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

One-way quantum state transfer in a lossy coupled-cavity array

Open Access Open Access

Abstract

Quantum state transfer plays an important role in quantum information processing, and it has been obtained many of the theoretical and experimental triumphs. But designing a dissipation-assisted scheme to transfer a quantum state is still by no means trivial. Here we put forward an easier scheme to dissipatively transfer an arbitrary quantum state from a sender to a receiver with two four-level atoms and three lasers in a lossy coupled-cavity array, and make the quantum state stable at the receiver via the photon loss of optical cavities. Owing to the assistance of the dissipation, the target state becomes the steady state of the whole process. Thus there is no requirement on external time-dependent controls. Furthermore, the atomic spontaneous emission can be significantly suppressed as the adiabatic elimination of the excited states. We also discuss the experimental feasibility of this scheme with the current experimental technologies and a high fidelity of the transferred state in the receiver can be above 98%.

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

1. Introduction

High fidelity quantum state transfer (QST) has always been the subject of intensive researches in the context of quantum information processing (QIP) [1,2]. It requires that an arbitrary quantum state can be coherently transferred from an “emitter” system to a “receiver” system. In order to realize a reliable QST, considerable schemes appear with various quantum data buses, such as optical photons [36], phonon modes for trapped ions [7,8], spin chains for spin qubits [913], cavity photon modes for superconducting qubits and spin qubits [1416], and so on.

It is well known that a major practical hurdle in achieving a perfect QST is the quantum dissipation since the inevitable interaction between an open system and its surrounding environment. To avoid the quantum dissipation destroying the desired coherence of the transferred quantum state, an enormous amount of efforts has been made by using either dynamical decoupling [17,18], decoherence-free subspaces [19,20], or quantum error corrections [2123]. Nevertheless, the above methods have to fight against the quantum dissipation. And they are all based on the unitary dynamics, where the precise designs on the time evolution are demanded to ensure the target state is obtained at the end of an experiment.

With the development of the quantum information, more and more quantum technologies, from the dissipative entanglement preparation [2444] to the autonomous quantum error correction [4557], have exploited the spectacular promise of the quantum dissipation to perform different QIP tasks rather than exhaustively resisting the quantum dissipation. However, there is no QST scheme regarding the dissipation as a useful resource until Wang and Gertler [58] recently proposed to dissipatively achieve QST without time-dependent controls. Their proposals were designed in superconducting circuit quantum electrodynamics, but the interaction term engineered by the four-wave mixing Hamiltonian was too weak to be obtained. And the logical states were encoded by three-atom superposition states, which caused a waste of quantum resource. Subsequently, Matsuzaki et al. [59] put forward another dissipative QST scheme with two qubits collectively coupled to a tailored cavity, where a rigorous condition that the two qubits owned different resonance frequencies was required. Inspired by these schemes, our group [60] executed a directional QST in a dissipative Rydberg-atom-cavity system. Although the system in [60] was not as special as the previous two literatures, it still contained five classical lasers and two seven-level Rydberg atoms and the Rydberg-mediated interaction will also increase the experimental difficulties. Therefore, we attempt to find a more simple and general model to realize a dissipative QST scheme.

In this work, we consider a dissipation-assisted scheme to achieve a high fidelity QST for an arbitrary quantum state. The system consists of three classical lasers and two identical atoms respectively trapped in an array of lossy coupled cavities. It will transfer an arbitrary quantum state from the first atom to the second atom. And utilizing the photon loss of the cavities, the transferred quantum state will be stabilized at the second atom without special controls on the evolution of the system. The adverse effect of the atomic spontaneous emission can be neglected as the adiabatic elimination of the excited states. Besides, the atoms trapped into two separated cavities are easier to operate during an experiment. We also investigate the experimental feasibility with the current experimental parameters and a high fidelity of the target state can be above $98\%$.

2. Original physical model

In Fig. 1, the model and the corresponding atomic energy levels have been illustrated. There are two identical four-level atoms trapped in a coupled-cavity array, and the atoms both consist of three ground states $|0\rangle ,|e\rangle ,|1\rangle$ and one excited state $|r\rangle$. The transitions $|e\rangle \leftrightarrow |r\rangle$ of the two atoms are respectively driven by one classical laser with Rabi frequency $\Omega$ and detuning $\Delta _1$. And the transitions $|0\rangle \leftrightarrow |r\rangle$ and $|1\rangle \leftrightarrow |r\rangle$ for the atom $1(2)$ are coupled to two cavity modes $a_{1(2)L}$ and $a_{1(2)R}$ with coupling strength $g_L,g_R$, detuning $\delta$. Moreover, the atom $1$ interacts with another classical laser (Rabi frequency $\omega$, detuning $\Delta _2$) to pump the ground state $|e\rangle$ to the excited state $|r\rangle$. And the cavity-cavity coupling is described by the photon-hopping strength $J$. In the interaction picture, the total Hamiltonian reads

$$\begin{aligned} H_I= & \sum_{j=1,2}g(|r\rangle_{jj}\langle0|a_{jL}+|r\rangle_{jj}\langle1|a_{jR})e^{i\delta t}+\Omega|r\rangle_{jj}\langle e|e^{i\Delta_1 t}+\omega|r\rangle_{11}\langle e|e^{i\Delta_2 t}\\ & +J(a_{2L}^\dagger a_{1L}+a_{2R}^\dagger a_{1R})+\textrm{H.c.}, \end{aligned}$$
where we set $g_L=g_R=g$ for simplicity and the states $|0\rangle$ and $|1\rangle$ are used as encoded quantum bits. Combining the photon loss of the cavities that can be described by $L_{jL(R)}=\sqrt {\kappa }a_{jL(R)}$, the evolution of the total system will be governed by the master equation
$$\dot\rho={-}i[H_I,\rho]+\mathcal{L}\rho,$$
where
$$\mathcal{L}\rho=\sum_{k=L,R}\sum_{j=1,2}L_{jk}\rho L_{jk}^\dagger-\frac{1}{2}(L_{jk}^\dagger L_{jk}\rho+\rho L_{jk}^\dagger L_{jk}),$$
and $\kappa$ is the rate of the photon loss.

 figure: Fig. 1.

Fig. 1. (a) Model for the one-way QST scheme. There are two identical four-level atoms trapped in a coupled-cavity array. The first atom interacts with two lasers and the second atom couples to one laser. (b) The diagram of corresponding atomic energy levels.

