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

Long-distance transmission of arbitrary quantum states between spatially separated microwave cavities

Open Access Open Access

Abstract

Long-distance transmission between spatially separated microwave cavities is a crucial area of quantum information science and technology. In this work, we present a method for achieving long-distance transmission of arbitrary quantum states between two microwave cavities, by using a hybrid system that comprises two microwave cavities, two nitrogen-vacancy center ensembles (NV ensembles), two optical cavities, and an optical fiber. Each NV ensemble serves as a quantum transducer, dispersively coupling with a microwave cavity and an optical cavity, which enables the conversion of quantum states between a microwave cavity and an optical cavity. The optical fiber acts as a connector between the two optical cavities. Numerical simulations demonstrate that our method allows for the transfer of an arbitrary photonic qubit state between two spatially separated microwave cavities with high fidelity. Furthermore, the method exhibits robustness against environmental decay, parameter fluctuations, and additive white Gaussian noise. Our approach offers a promising way for achieving long-distance transmission of quantum states between two spatially separated microwave cavities, which may have practical applications in networked large-scale quantum information processing and quantum communication.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Different quantum systems may be combined in a hybrid device for exploring new physical phenomenon or developing new quantum technology [19]. As an important component of hybrid systems, microwave cavities or resonators are often used in state-of-the-art science and technology [10]. Over the past years, microwave cavities or photons have been widely used for non-networked small-scale quantum computing in local area. To achieve a networked large-scale quantum computation, it is necessary to implement quantum state transmission between spatially separated microwave cavities.

Microwave photons can be effectively generated and manipulated. While, optical photons can be coherently transmitted by optical fibers. The combination of both enables long-distance quantum communication and quantum information processing [1115]. As optical photons do not directly interact with microwave photons, a transducer is often required to exchange quantum states between microwave photons and optical photons [1618]. Thus, quantum transducers are particularly attractive and play an important role in quantum information science and technology [15,1922].

In recent years, several theoretical proposals for a transducer have been put forward by using various physical systems, such as optomechanical systems [23], NV ensembles [24], atoms [15,25], electro-optical systems [26], magnons [27,28], superconducting circuits [29], and so on. With the help of a transducer, the coherent conversion between microwave and optical photons has been widely studied and has also been demonstrated experimentally [2935].

NV ensembles are one of the most promising candidates for a microwave-to-optical transducer because they can strongly couple to both microwave and optical cavities [24]. The strong coupling of NV ensembles with superconducting and optical cavities has been realized [3639]. On the other hand, NV ensembles have been considered as good quantum memory elements in quantum information processing. Long lifetime of NV ensembles has been experimentally reported [40]. Moreover, various quantum operations such as information transfer, entanglement preparation, and quantum gates have been realized with NV ensembles [4149].

As shown below, the goal of this work is to propose a method for implementing long-distance transmission of arbitrary quantum states between two spatially separated microwave cavities. In this proposal, we use a NV ensemble as a quantum transducer which enables the local transfer of quantum states between a microwave cavity and an optical cavity. In addition, we use an optical fiber to connect two optical cavities, which enables the long-distance transfer of quantum states between the two optical cavities. This method has the following features and advantages.

(1) Remote transmission can be realized by using this method.

(2) The local microwave-optical conversion of photons can be implemented using a single-step operation, and the whole scheme only needs three steps to complete the long-distance transmission of arbitrary quantum states between two spatially separated microwave cavities.

(3) Because the NV ensembles are unexcited during the entire process, decoherence from the NV ensembles is greatly suppressed.

(4) Arbitrary quantum states of microwave photons can be high-fidelity transferred between the spatially separated microwave cavities.

(5) The two optical cavities and the receiver microwave cavity have no requirements for the initial states, and they do not need to be prepared in specific initial states.

(6) The method is robust against the decay caused by environment, the parameter fluctuation, and the additive white Gaussian noise (AWGN).

The structure of this paper is as follows. In Sec. 2, we introduce our hybrid system and derive the effective Hamiltonian. In Sec.3, we explicitly show how to transfer an arbitrary quantum state between two spatially separated microwave cavities. In Sec. 4, we discuss the possible experimental implementation of our proposal and numerically calculate the operational fidelity for transferring an arbitrary quantum state of a microwave photonic qubit in a realistic case. In Sec. 5, we discuss the interferences caused by the real environment. We separately investigate the robustness against stochastic parameter fluctuation and the influences caused by the additive white Gaussian noise (AWGN), which is the most common noise in real experiments. A brief conclusion is given in Sec. 6.

