-loss are plotted in (c) and (d). The black-solid and red-dashed curves in (a) and (b) represent the system gain and loss balance, while the blue-solid and green-dashed curves represent the system without mechanical gain. The black-solid and red-dashed curves in (c) and (d) are calculated with $\gamma = \kappa /2$, and the blue-solid and green-dashed are calculated with $\gamma = 2\kappa$. Broken and unbroken $\mathcal {PT}$-symmetric regimes are distinguished by different background colors.
Fig. 5.
Fig. 5. Transmission coefficient $T$ and the dispersion Im($t_p$) versus detuning $\Delta$ with $J/\gamma = 0.425$ are shown in (a) and (b), respectively. The transmission coefficient $T$ as a function of detuning $\Delta$ for different coupling strengths $J$ of the gain-loss system is plotted in (c) and (d), and the loss-loss system is plotted in (e) and (f). The other system parameters are $\gamma = 0.4 \text {MHz}$, $\kappa = 0.4 \text {MHz}$, and $\kappa ^{\prime } = 0.2 \text {MHz}$.
Fig. 6.
Fig. 6. The logarithm of the transmission coefficient $T$ versus coupling strength $J$ for detuning $\Delta = 0$. The black solid line represents the case in which gain and loss are balanced. The red solid line indicates the case where there is no gain in the system.
Fig. 7.
Fig. 7. The transmission coefficient $T$ and dispersion Im($t_p$) versus the detuning $\Delta$ with $J = 11\gamma$ are plotted in (a) and (b), respectively. The blue-solid curve (red-dot-solid curve) represents the gain-loss system (loss-loss system). Here, we set $\gamma = 0.1\text {MHz}$, $\kappa = 400 \gamma$, and $\kappa ^{\prime } = 1/2 \kappa$.
Fig. 8.
Fig. 8. Phase $\theta$ and group delay $\tau$ for the $\mathcal {PT}$-symmetric system with gain and loss (see the green-solid curve) and the system without gain (see the orange-dashed curve). The parameters are the same as Fig. 7.

Equations (34)