Download Full Size | PDF

Then we introduce delocalized bosonic modes,

$$\begin{aligned}b_{{\pm} L}=\frac{1}{\sqrt{2}}(a_{1L}\pm a_{2L}) \end{aligned}$$
$$\begin{aligned}b_{{\pm} R}=\frac{1}{\sqrt{2}}(a_{1R}\pm a_{2R}) \end{aligned}$$
to rewrite the Hamiltonian in a rotating frame with respect to $U_0=\sum _{k=L,R}-itJ(b^{\dagger }_{+k}b_{+k}-b^{\dagger }_{-k}b_{-k})$
$$\begin{aligned} H_I&=\frac{g}{\sqrt{2}}(|r\rangle_{11}\langle0|b_{{+}L}+|r\rangle_{11}\langle1|b_{{+}R}+|r\rangle_{22}\langle0|b_{{+}L}+|r\rangle_{22}\langle1|b_{{+}R})e^{{-}i(J-\delta) t}\\ &\quad +\frac{g}{\sqrt{2}}(|r\rangle_{11}\langle0|b_{{-}L}+|r\rangle_{11}\langle1|b_{{-}R}-|r\rangle_{11}\langle0|b_{{-}L}-|r\rangle_{11}\langle1|b_{{-}R})e^{i(J+\delta) t}\\ &\quad +\Omega|r\rangle_{jj}\langle e|e^{i\Delta_1 t}+\omega|r\rangle_{11}\langle e|e^{i\Delta_2 t}+\textrm{H.c.}. \end{aligned}$$
After assuming $\Delta _1=J-\delta$ and $\Delta _2=J+\delta$, the corresponding Hamiltonian can be divided as
$$H_I=H_{s}+H_{d},$$
with
$$\begin{aligned} H_s&=\sum_{j=1,2}\frac{g}{\sqrt{2}}(|r\rangle_{jj}\langle 0|b_{{+}L}+|r\rangle_{jj}\langle1|b_{{+}R})e^{{-}i\Delta_1 t}+\Omega|r\rangle_{jj}\langle e|e^{i\Delta_1 t}+\textrm{H.c.},\\ H_d&=\sum_{j=1,2}({-}1)^{j+1}\frac{g}{\sqrt{2}}(|r\rangle_{jj}\langle 0|b_{{-}L}+|r\rangle_{jj}\langle1|b_{{-}R})e^{i\Delta_2 t}+\omega|r\rangle_{11}\langle e|e^{i\Delta_2 t}+\textrm{H.c.}. \end{aligned}$$
Here, $H_s$ aims to transfer an arbitrary quantum state $|\psi \rangle =\alpha |0\rangle +\beta |1\rangle$ ($|\alpha |^2+|\beta |^2=1$) from atom $1$ to atom $2$. At the same time, $H_d$ cooperating with the photon loss of the cavity stabilizes the quantum state $|\psi \rangle$ at the atom $2$ and the total QST scheme is finished.

3. Elaboration of effective master equation

First, we interpret the purpose of the $H_s$ in detail. The initial state of the two atoms is selected as $|\psi \rangle _1|e\rangle _2$ and all the cavity modes are initialized at vacuum states. In the large detuning condition $\Delta _2\gg g$, the Hamiltonian $H_d$ will be ineffective on the initial state while the Hamiltonian $H_s$ of Eq. (6) can be reformulated as

$$H_s=H_\Omega+H_g,$$
with
$$\begin{aligned} H_\Omega&=\Omega(|0e\rangle\langle 0r|+|r0\rangle\langle e0|+|1e\rangle\langle1r|+|r1\rangle\langle e1|)\otimes|\textbf{0}\rangle_{cc}\langle\textbf{0}|+\textrm{H.c.},\\ H_g&=\frac{g}{\sqrt{2}}(|0r\rangle\langle00|+|r0\rangle\langle00|+|1r\rangle\langle10|+|r1\rangle\langle01|)\otimes|0\rangle_{{+}L,+L}\langle1|+\frac{g}{\sqrt{2}}(|0r\rangle\langle01|+|r0\rangle\langle10|\\ &\quad +|1r\rangle\langle11|+|r1\rangle\langle11|)\otimes|0\rangle_{{+}R,+R}\langle1|+\textrm{H.c.}+\Delta_1(|0r\rangle\langle0r|+|r0\rangle\langle r0|+|1r\rangle\langle1r|\\ &\quad +|r1\rangle\langle r1|)\otimes|\textbf{0}\rangle_{cc}\langle\textbf{0}|+2\Delta_1(|00\rangle\langle00|+|01\rangle\langle01|+|10\rangle\langle10|)\otimes|1\rangle_{{+}L,+L}\langle1|+2\Delta_1(|11\rangle\langle11|\\ &\quad +|01\rangle\langle01|+|10\rangle\langle10|)\otimes|1\rangle_{{+}R,+R}\langle1|, \end{aligned}$$
where the state $|n\rangle _{m}$ ($n=0,1$ and $m=\pm L,\pm R$) describes only the cavity mode $b_{m}$ at state $|n\rangle$ and the other cavity modes at vacuum state. And $|\textbf {0}\rangle _c$ stands for all the cavity modes at the vacuum state. Then we can rewrite the Eq. (7) by diagonalizing the $H_g$,
$$\begin{aligned} H_s~=~~ & \Omega_M^-[(|e0\rangle+|0e\rangle)|\textbf{0}\rangle_{c}\langle T_1|+(|e1\rangle+|1e\rangle)|\textbf{0}\rangle_c\langle S_1|]+\Omega_K^+[(|e0\rangle-|0e\rangle)|\textbf{0}\rangle_c\langle T_2|\\ & +(|e1\rangle-|1e\rangle)|\textbf{0}\rangle_c\langle S_2|]+\Omega_K^-[(|e0\rangle-|0e\rangle)|\textbf{0}\rangle_c\langle T_4|+(|e1\rangle-|1e\rangle)|\textbf{0}\rangle_c\langle S_4|]\\ & +\Omega_M^+[(|e0\rangle+|0e\rangle)|\textbf{0}\rangle_c\langle T_5|+(|e1\rangle+|1e\rangle)|\textbf{0}\rangle_c\langle S_5|]+\textrm{H.c.}+\sum_{i=1}^5E_i(|T_i\rangle\langle T_i|\\ & +|S_i\rangle\langle S_i|), \end{aligned}$$
where $\Omega _M^\pm =\Omega \sqrt {1\pm \Delta _1/M}/2$, $\Omega _K^\pm =g\Omega /\sqrt {2K(K\pm \Delta _1)}$, $M=\sqrt {6g^2+\Delta _1^2}$, and $K=\sqrt {2g^2+\Delta _1^2}$. The states $|T_i\rangle$ and $|S_i\rangle$ are the eigenvectors of the $H_g$ with respect to the eigenvalue $E_i$ $[E1=(3\Delta _1+M)/2,E_2=(3\Delta _1+K)/2,E_3=2\Delta _1,E_4=(3\Delta _1-K)/2,E_5=(3\Delta _1-M)/2]$. The concrete form of $|T_i\rangle$ can be shown as
$$\begin{aligned} |T_1\rangle&=\frac{1}{2}\sqrt{1-\frac{\Delta_1}{M}}(|r0\rangle+|0r\rangle)|\textbf{0}\rangle_c+\frac{g}{\sqrt{2M(M-\Delta_1)}}(2|00\rangle|1\rangle_{{+}L}+|01\rangle|1\rangle_{{+}R}+|10\rangle|1\rangle_{{+}R}),\\ |T_2\rangle&=\frac{g}{\sqrt{2K(K+\Delta_1)}}(|r0\rangle-|0r\rangle)|\textbf{0}\rangle_c+\frac{1}{2}\sqrt{1+\frac{\Delta_1}{K}}(|10\rangle|1\rangle_{{+}R}-|01\rangle|1\rangle_{{+}R}),\\ |T_3\rangle&=\frac{1}{\sqrt{3}}(|10\rangle|1\rangle_{{+}R}+|01\rangle|1\rangle_{{+}R}-|00\rangle|1\rangle_{{+}L}),\\ |T_4\rangle&=\frac{g}{\sqrt{2K(K-\Delta_1)}}(|r0\rangle-|0r\rangle)|\textbf{0}\rangle_c+\frac{1}{2}\sqrt{1-\frac{\Delta_1}{K}}(|01\rangle|1\rangle_{{+}R}-|10\rangle|1\rangle_{{+}R}),\\ |T_5\rangle&=\frac{1}{2}\sqrt{1+\frac{\Delta_1}{M}}(|r0\rangle+|0r\rangle)|\textbf{0}\rangle_c-\frac{g}{\sqrt{2M(M+\Delta_1)}}(2|00\rangle|1\rangle_{{+}L}+|01\rangle|1\rangle_{{+}R}+|10\rangle|1\rangle_{{+}R}). \end{aligned}$$
And the form of $|S_i\rangle$ can be obtained by successively replacing $|00\rangle |1\rangle _{+L},~|0r\rangle |\textbf {0}\rangle _c,~|r0\rangle |\textbf {0}\rangle _c,~|01\rangle |1\rangle _{+R}$, and $|10\rangle |1\rangle _{+R}$ of $|T_i\rangle$ with $|11\rangle |1\rangle _{+R},~|1r\rangle |\textbf {0}\rangle _c,~|r1\rangle |\textbf {0}\rangle _c,~|10\rangle |1\rangle _{+L}$, and $|01\rangle |1\rangle _{+L}$.