2. Hybrid system and effective Hamiltonian

We consider a hybrid system in Fig. 1 which contains of two parts 1 and 2. Each part contains a microwave cavity, an NV ensemble, and an optical cavity. Each microwave cavity here is a superconducting coplanar waveguide (CPW) resonator. In each part, an NV ensemble is coupled to a microwave cavity via a magnetic coupling and simultaneously interacts with an optical cavity through an optical transition. The two optical cavities (or the two parts 1 and 2) are linked by an optical fiber.

 figure: Fig. 1.

Fig. 1. Schematic of a hybrid system. For each part, an optical cavity with an embedded NV ensemble is placed above a superconducting coplanar waveguide (CPW) cavity. The two optical cavities are linked by an optical fiber.

Download Full Size | PDF

As shown in Fig. 2(a), the NV defect center in diamond consists of a substitutional nitrogen atom (N) associated with a vacancy (V) in an adjacent lattice site of the diamond crystalline matrix. As shown in Fig. 2(b), the energy-level structure of an NV center is formed by a ground state $^{3}A$, an excited state $^{3}E$, and a metastable state $^{1}A$. Both $^{3}A$ and $^{3}E$ are spin triplet states while the metastable $^{1}A$ is a spin singlet state [50]. Owing to the $C_{3v}$ symmetry of the NV center, the ground-state spin sublevels $m_{s}=\pm 1$ are degenerate and the zero-field splitting from $m_{s}=0$ is $D_{gs}=2.87\,\mathrm {GHz}$ [51]. For simplicity, we label the ground state $|^{3}A\rangle \bigotimes |m_{s}=0\rangle$, the first excited state $| ^{3}A\rangle \bigotimes |m_{s}=+1\rangle$, and the second excited state $|^{3}E\rangle \bigotimes |m_{s}=0\rangle$ of a spin in each ensemble as $|0\rangle$, $|1\rangle$, and $|2\rangle$, respectively.

 figure: Fig. 2.

Fig. 2. (a) Atomic structure of an NV center. A single substitutional nitrogen atom (N) is accompanied by a vacancy (V) at a nearest neighbor lattice position. (b) Simplified energy-level diagram of an NV center. The energy-level structure consists of a ground state $^{3}A$, an excited state $^{3}E$, and a metastable state $^{1}A$. Both $^{3}A$ and $^{3}E$ are spin triplet states while the metastable $^{1}A$ is a spin singlet state. Here $D_{gs}$ and $D_{es}$ correspond, respectively, to the zero-field splitting between $m_{s}=0$ and $m_{s}=\pm 1$ in the triplet ground state $^{3}A$ and in the triplet excited state $^{3}E$. (c) The interactions in each part of the hybrid system. For each part, a microwave cavity couples the $|0\rangle _{j}\leftrightarrow |1\rangle _{j}$ transition, an optical cavity couples the $|0\rangle _{j}\leftrightarrow |2\rangle _{j}$, while a classical field drives the $|1\rangle _{j}\leftrightarrow |2\rangle _{j}$ transition of spin $j$ in each NV ensemble, with a Rabi frequency $\Omega$.

Download Full Size | PDF

As usual, in the low excitation limit, the collective spin excitations can be mapped to a boson mode labelled as c and then could be used as a medium for the conversion. In Fig. 1, each part can be described by a Hamiltonian composed of two Jaynes-Cummings (JC) interactions. One is the interaction between the microwave cavity mode $a_{1(2)}$ and the collective spin mode $c_{1(2)}$, while the other is the interaction between the optical cavity mode $b_{1(2)}$ and the collective spin mode $c_{1(2)}$. Here and below, the subscript 1 represents part 1 while the subscript 2 represents part 2 in the hybrid system (Fig. 1).

For each part, we consider the NV ensemble contains $N$ spins. We define the three states $|0\rangle$, $|1\rangle$, and $|2\rangle$ of spin $j$ in the NV ensemble as $|0\rangle _{j}$, $|1\rangle _{j}$, and $|2\rangle _{j}$, respectively (Fig. 2(c)). Here, $j=1,2,\ldots N$. In each part, a microwave cavity couples the $|0\rangle _{j}\leftrightarrow |1\rangle _{j}$ transition, an optical cavity couples the $|0\rangle _{j}\leftrightarrow |2\rangle _{j}$, while a classical field drives the $|1\rangle _{j}\leftrightarrow |2\rangle _{j}$ transition of spin $j$ in each NV ensemble, with a Rabi frequency $\Omega$. The detunings for these transitions are $\Delta _{j,1}=\omega _{j,10}-\omega _{m},\,\Delta _{j,2}=\omega _{j,20}-\omega _{o},$ and $\Delta _{j,3}=\omega _{j,12}-\omega _{p}$, Here $\omega _{j,10}$, $\omega _{j,20}$, and $\omega _{j,21}$ are the $|0\rangle _{j}\leftrightarrow |1\rangle _{j}$, $|0\rangle _{j}\leftrightarrow |2\rangle _{j}$, and $|1\rangle _{j}\leftrightarrow |2\rangle _{j}$ transition frequencies of the $j$th spin for each ensemble, respectively. While $\omega _{m},\,\omega _{o}$ and $\omega _{p}$ are the frequencies of the microwave cavity, the optical cavity, and the laser pulse, respectively.