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

H q i = 1 2 ϵ i ( Φ z i ) σ ¯ z i + Ω i ( Φ x i ) σ ¯ x i ,
H q i = 1 2 ω q i σ z i + g i σ z i ( b i + b i ) + η i ε p i cos ( ω p i t ) σ x i ,
H c = J ( b 1 + b 1 ) ( b 2 + b 2 ) ,
J = 1 C 2 C w 2 U 2 ( 2 C + C w ) 3 b i 2 h 2 .
H total = i = 1 , 2 H r i + i = 1 , 2 H q i + H c .
H total = i = 1 , 2 ω r i b i b i + i = 1 , 2 1 2 Δ i σ z i + i = 1 , 2 g i σ z i ( b i + b i ) + i = 1 , 2 η i ε i ( σ + i + σ i ) + J ( b 1 b 2 + b 2 b 1 ) ,
H 1 = e S H total e S ,
H 1 = i = 1 , 2 ( 1 2 Δ i σ z i + ω r b i b i ) + i = 1 , 2 [ η i ε i σ + i e 2 β i ( b i b i ) + H.c. ] β 1 σ z 1 ( b 1 + b 1 ) β 2 σ z 2 ( b 2 + b 2 ) + J ( b 1 b 2 + b 2 b 1 ) ,
H 2 = i = 1 , 2 2 η i ϵ i β i [ σ + i b i e i ( Δ i + ω r ) t + σ i b i e i ( Δ i + ω r ) t + σ + i b i e i ( Δ i ω r ) t + σ i b i e i ( Δ i ω r ) t ] + β 1 σ z 1 ( b 1 e i ω r t + b 1 e i ω r t ) + β 2 σ z 2 ( b 2 e i ω r t + b 2 e i ω r t ) + J ( b 1 b 2 + b 2 b 1 ) .
H 3 = G 1 ( σ + 1 e i ω r t + σ 1 e i ω r t ) ( b 1 e i ω r t + b 1 e i ω r t ) + G 2 ( σ + 2 e i ω r t + σ 2 e i ω r t ) ( b 2 e i ω r t + b 2 e i ω r t ) + β 1 σ z 1 ( b 1 e i ω r t + b 1 e i ω r t ) + β 2 σ z 2 ( b 2 e i ω r t + b 2 e i ω r t ) + J ( b 1 b 2 + b 2 b 1 ) .
d ρ ( t ) d t = i [ H 3 , ρ ( t ) ] + i = 1 , 2 Γ i D [ σ i ] + n t h ξ i D [ b i ] ρ ( t ) + ( n t h + 1 ) ξ i D [ b i ] ρ ( t ) .
d μ 1 ( t ) d t = D [ b 1 , 2 G 1 2 Γ 1 + n t h ξ 1 ] μ 1 + D [ b 1 , 2 G 1 2 Γ 1 + 2 i ω r + ( n t h + 1 ) ξ 1 ) ] μ 1 d μ 1 ( t ) d t = D [ b 2 , 2 G 2 2 Γ 2 + 2 i ω r + n t h ξ 2 ] μ 2 + D [ b 2 , 2 G 2 2 Γ 2 + ( n t h + 1 ) ξ 2 ) ] μ 2 .
d n ¯ 1 d t = ( 2 G 1 2 Γ 1 ξ 1 ) n ¯ 1 + ( 2 G 1 2 Γ 1 + n t h ξ 1 ) , d n ¯ 2 d t = ( 2 G 2 2 Γ 2 + ξ 2 ) n ¯ 2 + n t h ξ 2 .
n ¯ 1 ( t ) = n ¯ 1 ( 0 ) + 1 e 2 G 1 2 t / Γ 1 1 , n ¯ 2 ( t ) = n ¯ 2 ( 0 ) e 2 G 2 2 t / Γ 2 .
H = ( ω r + i γ 2 ) b 1 b 1 + ( ω r i κ 2 ) b 1 b 1 + J ( b 1 b 2 + b 2 b 1 ) ,
ω ± = ω r + i 4 ( γ κ ) ± J 2 ( γ + κ ) 16 .
H = ( ω r + i γ 2 ) b 1 b 1 + ( ω r i κ 2 ) b 2 b 2 + J ( b 1 b 2 + b 2 b 1 ) + ε p ( b 2 e i ω p t + b 2 e i ω p t ) ,
d b 1 d t = i ω r b 1 + γ 2 b 1 i J b 2 d b 2 d t = i ω r b 2 κ 2 b 2 i J b 1 i ε p e i ω r t κ b in2 ,
d B 1 d t = i ( ω ω r ) B 1 + γ 2 B 1 i J B 2 d B 2 d t = i ( ω ω r ) B 2 κ 2 B 2 i J B 1 i ε p e i ω r t κ B in2 .
B 1 = i J κ J 2 + ( i Δ κ / 2 ) ( i Δ + γ / 2 ) B in2 B 2 = κ ( i Δ + γ / 2 ) J 2 + ( i Δ κ / 2 ) ( i Δ + γ / 2 ) B in2 ,
T = | B out2 B in2 | 2 = | 1 + κ ( i Δ + γ / 2 ) J 2 + ( i Δ κ / 2 ) ( i Δ + γ / 2 ) | 2 .
T = | 1 + κ γ / 2 J 2 κ γ / 4 | 2 ,
τ = d θ d Δ ,
H gain = G 1 ( σ + 1 e i ω r t + σ 1 e i ω r t ) ( b 1 e i ω r t + b 1 e i ω r t ) .
H A 1 = H gain + V ( t ) .
H A 2 = G 1 U ^ ( t ) ( σ + 1 e i ω r t + σ 1 e i ω r t ) U ^ ( t ) ( b 1 e i ω r t + b 1 e i ω r t ) ,
d μ 1 ( t ) d t = 0 t d t tr q tr B [ H A 2 ( t ) , [ H A 2 ( t ) , μ ( t ) ρ q ( t ) ρ B ] ] .
tr q tr B [ H A 2 ( t ) , ρ ^ ( 0 ) ] = 0 ,
d μ 1 ( t ) d t = 0 t d t tr q tr B [ H A 2 ( t ) , [ H A 2 ( t ) , μ ( t ) ρ q ( t ) ρ B ] ] = 0 t d t tr q tr B [ H A 2 ( t ) μ ( t ) ρ q ( t ) ρ B H A 2 ( t ) H A 2 ( t ) H A 2 ( t ) μ ( t ) ρ q ( t ) ρ B ] + H.c. ,
d μ 1 ( t ) d t = 0 t d s tr q tr B [ C ( t s ) ρ q ( t s ) ρ B C ( t ) ] e i ω r s b 1 μ ( t ) b 1 + 0 t d s tr q tr B [ C ( t s ) ρ q ( t s ) ρ B C ( t ) ] e i ω r s b 1 μ ( t ) b 1 0 t d s tr q tr B [ C ( t ) C ( t s ) ρ q ( t s ) ρ B ] e i ω r s b 1 b 1 μ ( t ) 0 t d s tr q tr B [ C ( t ) C ( t s ) ρ q ( t s ) ρ B ] e i ω r s b 1 b 1 μ ( t ) + H.c .
0 t d s tr q tr B [ C ( t s ) ρ q ( t s ) ρ B C ( t ) ] e i ω r s = 0 t d s tr q tr B [ C ( t ) C ( t s ) ρ q ( t s ) ρ B ] e i ω r s = A + 1 ( t ) , 0 t d s tr q tr B [ C ( t s ) ρ q ( t s ) ρ B C ( t ) ] e i ω r s = 0 t d s tr q tr B [ C ( t ) C ( t s ) ρ q ( t s ) ρ B ] e i ω r s = A 1 ( t ) ,
d μ 1 ( t ) d t = A + 1 ( t ) [ b 1 μ 1 ( t ) b 1 b 1 b 1 μ 1 ( t ) ] + A 1 ( t ) [ b 1 μ 1 ( t ) b 1 b 1 b 1 μ 1 ( t ) ] + H.c .
A + 1 = G 1 2 0 t ρ q , e e ( t s ) e 2 i ω r s e ( Γ 1 / 2 ) s d s + G 1 2 0 t ρ q , g g ( t s ) e ( Γ 1 / 2 ) s d s 2 Γ 1 , A 1 = G 1 2 0 t ρ q , e e ( t s ) e ( Γ 1 / 2 ) s d s + G 1 2 0 t ρ q , g g ( t s ) e 2 i ω r s e ( Γ 1 / 2 ) s d s G 1 2 Γ 1 + 2 i ω r ,
d μ 1 ( t ) d t = G 1 2 Γ 1 [ 2 b 1 μ 1 ( t ) b 1 b 1 b 1 μ 1 ( t ) μ 1 ( t ) b 1 b 1 ] + G 1 2 Γ 1 + 2 i ω r [ b 1 μ 1 ( t ) b 1 b 1 b 1 μ 1 ( t ) ] + G 1 2 Γ 1 2 i ω r [ b 1 μ 1 ( t ) b 1 μ 1 ( t ) b 1 b 1 ]
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.