Furthermore, while $1\gg \max \{\Omega ^-_M/E_1,\Omega ^+_K/E_2,\Omega ^-_K/E_4,\Omega ^+_M/E_5\}$, we can adiabatically eliminate the excited states of the Eq. (8). The corresponding Hamiltonian is

$$\begin{aligned} H_s^\textrm{eff}~=~~ & \tilde\Omega(|e0\rangle\langle 0e|+|e1\rangle\langle 1e|)\otimes|\textbf{0}\rangle_{cc}\langle\textbf{0}|+\textrm{H.c.}-\tilde\Delta(|0e\rangle\langle0e|+|e0\rangle\langle e0|+|1e\rangle\langle1e|\\ & +|e1\rangle\langle e1|)\otimes|\textbf{0}\rangle_{cc}\langle\textbf{0}|. \end{aligned}$$
Here we have abbreviated $\tilde \Omega =(\Omega ^{+2}_K/E_2+\Omega ^{-2}_K/E_4-\Omega ^{-2}_M/E_1-\Omega ^{+2}_M/E_5)$ and $\tilde \Delta =(\Omega ^{-2}_M/E_1+\Omega ^{+2}_K/E_2+\Omega ^{-2}_K/E_4+\Omega ^{+2}_M/E_5)$, and the terms of $\tilde \Delta$ stand for the Stark shifts. Moreover, we have ignored the term $E_3(|T_3\rangle \langle T_3|+|S_3\rangle \langle S_3|)$ because the state $|T(S)_3\rangle$ will never be populated during the whole process.

According to the Eq. (9), $H_s$ can swap an arbitrary quantum state between the two atoms, i.e., $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c\leftrightarrow |e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$. Particularly, the adverse impacts of dissipation in the swap process are significantly suppressed as the cavity states of the swap process are restrained into the vacuum state and the excited states are adiabatically eliminated.

In Fig. 2, we show the evolution of the fidelity for the states $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ and $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ in the swap process, where the measure of fidelity for state $|i\rangle$ uses the definition Tr${\bigg [}\rho _i^{1/2}\rho \rho _i^{1/2}{\bigg ]}^{1/2}$ and $\rho _i=|i\rangle \langle i|$. The solid line and the empty circles describe the fidelity of the state $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ simulated by $H_s$ and $H_s^\textrm {eff}$, respectively. And the appearance of the former in accord with that of the latter proves the validity of the derivation of the $H_s^\textrm {eff}$. Besides, the periodical oscillations of the solid and dashed lines exactly express the swap process $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c\leftrightarrow |e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$, where the dashed line is the fidelity of the state $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ governed by the original Hamiltonian $H_s$.

 figure: Fig. 2.

Fig. 2. The evolution of the fidelity for the states $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ and $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ in the swap process. The solid line and the empty circles describe the fidelity of the state $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ simulated by $H_s$ and $H_s^\textrm {eff}$, respectively. And the dashed line stands for the fidelity of the state $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ governed by the Hamiltonian $H_s$. The transferred quantum state $|\psi \rangle$ is randomly selected as $|\psi \rangle =\alpha |0\rangle +\beta |1\rangle$ with $\alpha =\sin \theta$, $\beta =\cos \theta$, and $\theta =\pi /3$. And the initial state is $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$. The relevant parameters are chosen as: $J=161g$, $\delta =160g$, and $\Omega =0.02g$.