Due to the fluctuating magnetic environment, there might be fluctuations in the detuning of the level $|1\rangle _{j}$. However, these environmental induced fluctuations are much smaller than the frequency detunings, and can be safely ignored [52]. The $j$th spin located at the position $r_{j}$ is coupled to the two cavities (a microwave cavity and an optical cavity) with the coupling strengths $G_{m,j}\propto B_{1}(r_{j})$ and $G_{o,j}\propto E_{2}(r_{j})$, where $B_{1}(r_{j})$, and $E_{2}(r_{j})$ are the zero-point magnetic and electric fields of the microwave and the optical cavity modes, respectively. As the collective enhanced couplings are employed, here we introduce

$$\begin{aligned} G_{m} & = & \sqrt{\sum_{j=1}^{N}|G_{m,j}(r_{j})|^{2}/N},\\ G_{o} & = & \sqrt{\sum_{j=1}^{N}|G_{o,j}(r_{j})|^{2}/N} \end{aligned}$$
to denote the average coupling strengths for each spin in an NV ensemble [53,54]. Moreover, we assume that the NV center spins in each ensemble are sufficiently far apart so that the direct spin-spin interactions can be ignored [55]. Typically, the NV-NV coupling strength is within a range of a few kilohertz to tens of kilohertz when the distance between NV centers is within 20 nm [56]. Recently, it has been reported that the effect of the direct NV-NV interactions on the dynamics of a microwave-cavity-NV ensemble system is negligible [57].

Since the two parts 1 and 2 in the hybrid system are the same, we first derive the effective Hamiltonian of part 1 for the sake of simplicity. In the interaction picture, after making the rotating-wave approximation, the Hamiltonian of the two cavities and the NV ensemble in part 1 is (assuming $\hbar =1$)

$$\begin{aligned} H_{I,1}= & \sum_{j=1}^{N}G_{m,1}(a^{\dagger}_{1}\sigma_{j,10}^{-}e^{{-}i\Delta_{j,1}t}+a_{1}\sigma_{j,10}^+e^{i\Delta_{j,1}t}) & \\ & +\sum_{j=1}^{N}G_{o,1}(b^{\dagger}_{1}\sigma_{j,20}^{-}e^{{-}i\Delta_{j,2}t}+b_{1}\sigma_{j,20}^+e^{i\Delta_{j,2}t}) & \\ & +\Omega_{1}(\sigma_{j,21}^{-}e^{{-}i\Delta_{j,3}t}+\sigma_{j,21}^+e^{i\Delta_{j,3}t}), & \\ \end{aligned}$$
where the subscript $1$ of $H_{I,1}$, $G_{m,1}$, $G_{o,1}$, $\Omega _{1}$, $a_{1}$, and $b_{1}$ represent part 1 in the hybrid system, $a_{1}$ and $a^{\dagger }_{1}$ ($b_{1}$ and $b^{\dagger }_{1}$) are the annihilation and creation operators for the microwave (optical) cavity in part 1, respectively. Here
$$\begin{aligned} \sigma_{j,10}^+{=}|1\rangle_{j}\langle 0|,\,\sigma_{j,10}^{-}=|0\rangle_{j}\langle 1|,\\ \sigma_{j,20}^+{=}|2\rangle_{j}\langle 0|,\,\sigma_{j,20}^{-}=|0\rangle_{j}\langle 2|,\\ \sigma_{j,21}^+{=}|2\rangle_{j}\langle 1|,\,\sigma_{j,21}^{-}=|1\rangle_{j}\langle 2|,\\ \end{aligned}$$
are the lowering and raising operators of the $j$th spin in the ensemble. And $\Omega _{1}$ is the Rabi frequency of the laser pulse. Note that the presence of random local strain may lead to inhomogeneous broadening in the transition frequencies, and then results in random shifts $\delta _{1,j}=\Delta _{1,j}-\Delta _{1},\, \delta _{2,j}=\Delta _{2,j}^{j}-\Delta _{2}$, and $\delta _{3,j}^{j}=\Delta _{3,j}-\Delta _{3}$ for the $j$th spin of the ensemble, where $\Delta _{1},\,\Delta _{2}$ and $\Delta _{3}$ are the average detunings. After considering the system under the large detuning condition: $|\Delta _{2}|\gg \{|G_{o,1}|,\,|\delta _{2,j}|\}$ and $|\Delta _{3}|\gg \{|\Omega _{1}|,\,|\delta _{3,j}|\}$, one can ignore the inhomogeneous broadening of the transition frequencies. Thus, the Hamiltonian (2) can be expressed as:
$$\begin{aligned} H_{I,1}= & \sum_{j=1}^{N}G_{m,1}(a^{\dagger}_{1}\sigma_{j,10}^{-}e^{{-}i\Delta_{1}t}+a_{1}\sigma_{j,10}^+e^{i\Delta_{1}t}) & \\ & +\sum_{j=1}^{N}G_{o,1}(b^{\dagger}_{1}\sigma_{j,20}^{-}e^{{-}i\Delta_{2}t}+b_{1}\sigma_{j,20}^+e^{i\Delta_{2}t}) & \\ & +\Omega_{1}(\sigma_{j,21}^{-}e^{{-}i\Delta_{3}t}+\sigma_{j,21}^+e^{i\Delta_{3}t}). & \\ \end{aligned}$$

Under the large detuning, in the low excitation limit, and for a large $N$, we could obtain the following effective Hamiltonian describing part 1 [58]:

$$\begin{aligned} H_{\mathrm{eff,1}}= & G_{\mathrm{eff,1}}a_{1}^{\dagger}b_{1}+\mathrm{H.c.}. & \\ \end{aligned}$$
where $G_{\mathrm {eff,1}}=\frac {G_{1,1}G_{2,1}}{2}(\frac {1}{\Delta _{1}}+\frac {1}{\Delta _{2}-\Delta _{3}})$, with $G_{1,1}=\frac {\sqrt {N}G_{o,1}\Omega _{1}}{2}(\frac {1}{\Delta _{2}}+\frac {1}{\Delta _{3}})$ and $G_{2,1}=\sqrt {N}G_{m,1}$. This Hamiltonian is the well-known JC Hamiltonian, which describes the coherent conversion between a microwave-cavity mode and an optical-cavity mode. For the details of deriving this effective Hamiltonian, please refer to Appendix A.

Similar to Part 1, the effective Hamiltonian of Part 2 can be expressed as

$$\begin{aligned} H_{\mathrm{eff,2}}= & G_{\mathrm{eff,2}}a_{2}^{\dagger}b_{2}+\mathrm{H.c.}, & \\ \end{aligned}$$
where $G_{\mathrm {eff,2}}=\frac {G_{1,2}G_{2,2}}{2}(\frac {1}{\Delta _{1}}+\frac {1}{\Delta _{2}-\Delta _{3}})$ with $G_{1,2}=\frac {\sqrt {N}G_{o,2}\Omega _{1}}{2}(\frac {1}{\Delta _{2}}+\frac {1}{\Delta _{3}})$ and $G_{2,2}=\sqrt {N}G_{m,2}$. Therefore, the effective Hamiltonian of the whole system can be written as
$$\begin{aligned} & H_{\mathrm{eff}}=H_{\mathrm{eff,1}}+H_{\mathrm{eff,2}}+H_{fiber}, & \\ & H_{\mathrm{eff,1}}=G_{\mathrm{eff,1}}a_{1}^{\dagger}b_{1}+\mathrm{H.c.}, & \\ & H_{\mathrm{eff,2}}=G_{\mathrm{eff,2}}a_{2}^{\dagger}b_{2}+\mathrm{H.c.}, & \\ & H_{\mathrm{fiber}}=J_{\mathrm{eff}}b_{1}^{\dagger}b_{2}+\mathrm{H.c.}. & \\ \end{aligned}$$

Here $J_{\mathrm {eff}}$ is the coupling strength between the two optical cavities, induced by the fiber. Taking into account a finite length $L$ of the fiber implies a quantization of the modes of the fiber with frequency spacing given by $2\pi c/L$ [59]. Thus, the number of modes, which would significantly interact with the cavities’ modes, is on the order of $n=(L\Gamma _{b})/(2\pi c)$, where $\Gamma _{b}$ is the decay rate of the cavities’ fields into a continuum of fiber modes [59]. Here we will focus on the case: short fiber limit [60], for which essentially only one (resonant) mode of the fiber will interact with the model of each optical cavity ($n\leq 1$). This condition applies in most realistic experimental situations: for instance, the fiber length meets $L\leq 1\mathrm {m}$ and the decay rate of the optical cavity satisfies $\Gamma _{b}\simeq 1\,\mathrm {GHz}$ [60]. It is also noted that under specific conditions [59,61], the fiber mode can be adiabatically eliminated to create an effective cavity-cavity coupling through exchanging cavity photons. Thus, the coupling strength $J_{\mathrm {eff}}$ between the two optical cavities, induced by the modes of a fiber, can be estimated as $J_{\mathrm {eff}}\simeq \sqrt {8\pi \Gamma _{b}c/L}$, which can be increased by decreasing the reflectivity of the cavity mirrors connected to the fiber [5962].