Download Full Size | PDF

On the other hand, once the system evolves into $|e0\rangle |\textbf {0}\rangle _c$ or $|e1\rangle |\textbf {0}\rangle _c$ by the swap process, the Hamiltonian $H_d$ will be in action and it can be expanded as

$$\begin{aligned} H_d~=~~ & \Big(\omega|e\rangle_{11}\langle r|\otimes|\textbf{0}\rangle_{cc}\langle\textbf{0}|+\frac{g}{\sqrt{2}}|1\rangle_{11}\langle r|\otimes|1\rangle_{{-}R,-R}\langle0|+\frac{g}{\sqrt{2}}|0\rangle_{11}\langle r|\otimes|1\rangle_{{-}L,-L}\langle0|\Big)\\ & \otimes(|0\rangle_{22}\langle0|+|1\rangle_{22}\langle1|)e^{{-}i\Delta_2t}+\textrm{H.c.}. \end{aligned}$$
It is significant that the $H_d$ will not change the transferred state $|\psi \rangle$ in atom 2. On account of the large detuning condition $\Delta _2\gg \{\omega ,g\}$, the Hamiltonian can be further simplified as
$$\begin{aligned} H_d^\textrm{eff}~=~~ & -\frac{g\omega}{\sqrt{2}\Delta_2}(|e\rangle_{11}\langle 1|\otimes|0\rangle_{{-}R,-R}\langle1|+|e\rangle_{11}\langle 0|\otimes|0\rangle_{{-}L,-L}\langle1|)\otimes(|0\rangle_{22}\langle0|+|1\rangle_{22}\langle1|)\\ & -\frac{g^2}{2\Delta_2}|1\rangle_{11}\langle0|\otimes|1\rangle_{{-}R,-L}\langle1|\otimes(|0\rangle_{22}\langle0|+|1\rangle_{22}\langle1|)+\textrm{H.c.,} \end{aligned}$$
where the Stark-shift terms are omitted since they can be compensated by the auxiliary levels.

Referring to the Eq. (11), the state $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ can evolve into the states $|0\rangle _1|\psi \rangle _2|1\rangle _{-L}$ and $|1\rangle _1|\psi \rangle _2|1\rangle _{-R}$. And these two states will be respectively stabilized at the steady states of the whole process, $|0\rangle _1|\psi \rangle _2|\textbf {0}\rangle _{c}$ and $|1\rangle _1|\psi \rangle _2|\textbf {0}\rangle _{c}$, via the photon loss of the cavities. Then the QST is finished.

In addition, we can describe the total dissipative dynamics with the effective master equation that can be acquired by replacing the $H_I$ of the Eq. (2) with $H_\textrm {eff}=H_s^\textrm {eff}+H_d^\textrm {eff}$. And the effective process certifies that our scheme successfully transfers an arbitrary quantum state $|\psi \rangle$ from atom 1 to atom 2 and stabilizes $|\psi \rangle$ at atom 2 by the ingenious integration of $H_s$, $H_d$, and the cavities decay.

4. Numerical results

In Fig. 3, we plot the dynamical evolution for the fidelity of the transferred state $|\psi \rangle$ in atom 2, $F=\textrm {Tr}{\bigg [}\rho _s^{1/2}\textrm {Tr}_{1\& c}(\rho )\rho _s^{1/2}{\bigg ]}^{1/2}$, to demonstrate the validity of the above analyses, where $\rho _s=|\psi \rangle \langle \psi |$ and $\textrm {Tr}_{1\& c}$ means the partial trace over atom 1 and all cavity modes. The parameters of the transferred state $|\psi \rangle$ are randomly chosen as $\alpha =\sin \theta$, $\beta =\cos \theta$, and $\theta =\pi /3$. The solid line and the empty circles respectively represent the dynamical evolution of $F$ governed by the original master equation and the effective master equation. The brilliant agreement of the two curves adequately proves the validity of the reduced system, which can help us to forecast and interpret the behaviors of the original system. Moreover, the fidelity of $|\psi \rangle$ in atom 2 is as great as $98.63\%$ at $gt=12000$, and the fidelity of the initial state $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ (dashed line) tends to be vanished as time goes by. These results also confirm the feasibility of our scheme, i.e., our scheme successfully realizes that an arbitrary quantum state is dissipatively transferred from an “emitter” system to a “receiver” system and stabilized at the “receiver” system without time-dependent controls.

 figure: Fig. 3.

Fig. 3. The fidelity as functions of $gt$ governed by the full master equation and the effective master equation, where $\rho _s=|\psi \rangle \langle \psi |$ with $|\psi \rangle =\alpha |0\rangle +\beta |1\rangle$, $\alpha =\sin \theta$, $\beta =\cos \theta$, and $\theta =\pi /3$. The initial state is selected as $\rho _0=|\psi \rangle _1\langle \psi |\otimes |e\rangle _2\langle e|\otimes |\textbf {0}\rangle _c\langle \textbf {0}|$. The other corresponding parameters are chosen as: $J=161g$, $\delta =160g$, $\Omega =0.02g$, $\omega =g$, and $\kappa =0.02g$.

Download Full Size | PDF

Finally, we investigate the experimental feasibility of our scheme. In [61], Spillane et al. discussed the suitability of toroidal microcavities for strong-coupling cavity quantum electrodynamics. And they reported the projected limit for a Fabry-Perot cavity that the coupling coefficient can arrive at $g/2\pi =770$ MHz. In terms of the corresponding critical photon number and critical atom number, one can obtain the rates of cavity decay and atomic spontaneous emission as $(\kappa ,\gamma )/2\pi =(21.7,2.6)$ MHz. Besides, referring to the experiment of [6264], we can also suppose that a cavity made of two niobium superconducting mirrors ($g/2\pi =50$ kHz and $\kappa /2\pi =1$ kHz) traps Rubidium atoms with $\gamma /2\pi =0.03$ kHz. Moreover, another group of experimental parameters can be obtained in [65] with $(g,\kappa ,\gamma )=2\pi \times (150,6.25,0.03)$ kHz. And we characterize the atomic spontaneous emission of our scheme by the Lindblad operators $L_{1(2)}=\sqrt {\gamma /3}|0\rangle _{1(2)}\langle r|$, $L_{3(4)}=\sqrt {\gamma /3}|e\rangle _{1(2)}\langle r|$, and $L_{5(6)}=\sqrt {\gamma /3}|1\rangle _{1(2)}\langle r|$. Another key element for our scheme is the photon-hopping strength $J$. Although the analytical expressions of $J$ are different as the types of cavities change [6669], a common important dependence is the distance between the cavities. Particularly, in [67], Underwood et al. experimentally investigated the suitability of coupled-transmission-line resonators for studies of quantum phase transitions of light. The hopping rate between resonators $i$ and $j$ is considered as $J=\nu C_c|\phi _i(x_i)\phi _j(x_j)|/2$, where $\nu$ stands for the resonator frequency, $C_c$ is the coupling capacitance, and $\phi _i(x_i)$ denotes the classical mode function of resonator $i$. Additionally, in [68], Shen et al. realized an optical diode for photonic transport with two spatially overlapping single-mode semiconductor microcavities. The coupling strength $J=|\sqrt {\omega _a\omega _b}\int \varepsilon _{ij}(\boldsymbol {r})\phi _{i,a}(\boldsymbol {r})\phi _{j,b}^\ast (\boldsymbol {r})/\varepsilon (\boldsymbol {r})d^3\boldsymbol {r}|$, where $\varepsilon _{ij}(\boldsymbol {r})$ is the relative dielectric permittivity tensor of the medium, $\varepsilon (\boldsymbol {r})$ is a spatially dependent scalar quantity, the frequencies of the two cavities are denoted by $\omega _a$ and $\omega _b$, and $\phi _{i,a}(\boldsymbol {r})$ is the $i$-direction ($i=x,y,z$) space wave function depending on the position $\boldsymbol {r}$ for cavity $a$. Nevertheless, it should be noted that the conclusion of our scheme is independent of the analytical expressions of $J$ and only the value of the $J$ at a given fixed distance of the two cavities is relevant. According to the experimental schemes [67,70,71], we can find that the distance between different sorts of coupled cavities can be tuned from 200 nm to 40 $\mu$m with a wide range of the coupled strength ($\sim$THz-MHz), which provides the possibility for us to find a suitable value of $J$.