In the next section, we will show in details how to use the effective Hamiltonian $H_{\mathrm {eff}}$ to transfer quantum states between two spatially separated microwave cavities.

3. Quantum state transfer between spatially separated microwave cavities

Now we show how to transfer quantum states between two spatially separated microwave cavities. Generally, the state transfer includes three steps: step 1, turn on the coupling $G_{\mathrm {eff,1}}$ (while $G_{\mathrm {eff,2}}= 0$) in a time interval $0<t<T_{1}$, which uses the spin mode $c_{1}$ to swap the quantum states between the microwave mode $a_{1}$ and the optical mode $b_{1}$ in part 1; step 2, turn off the coupling $G_{\mathrm {eff,1}}$ and $G_{\mathrm {eff,2}}$ in a time interval $T_{1}<t<T_{2}$, so that quantum states are transferred from the optical mode $b_{1}$ in part 1 to the distant optical mode $b_{2}$ in part 2; step 3, turn on the coupling $G_{\mathrm {eff,2}}$ (while $G_{\mathrm {eff,1}}= 0$) in a time interval $T_{2}<t<T_{3}$, which uses the spin mode $c_{2}$ to swap the quantum states between the optical mode $b_{2}$ and the microwave mode $a_{2}$ in part 2. Notice that in the above steps, the coupling between the fiber and the optical cavities can be manipulated through the application of an optical switch [62,63]. Due to potential mismatch between the modes of the fiber and the optical cavities, a coupling loss is typically introduced. By adjusting the coupling loss appropriately, a switching control can be achieved. Here the coupling loss can be adjusted by incorporating a lens to match the mode of the optical cavities, similar to the Q-adjusting cavity scheme which control the leakage of the cavity by adjusting the coupling between the cavity and the environment [59,61]. The effective Hamiltonian of the system reads

$$H_{\mathrm{eff}}= \begin{cases} G_{\mathrm{eff,1}}a_{1}^{\dagger}b_{1}+\mathrm{H.c.}\,\,0<t<T_{1},\\ J_{\mathrm{eff}}b_{1}^{\dagger}b_{2}+\mathrm{H.c.}\,\,T_{1}<t<T_{2},\\ G_{\mathrm{eff,2}}a_{2}^{\dagger}b_{2}+\mathrm{H.c.}\,\,T_{2}<t<T_{3}.\\ \end{cases}$$
Without loss of generality, we suppose that the two microwave cavities ($a_{1},a_{2}$) and the two optical cavities ($b_{1},b_{2}$) are initially in arbitrary pure states $|\psi \rangle _{a_{1}}=\sum _{m=0}^{\infty }c_{m,a_{1}}|m\rangle _{a_{1}}$, $|\phi \rangle _{b_{1}}=\sum _{n=0}^{\infty }c_{n,b_{1}}|n\rangle _{b_{1}}$, $|\eta \rangle _{b_{2}}=\sum _{k=0}^{\infty }c_{l,b_{2}}|k\rangle _{b_{2}}$, and $|\xi \rangle _{a_{2}}=\sum _{l=0}^{\infty }c_{k,a_{2}}|l\rangle _{a_{2}}$, respectively. Here we have $|m\rangle _{a_{1}}=\frac {(a_{1}^{\dagger })^{m}}{\sqrt {m!}}|0\rangle _{a_{1}}$, $|n\rangle _{b_{1}}=\frac {(b_{1}^{\dagger })^{n}}{\sqrt {n!}}|0\rangle _{b_{1}}$, $|k\rangle _{b_{2}}=\frac {(b_{2}^{\dagger })^{k}}{\sqrt {k!}}|0\rangle _{b_{2}}$, and $|l\rangle _{a_{2}}=\frac {(a_{2}^{\dagger })^{l}}{\sqrt {l!}}|0\rangle _{a_{2}}$.