Using the above groups of parameters, we numerically calculate the dynamical evolution for the fidelity of the transferred state $|\psi \rangle$ at atom 2 in Fig. 4, and the results can be stable at $98.33\%$ (empty circles), $98.45\%$ (empty squares), and $98.56\%$ (empty triangles). These consequences thoroughly demonstrate the experimental feasibility of our scheme. In addition, the inset of Fig. 4 depicts the evolution of the fidelity for the transferred states with different $\theta$, where the tendencies of these curves are in full accord with each others. It faithfully reflects that our scheme is achievable with an arbitrary quantum state.

 figure: Fig. 4.

Fig. 4. The dynamical evolution for the fidelity of the transferred state $|\psi \rangle$ in atom 2 with experimental parameters. The other corresponding parameters are chosen as: $J=161g$, $\delta =160g$, $\Omega =0.02g$, $\omega =g$, and $\theta =\pi /3$. The inset shows the corresponding fidelity as function of $gt$ with different $\theta$, where the parameters are $g=2\pi \times 770$ MHz, $\kappa =21.7$ MHz, $\gamma =2.6$ MHz, $J=161g$, $\delta =160g$, $\Omega =0.02g$, and $\omega =g$.

Download Full Size | PDF

5. Summary

In summary, we successfully implement a one-way QST scheme in a lossy coupled-cavity array. Due to the photon loss of the cavities, an arbitrary quantum state can be not only transferred from an “emitter” system to a “receiver” system, but also stabilized at the latter without external time-dependent control. Moreover, the atomic spontaneous emission can be dramatically inhibited since the adiabatic elimination of the excited states with suitable conditions. In addition, compared with the previous known proposals, the present scheme is easier to be actualized in an experiment. We also discuss the feasibility of the scheme with the current experimental parameter and a fidelity for the transferred state at the “receiver” system can be above $98\%$. We believe the present scheme supplies a viable prospect for QST scheme via dissipation.

Funding

National Natural Science Foundation of China (11774047).

References

1. D. P. DiVincenzo, “The physical implementation of quantum computation,” Fortschr. Phys. 48(9-11), 771–783 (2000). [CrossRef]  

2. H. J. Kimble, “The quantum internet,” Nature 453(7198), 1023–1030 (2008). [CrossRef]  

3. C. Cabrillo, J. I. Cirac, P. García-Fernández, and P. Zoller, “Creation of entangled states of distant atoms by interference,” Phys. Rev. A 59(2), 1025–1033 (1999). [CrossRef]  

4. S. Bose, P. L. Knight, M. B. Plenio, and V. Vedral, “Proposal for teleportation of an atomic state via cavity decay,” Phys. Rev. Lett. 83(24), 5158–5161 (1999). [CrossRef]  

5. Y. L. Lim, A. Beige, and L. C. Kwek, “Repeat-until-success linear optics distributed quantum computing,” Phys. Rev. Lett. 95(3), 030505 (2005). [CrossRef]  

6. D. N. Matsukevich, P. Maunz, D. L. Moehring, S. Olmschenk, and C. Monroe, “Bell inequality violation with two remote atomic qubits,” Phys. Rev. Lett. 100(15), 150404 (2008). [CrossRef]  

7. J. I. Cirac and P. Zoller, “Quantum computations with cold trapped ions,” Phys. Rev. Lett. 74(20), 4091–4094 (1995). [CrossRef]  

8. R. Blatt and D. Wineland, “Entangled states of trapped atomic ions,” Nature 453(7198), 1008–1015 (2008). [CrossRef]  

9. S. Bose, “Quantum communication through an unmodulated spin chain,” Phys. Rev. Lett. 91(20), 207901 (2003). [CrossRef]  

10. T. Shi, Y. Li, Z. Song, and C.-P. Sun, “Quantum-state transfer via the ferromagnetic chain in a spatially modulated field,” Phys. Rev. A 71(3), 032309 (2005). [CrossRef]  

11. A. Lyakhov and C. Bruder, “Quantum state transfer in arrays of flux qubits,” New J. Phys. 7, 181 (2005). [CrossRef]  

12. M. Friesen, A. Biswas, X. Hu, and D. Lidar, “Efficient multiqubit entanglement via a spin bus,” Phys. Rev. Lett. 98(23), 230503 (2007). [CrossRef]  

13. F. A. Mohiyaddin, R. Kalra, A. Laucht, R. Rahman, G. Klimeck, and A. Morello, “Transport of spin qubits with donor chains under realistic experimental conditions,” Phys. Rev. B 94(4), 045314 (2016). [CrossRef]  

14. A. Imamoḡlu, D. D. Awschalom, G. Burkard, D. P. DiVincenzo, D. Loss, M. Sherwin, and A. Small, “Quantum information processing using quantum dot spins and cavity qed,” Phys. Rev. Lett. 83(20), 4204–4207 (1999). [CrossRef]  

15. A. A. Houck, D. I. Schuster, J. M. Gambetta, J. A. Schreier, B. R. Johnson, J. M. Chow, L. Frunzio, J. Majer, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, “Generating single microwave photons in a circuit,” Nature 449(7160), 328–331 (2007). [CrossRef]  

16. J. Majer, J. M. Chow, J. M. Gambetta, J. Koch, B. R. Johnson, J. A. Schreier, L. Frunzio, D. I. Schuster, A. A. Houck, A. Wallraff, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, “Coupling superconducting qubits via a cavity bus,” Nature 449(7161), 443–447 (2007). [CrossRef]  

17. L. Viola, E. Knill, and S. Lloyd, “Dynamical decoupling of open quantum systems,” Phys. Rev. Lett. 82(12), 2417–2421 (1999). [CrossRef]  