3.1 Quantum state transfer from a microwave cavity to an optical cavity

In part 1, the collective spin mode $c_{1}$ of the NV ensemble interacts with the microwave mode $a_{1}$ and the optical mode $b_{1}$, respectively. In the following, we show that the spin mode can be used to swap the quantum states between the microwave mode $a_{1}$ and the optical mode $b_{1}$.

In the time interval $0<t<T_{1}$, according to Eq. (8), the dynamics of the operators can be derived by solving the Heisenberg equations as

$$\begin{aligned} a^{\dagger}_{1}(t)=\mathrm{cos}(G_{\mathrm{eff,1}}t)a^{\dagger}_{1}+i\mathrm{sin}(G_{\mathrm{eff,1}}t)b^{\dagger}_{1},\\ b^{\dagger}_{1}(t)=\mathrm{cos}(G_{\mathrm{eff,1}}t)b^{\dagger}_{1}+i\mathrm{sin}(G_{\mathrm{eff,1}}t)a^{\dagger}_{1}.\\ \end{aligned}$$
When we set $t=\pi /2G_{\mathrm {eff,1}}$, the solution turns to
$$\begin{aligned} a^{\dagger}_{1}(t)=ib^{\dagger}_{1},\\ b^{\dagger}_{1}(t)=ia^{\dagger}_{1}.\\ \end{aligned}$$

From Eq. (8), one can see that for the time interval $0<t<T_{1}$, the Hamiltonian given in Eq. (8) is the Hamiltonian $H_{\mathrm {eff,1}}$ given above. After the evolution time $t=\pi /2|G_{\mathrm {eff,1}}|$, the state $|\psi \rangle _{a_{1}}|\phi \rangle _{b_{1}}|\eta \rangle _{b_{2}}|\xi \rangle _{a_{2}}$ of the system evolves into

$$\begin{aligned} & e^{{-}iH_{\mathrm{eff,1}}t}|\psi\rangle_{a_{1}}|\phi\rangle_{b_{1}}|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & e^{{-}iH_{\mathrm{eff,1}}t}(|\psi\rangle_{a_{1}}|\phi\rangle_{b_{1}})|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & \sum_{m=0}^{\infty}\sum_{n=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}\frac{c_{n,b_{1}}}{\sqrt{n!}}e^{{-}iH_{\mathrm{eff,1}}t}(a_{1}^{\dagger})^{m}(b_{1}^{\dagger})^{n}|0\rangle_{a_{1}}|0\rangle_{b_{1}}\bigotimes|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & \sum_{m=0}^{\infty}\sum_{n=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}\frac{c_{n,b_{1}}}{\sqrt{n!}}[e^{iH_{\mathrm{eff,1}}t}(a_{1}^{\dagger})^{m}e^{{-}iH_{\mathrm{eff,1}}t}][e^{iH_{\mathrm{eff,1}}t}(b_{1}^{\dagger})^{n}e^{{-}iH_{\mathrm{eff,1}}t}]|0\rangle_{a_{1}}|0\rangle_{b_{1}}\bigotimes|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & \sum_{n=0}^{\infty}\frac{c_{n,b_{1}}}{\sqrt{n!}}(ia_{1}^{\dagger})^{n}|0\rangle_{a_{1}} \sum_{m=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}(ib_{1}^{\dagger})^{m}|0\rangle_{b_{1}}\bigotimes|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & \sum_{n=0}^{\infty}e^{i\frac{\pi}{2}n}c_{n,b_{1}}|n\rangle_{a_{1}} \sum_{m=0}^{\infty}e^{i\frac{\pi}{2}m}c_{m,a_{1}}|m\rangle_{b_{1}}\bigotimes|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\psi\rangle_{b_{1}}|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}}, & \end{aligned}$$
where we have used $a_{1}|0\rangle _{a_{1}}=0$, $b_{1}|0\rangle _{b_{1}}=0$, $(i)^{m}=e^{i\frac {\pi }{2}m}$, and $(i)^{n}=e^{i\frac {\pi }{2}n}$, and we have dropped the phase shifts $e^{i\frac {\pi }{2}m}$ and $e^{i\frac {\pi }{2}n}$, which can be corrected by the local rotations [16]. Thus, through the first step, we successfully transferred the quantum state $|\psi \rangle$ from the microwave cavity to the optical cavity in part 1. Here, the $|\psi \rangle _{a_{1}}$, $|\phi \rangle _{b_{1}}$, $|\eta \rangle _{b_{2}}$, and $|\xi \rangle _{a_{2}}$ states can be arbitrary discrete-variable states (e.g., Fock states) or arbitrary continuous-variable states (e.g., cat states).

3.2 Quantum state transfer between spatially separated optical cavities via the optical fiber

In the time interval $T_{1}<t<T_{2}$, we turn off both couplings $G_{\mathrm {eff,1}}$ and $G_{\mathrm {eff,2}}$, the quantum states could be exchanged between the optical modes $b_{1}$ and $b_{2}$ with assistance of the optical fiber. This state exchange can be described as follows:

$$\begin{aligned} & e^{{-}iH_{\mathrm{fiber}}t}|\phi\rangle_{a_{1}}|\psi\rangle_{b_{1}}|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}e^{{-}iH_{\mathrm{fiber}}t}(|\psi\rangle_{b_{1}}|\eta\rangle_{b_{2}})|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}|\psi\rangle_{b_{2}}|\xi\rangle_{a_{2}}. & \\ \end{aligned}$$

For the detailed derivation, please see Appendix B. From Eq. (12), one can see that an arbitrary quantum state $|\psi \rangle$ of the optical cavity $b_{1}$ is transferred onto the optical cavity $b_{2}$.

3.3 Quantum states transfer from an optical cavity to a microwave cavity

In the time interval $T_{2}<t<T_{3}$, we turn on the coupling $G_{\mathrm {eff,2}}$ while turn off the coupling $G_{\mathrm {eff,1}}$. The quantum states could be exchanged between the optical mode $b_{2}$ and the microwave mode $a_{2}$ as follows:

$$\begin{aligned} & e^{{-}iH_{\mathrm{eff,2}}t}|\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}|\psi\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}e^{{-}iH_{\mathrm{eff,2}}t}(|\psi\rangle_{b_{2}}|\xi\rangle_{a_{2}}) & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}|\xi\rangle_{b_{2}}|\psi\rangle_{a_{2}}. & \\ \end{aligned}$$

In Appendix B, we provide a detailed derivation of Eq. (13). From Eq. (13), one can see that, through the third step, we successfully transfer the quantum state $|\psi \rangle$ from the optical cavity $b_{2}$ to the microwave cavity $a_{2}$. Hence, this three-step procedure completes a perfect state transmission between two spatially separated microwave cavities, which is useful in networked long-distance quantum communication and quantum information processing.

4. Analysis and numerical simulation

To better understand the inner principle of the scheme, we numerically simulate the transfer process. For simplicity, we set $G_{m,1}=G_{m,2}=G_{m}$, $G_{o,1}=G_{o,2}=G_{o}$, and $\Omega _{1}=\Omega _{2}=\Omega$. Since any quantum system would suffer from decoherence, we consider the dissipations from the microwave cavities, the NV ensemble collective spin modes, and the optical cavities. In the low-temperature environment, namely, the thermal photon occupation number is nearly zero, the dynamics of the system can be described by the following master equation

$$\begin{aligned} \dot{\rho} & ={-}i[H_{\mathrm{eff}},\rho]+\mathcal{P}, & \\ \mathcal{P} & =\kappa_{a_{1}}\mathcal{L}(a_{1})+\kappa_{b_{1}}\mathcal{L}(b_{1})+\gamma_{c_{1}}\mathcal{L}(c_{1}) & \\ & +\kappa_{a_{2}}\mathcal{L}(a_{2})+\kappa_{b_{2}}\mathcal{L}(b_{2})+\gamma_{c_{2}}\mathcal{L}(c_{2}), & \\ \mathcal{L}(o) & =2o\rho o^{\dagger}-o^{\dagger}o\rho-\rho o^{\dagger}o, & \end{aligned}$$
where $\rho$ is the density matrix of the whole system. When an optical fiber is coupled to optical cavities, the leakage of the optical cavity mode $b_{1(2)}$ is divided into two parts: leaking into the optical fiber with decay rate $\Gamma _{b_{1(2)}}$ and leaking into the surrounding environments with decay rate $\kappa _{b_{1(2)}}$. Similarly, the microwave cavity mode $a_{1(2)}$ leaks into the surrounding environments with decay rate $\kappa _{a_{1(2)}}$ and the NV ensembles mode $c_{1(2)}$ leaks into the surrounding environments with decay rate $\gamma _{c_{1(2)}}$. Here, $H_{\mathrm {eff}}$ is given by Eq. (8). As an example, we consider the two optical cavities ($b_{1},b_{2}$) and the microwave cavity ($a_{2}$) are initially in the vacuum state. Since microwave-photon coding has important applications in network communication, we take into account the transmission of an arbitrary quantum states of a microwave photonic qubit from the microwave cavity $a_{1}$ to the distant microwave cavity $a_{2}$ . In this case, the initial state of the microwave cavity $a_{1}$ is an arbitrary quantum state of a microwave photonic qubit, i.e., $|\psi \rangle _{a_{1}}=\mathrm {cos}\frac {\theta }{2}|0\rangle _{a_{1}}+e^{i\phi }\mathrm {sin}\frac {\theta }{2}|1\rangle _{a_{1}}$. This initial state of the microwave cavity can be simply generated by several methods [6466], i.e., one can position a superconducting qubit in a microwave cavity and this device not only functions as a deterministic single-photon source but also enables the generation of various Fock states and arbitrary superpositions of Fock states within the cavity field. The controllable interaction between the cavity field and the qubit can be realized by adjusting the external parameters of superconducting qubit [66]. Thus, the initial state of the whole hybrid system is given by:
$$|\psi_{0}\rangle=|\psi\rangle_{a_{1}}|0\rangle_{b_{1}}|0\rangle_{b_{2}}|0\rangle_{a_{2}}.$$