18. G. Xu and G. Long, “Protecting geometric gates by dynamical decoupling,” Phys. Rev. A 90(2), 022323 (2014). [CrossRef]  

19. L.-M. Duan and G.-C. Guo, “Preserving coherence in quantum computation by pairing quantum bits,” Phys. Rev. Lett. 79(10), 1953–1956 (1997). [CrossRef]  

20. W. Qin, C. Wang, and X. Zhang, “Protected quantum-state transfer in decoherence-free subspaces,” Phys. Rev. A 91(4), 042303 (2015). [CrossRef]  

21. A. Y. Kitaev, “Quantum computations: algorithms and error correction,” Russ. Math. Surv. 52(6), 1191–1249 (1997). [CrossRef]  

22. B. M. Terhal and G. Burkard, “Fault-tolerant quantum computation for local non-markovian noise,” Phys. Rev. A 71(1), 012336 (2005). [CrossRef]  

23. D. Aharonov, A. Kitaev, and J. Preskill, “Fault-tolerant quantum computation with long-range correlated noise,” Phys. Rev. Lett. 96(5), 050504 (2006). [CrossRef]  

24. M. B. Plenio, S. F. Huelga, A. Beige, and P. L. Knight, “Cavity-loss-induced generation of entangled atoms,” Phys. Rev. A 59(3), 2468–2475 (1999). [CrossRef]  

25. M. B. Plenio and S. F. Huelga, “Entangled light from white noise,” Phys. Rev. Lett. 88(19), 197901 (2002). [CrossRef]  

26. X. X. Yi, C. S. Yu, L. Zhou, and H. S. Song, “Noise-assisted preparation of entangled atoms,” Phys. Rev. A 68(5), 052304 (2003). [CrossRef]  

27. S. Pielawa, G. Morigi, D. Vitali, and L. Davidovich, “Generation of einstein-podolsky-rosen-entangled radiation through an atomic reservoir,” Phys. Rev. Lett. 98(24), 240401 (2007). [CrossRef]  

28. G. Vacanti and A. Beige, “Cooling atoms into entangled states,” New J. Phys. 11(8), 083008 (2009). [CrossRef]  

29. M. J. Kastoryano, F. Reiter, and A. S. Sørensen, “Dissipative preparation of entanglement in optical cavities,” Phys. Rev. Lett. 106(9), 090502 (2011). [CrossRef]  

30. L. T. Shen, X. Y. Chen, Z. B. Yang, H. Z. Wu, and S. B. Zheng, “Steady-state entanglement for distant atoms by dissipation in coupled cavities,” Phys. Rev. A 84(6), 064302 (2011). [CrossRef]  

31. A. W. Carr and M. Saffman, “Preparation of entangled and antiferromagnetic states by dissipative rydberg pumping,” Phys. Rev. Lett. 111(3), 033607 (2013). [CrossRef]  

32. Y. Lin, J. P. Gaebler, F. Reiter, T. R. Tan, R. Bowler, A. S. Sørensen, D. Leibfried, and D. J. Wineland, “Dissipative production of a maximally entangled steady state of two quantum bits,” Nature 504(7480), 415–418 (2013). [CrossRef]  

33. S. Shankar, M. Hatridge, Z. Leghtas, K. M. Sliwa, A. Narla, U. Vool, S. M. Girvin, L. Frunzio, M. Mirrahimi, and M. H. Devoret, “Autonomously stabilized entanglement between two superconducting quantum bits,” Nature 504(7480), 419–422 (2013). [CrossRef]  

34. F. Reiter, D. Reeb, and A. S. Sørensen, “Scalable dissipative preparation of many-body entanglement,” Phys. Rev. Lett. 117(4), 040501 (2016). [CrossRef]  

35. X. Q. Shao, J. H. Wu, and X. X. Yi, “Dissipative stabilization of quantum-feedback-based multipartite entanglement with rydberg atoms,” Phys. Rev. A 95(2), 022317 (2017). [CrossRef]  

36. X. Q. Shao, J. H. Wu, and X. X. Yi, “Dissipation-based entanglement via quantum zeno dynamics and rydberg antiblockade,” Phys. Rev. A 95(6), 062339 (2017). [CrossRef]  

37. X. Q. Shao, J. H. Wu, X. X. Yi, and G.-L. Long, “Dissipative preparation of steady greenberger-horne-zeilinger states for rydberg atoms with quantum zeno dynamics,” Phys. Rev. A 96(6), 062315 (2017). [CrossRef]  

38. D. X. Li, X. Q. Shao, J. H. Wu, and X. X. Yi, “Engineering steady-state entanglement via dissipation and quantum zeno dynamics in an optical cavity,” Opt. Lett. 42(19), 3904–3907 (2017). [CrossRef]  

39. D.-X. Li, X.-Q. Shao, J.-H. Wu, and X. X. Yi, “Noise-induced distributed entanglement in atom-cavity-fiber system,” Opt. Express 25(26), 33359–33378 (2017). [CrossRef]  

40. Z. Jin, S.-L. Su, A.-D. Zhu, H.-F. Wang, and S. Zhang, “Dissipative preparation of distributed steady entanglement: an approach of unilateral qubit driving,” Opt. Express 25(1), 88–101 (2017). [CrossRef]  

41. X.-Q. Shao, “Engineering steady entanglement for trapped ions at finite temperature by dissipation,” Phys. Rev. A 98(4), 042310 (2018). [CrossRef]  

42. D.-X. Li, X.-Q. Shao, J.-H. Wu, and X. X. Yi, “Dissipation-induced w state in a rydberg-atom-cavity system,” Opt. Lett. 43(8), 1639–1642 (2018). [CrossRef]  

43. D.-X. Li, X.-Q. Shao, J.-H. Wu, X. X. Yi, and T.-Y. Zheng, “Engineering steady knill-laflamme-milburn state of rydberg atoms by dissipation,” Opt. Express 26(3), 2292–2302 (2018). [CrossRef]  

44. C. Yang, D. Li, and X. Shao, “Dissipative preparation of bell states with parallel quantum zeno dynamics,” Sci. China: Phys., Mech. Astron. 62(11), 110312 (2019). [CrossRef]  

45. J. P. Paz and W. H. Zurek, “Continuous error correction,” Proc. R. Soc. London, Ser. A 454(1969), 355–364 (1998). [CrossRef]  

46. C. Ahn, A. C. Doherty, and A. J. Landahl, “Continuous quantum error correction via quantum feedback control,” Phys. Rev. A 65(4), 042301 (2002). [CrossRef]  

47. J. Kerckhoff, H. I. Nurdin, D. S. Pavlichin, and H. Mabuchi, “Designing quantum memories with embedded control: Photonic circuits for autonomous quantum error correction,” Phys. Rev. Lett. 105(4), 040502 (2010). [CrossRef]  

48. K. Fujii, M. Negoro, N. Imoto, and M. Kitagawa, “Measurement-free topological protection using dissipative feedback,” Phys. Rev. X 4(4), 041039 (2014). [CrossRef]  

49. Z. Leghtas, S. Touzard, I. M. Pop, A. Kou, B. Vlastakis, A. Petrenko, K. M. Sliwa, A. Narla, S. Shankar, M. J. Hatridge, M. Reagor, L. Frunzio, R. J. Schoelkopf, M. Mirrahimi, and M. H. Devoret, “Confining the state of light to a quantum manifold by engineered two-photon loss,” Science 347(6224), 853–857 (2015). [CrossRef]  

50. M. Ippoliti, L. Mazza, M. Rizzi, and V. Giovannetti, “Perturbative approach to continuous-time quantum error correction,” Phys. Rev. A 91(4), 042322 (2015). [CrossRef]  

51. E. Kapit, J. T. Chalker, and S. H. Simon, “Passive correction of quantum logical errors in a driven, dissipative system: A blueprint for an analog quantum code fabric,” Phys. Rev. A 91(6), 062324 (2015). [CrossRef]  

52. E. Kapit, “Hardware-efficient and fully autonomous quantum error correction in superconducting circuits,” Phys. Rev. Lett. 116(15), 150501 (2016). [CrossRef]  

53. C. D. Freeman, C. M. Herdman, and K. B. Whaley, “Engineering autonomous error correction in stabilizer codes at finite temperature,” Phys. Rev. A 96(1), 012311 (2017). [CrossRef]  

54. C. Shen, K. Noh, V. V. Albert, S. Krastanov, M. H. Devoret, R. J. Schoelkopf, S. M. Girvin, and L. Jiang, “Quantum channel construction with circuit quantum electrodynamics,” Phys. Rev. B 95(13), 134501 (2017). [CrossRef]  

55. F. Reiter, A. S. Sørensen, P. Zoller, and C. A. Muschik, “Dissipative quantum error correction and application to quantum sensing with trapped ions,” Nat. Commun. 8(1), 1822 (2017). [CrossRef]  

56. J.-M. Lihm, K. Noh, and U. R. Fischer, “Implementation-independent sufficient condition of the knill-laflamme type for the autonomous protection of logical qudits by strong engineered dissipation,” Phys. Rev. A 98(1), 012317 (2018). [CrossRef]  

57. D. X. Li and X. Q. Shao, “Unconventional rydberg pumping and applications in quantum information processing,” Phys. Rev. A 98(6), 062338 (2018). [CrossRef]  

58. C. Wang and J. M. Gertler, “Autonomous quantum state transfer by dissipation,” ArXiv e-prints 1809.03571v2 (2018).

59. Y. Matsuzaki, V. M. Bastidas, Y. Takeuchi, W. J. Munro, and S. Saito, “One-way transfer of quantum states via decoherence,” ArXiv e-prints 1810.02995 (2018).

60. D. X. Li and X. Q. Shao, “Directional quantum state transfer in a dissipative rydberg-atom-cavity system,” Phys. Rev. A 99(3), 032348 (2019). [CrossRef]  

61. S. M. Spillane, T. J. Kippenberg, K. J. Vahala, K. W. Goh, E. Wilcut, and H. J. Kimble, “Ultrahigh-q toroidal microresonators for cavity quantum electrodynamics,” Phys. Rev. A 71(1), 013817 (2005). [CrossRef]  

62. M. Brune, F. Schmidt-Kaler, A. Maali, J. Dreyer, E. Hagley, J. M. Raimond, and S. Haroche, “Quantum rabi oscillation: A direct test of field quantization in a cavity,” Phys. Rev. Lett. 76(11), 1800–1803 (1996). [CrossRef]  

63. S. Osnaghi, P. Bertet, A. Auffeves, P. Maioli, M. Brune, J. M. Raimond, and S. Haroche, “Coherent control of an atomic collision in a cavity,” Phys. Rev. Lett. 87(3), 037902 (2001). [CrossRef]  

64. F. O. Prado, F. S. Luiz, J. M. Villas-Bôas, A. M. Alcalde, E. I. Duzzioni, and L. Sanz, “Atom-mediated effective interactions between modes of a bimodal cavity,” Phys. Rev. A 84(5), 053839 (2011). [CrossRef]  

65. C. Lazarou and B. M. Garraway, “Adiabatic entanglement in two-atom cavity qed,” Phys. Rev. A 77(2), 023818 (2008). [CrossRef]  

66. A. Nunnenkamp, J. Koch, and S. M. Girvin, “Synthetic gauge fields and homodyne transmission in jaynes–cummings lattices,” New J. Phys. 13(9), 095008 (2011). [CrossRef]  

67. D. L. Underwood, W. E. Shanks, J. Koch, and A. A. Houck, “Low-disorder microwave cavity lattices for quantum simulation with photons,” Phys. Rev. A 86(2), 023837 (2012). [CrossRef]  

68. H. Z. Shen, Y. H. Zhou, and X. X. Yi, “Quantum optical diode with semiconductor microcavities,” Phys. Rev. A 90(2), 023849 (2014). [CrossRef]  

69. H. Z. Shen, Y. H. Zhou, and X. X. Yi, “Tunable photon blockade in coupled semiconductor cavities,” Phys. Rev. A 91(6), 063808 (2015). [CrossRef]  

70. A. Majumdar, A. Rundquist, M. Bajcsy, V. D. Dasika, S. R. Bank, and J. Vučković, “Design and analysis of photonic crystal coupled cavity arrays for quantum simulation,” Phys. Rev. B 86(19), 195312 (2012). [CrossRef]  

71. Y. He, “Light storage and cavity supermodes in two coupled optomechanical cavities,” Phys. Rev. A 94(6), 063804 (2016). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. (a) Model for the one-way QST scheme. There are two identical four-level atoms trapped in a coupled-cavity array. The first atom interacts with two lasers and the second atom couples to one laser. (b) The diagram of corresponding atomic energy levels.
Fig. 2.
Fig. 2. The evolution of the fidelity for the states $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ and $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ in the swap process. The solid line and the empty circles describe the fidelity of the state $|e\rangle _1|\psi \rangle _2|\textbf {0}\rangle _c$ simulated by $H_s$ and $H_s^\textrm {eff}$, respectively. And the dashed line stands for the fidelity of the state $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$ governed by the Hamiltonian $H_s$. The transferred quantum state $|\psi \rangle$ is randomly selected as $|\psi \rangle =\alpha |0\rangle +\beta |1\rangle$ with $\alpha =\sin \theta$, $\beta =\cos \theta$, and $\theta =\pi /3$. And the initial state is $|\psi \rangle _1|e\rangle _2|\textbf {0}\rangle _c$. The relevant parameters are chosen as: $J=161g$, $\delta =160g$, and $\Omega =0.02g$.
Fig. 3.
Fig. 3. The fidelity as functions of $gt$ governed by the full master equation and the effective master equation, where $\rho _s=|\psi \rangle \langle \psi |$ with $|\psi \rangle =\alpha |0\rangle +\beta |1\rangle$, $\alpha =\sin \theta$, $\beta =\cos \theta$, and $\theta =\pi /3$. The initial state is selected as $\rho _0=|\psi \rangle _1\langle \psi |\otimes |e\rangle _2\langle e|\otimes |\textbf {0}\rangle _c\langle \textbf {0}|$. The other corresponding parameters are chosen as: $J=161g$, $\delta =160g$, $\Omega =0.02g$, $\omega =g$, and $\kappa =0.02g$.
Fig. 4.
Fig. 4. The dynamical evolution for the fidelity of the transferred state $|\psi \rangle$ in atom 2 with experimental parameters. The other corresponding parameters are chosen as: $J=161g$, $\delta =160g$, $\Omega =0.02g$, $\omega =g$, and $\theta =\pi /3$. The inset shows the corresponding fidelity as function of $gt$ with different $\theta$, where the parameters are $g=2\pi \times 770$ MHz, $\kappa =21.7$ MHz, $\gamma =2.6$ MHz, $J=161g$, $\delta =160g$, $\Omega =0.02g$, and $\omega =g$.

Equations (15)

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

H I = j = 1 , 2 g ( | r j j 0 | a j L + | r j j 1 | a j R ) e i δ t + Ω | r j j e | e i Δ 1 t + ω | r 11 e | e i Δ 2 t + J ( a 2 L a 1 L + a 2 R a 1 R ) + H.c. ,
ρ ˙ = i [ H I , ρ ] + L ρ ,
L ρ = k = L , R j = 1 , 2 L j k ρ L j k 1 2 ( L j k L j k ρ + ρ L j k L j k ) ,
b ± L = 1 2 ( a 1 L ± a 2 L )
b ± R = 1 2 ( a 1 R ± a 2 R )
H I = g 2 ( | r 11 0 | b + L + | r 11 1 | b + R + | r 22 0 | b + L + | r 22 1 | b + R ) e i ( J δ ) t + g 2 ( | r 11 0 | b L + | r 11 1 | b R | r 11 0 | b L | r 11 1 | b R ) e i ( J + δ ) t + Ω | r j j e | e i Δ 1 t + ω | r 11 e | e i Δ 2 t + H.c. .
H I = H s + H d ,
H s = j = 1 , 2 g 2 ( | r j j 0 | b + L + | r j j 1 | b + R ) e i Δ 1 t + Ω | r j j e | e i Δ 1 t + H.c. , H d = j = 1 , 2 ( 1 ) j + 1 g 2 ( | r j j 0 | b L + | r j j 1 | b R ) e i Δ 2 t + ω | r 11 e | e i Δ 2 t + H.c. .
H s = H Ω + H g ,
H Ω = Ω ( | 0 e 0 r | + | r 0 e 0 | + | 1 e 1 r | + | r 1 e 1 | ) | 0 c c 0 | + H.c. , H g = g 2 ( | 0 r 00 | + | r 0 00 | + | 1 r 10 | + | r 1 01 | ) | 0 + L , + L 1 | + g 2 ( | 0 r 01 | + | r 0 10 | + | 1 r 11 | + | r 1 11 | ) | 0 + R , + R 1 | + H.c. + Δ 1 ( | 0 r 0 r | + | r 0 r 0 | + | 1 r 1 r | + | r 1 r 1 | ) | 0 c c 0 | + 2 Δ 1 ( | 00 00 | + | 01 01 | + | 10 10 | ) | 1 + L , + L 1 | + 2 Δ 1 ( | 11 11 | + | 01 01 | + | 10 10 | ) | 1 + R , + R 1 | ,
H s   =     Ω M [ ( | e 0 + | 0 e ) | 0 c T 1 | + ( | e 1 + | 1 e ) | 0 c S 1 | ] + Ω K + [ ( | e 0 | 0 e ) | 0 c T 2 | + ( | e 1 | 1 e ) | 0 c S 2 | ] + Ω K [ ( | e 0 | 0 e ) | 0 c T 4 | + ( | e 1 | 1 e ) | 0 c S 4 | ] + Ω M + [ ( | e 0 + | 0 e ) | 0 c T 5 | + ( | e 1 + | 1 e ) | 0 c S 5 | ] + H.c. + i = 1 5 E i ( | T i T i | + | S i S i | ) ,
| T 1 = 1 2 1 Δ 1 M ( | r 0 + | 0 r ) | 0 c + g 2 M ( M Δ 1 ) ( 2 | 00 | 1 + L + | 01 | 1 + R + | 10 | 1 + R ) , | T 2 = g 2 K ( K + Δ 1 ) ( | r 0 | 0 r ) | 0 c + 1 2 1 + Δ 1 K ( | 10 | 1 + R | 01 | 1 + R ) , | T 3 = 1 3 ( | 10 | 1 + R + | 01 | 1 + R | 00 | 1 + L ) , | T 4 = g 2 K ( K Δ 1 ) ( | r 0 | 0 r ) | 0 c + 1 2 1 Δ 1 K ( | 01 | 1 + R | 10 | 1 + R ) , | T 5 = 1 2 1 + Δ 1 M ( | r 0 + | 0 r ) | 0 c g 2 M ( M + Δ 1 ) ( 2 | 00 | 1 + L + | 01 | 1 + R + | 10 | 1 + R ) .
H s eff   =     Ω ~ ( | e 0 0 e | + | e 1 1 e | ) | 0 c c 0 | + H.c. Δ ~ ( | 0 e 0 e | + | e 0 e 0 | + | 1 e 1 e | + | e 1 e 1 | ) | 0 c c 0 | .
H d   =     ( ω | e 11 r | | 0 c c 0 | + g 2 | 1 11 r | | 1 R , R 0 | + g 2 | 0 11 r | | 1 L , L 0 | ) ( | 0 22 0 | + | 1 22 1 | ) e i Δ 2 t + H.c. .
H d eff   =     g ω 2 Δ 2 ( | e 11 1 | | 0 R , R 1 | + | e 11 0 | | 0 L , L 1 | ) ( | 0 22 0 | + | 1 22 1 | ) g 2 2 Δ 2 | 1 11 0 | | 1 R , L 1 | ( | 0 22 0 | + | 1 22 1 | ) + H.c.,
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.