Accordingly, the ideal target state is given by

$$|S\rangle= |0\rangle_{a_{1}}|0\rangle_{b_{1}}|0\rangle_{b_{2}}|\psi\rangle_{a_{2}}.$$

The state transmission efficiency can be evaluated by the average fidelity of the operation, which is defined as

$$\mathcal{F}=\frac{1}{4\pi^{2}}\int_{0}^{2\pi}\int_{0}^{2\pi}\sqrt{\langle S|\rho|S\rangle}\mathrm{d}\theta\mathrm{d}\phi,$$
where $\rho$ is the final practical density operator of the system when the operation is performed in a realistic situation.

We perform numerical simulations by solving the master Eq. (14). The parameters we chosen in the numerical simulations are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$ and $\sqrt {N}G_{o}=2\pi \times 350\,\mathrm {MHz}$ with the number of NV centers $N\sim 10^{12}$ in each NV ensemble [38,67]. We assume the inhomogeneous broadening of the transition frequencies with $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, and $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, which can be achieved experimentally [43,47,48,68]. In addition, we choose $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$ [69]. At the temperature of $T\sim 20\,\mathrm {mK}$, the equilibrium thermal photon occupation number is $n<0.01$, which can be neglected [16].

We first consider the transmission in the ideal case that the decoherence process is not taken into account. As the black-solid line shown in Fig. 3, the spin mode can be used to successfully transfer the quantum state from the microwave cavity $a_{1}$ to the distant microwave cavity $a_{2}$ when selecting a suitable evolution time. In a realistic situation, the decoherence should be taken into account. Thus, we consider the dissipation process with the decay rates $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$ for the microwave cavities, $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$ for the collective spin modes, and $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$ for the optical cavities. As the red-dash line shown in Fig. 3, a high fidelity (0.988) for the state transmission can still be achieved. Therefore, our method works well for the long-distance transmission of an arbitrary quantum state of a microwave photonic qubit between two spatially separated microwave cavities in a realistic case.

 figure: Fig. 3.

Fig. 3. Average fidelity of the operation. The parameters used in the numerical simulation are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$, $\sqrt {N}G_{o}=2\pi \times 350 \,\mathrm {MHz}$, $N\sim 10^{12}$, $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$.

Download Full Size | PDF

5. Interference of the real environment

5.1 Robustness to the stochastic parameter fluctuation

In order to find out whether the scheme is effective in practice, one should take the robustness against stochastic parameter fluctuation into consideration. In this section, we set the amplitude noise as $\beta H_{k}$ and the time derivative of the Brownian motion $W_{t}$ as $\alpha (t)$. Therefore, a stochastic Schrödinger equation in a close system can be written as

$$\begin{aligned} \dot{\phi}_{\alpha}(t) & ={-}i (H_{\mathrm{eff}}+\beta H_{k}\alpha(t))\phi_{\alpha}(t), & \\ \alpha(t) & =\frac{\partial W_{t}}{\partial t}, & \\ H_{k} & =\sum_{i=1,2}G_{\mathrm{eff,i}}a_{i}^{\dagger}b_{i}+\mathrm{H.c.}, \end{aligned}$$
where $\alpha (t)$ satisfies the definition in Ref. [70], i.e., $\langle \alpha (t)\rangle =0$ and $\langle \alpha (t)\alpha (t')\rangle =\delta (t-t')$. This definition ensures that the noise is meaningless and orthorhombic in different times. Then, the dynamics of the system in this case is modeled as (without dissipation terms)
$$\begin{aligned} \dot{\rho}_{\alpha}={-}i[H_\mathrm{eff},\rho_{\alpha}]-i[H_{k},\alpha\rho_{\alpha}], \end{aligned}$$
where $\rho _{\alpha }(t)=|\phi _{\alpha }(t)\rangle \langle \phi _{\alpha }(t)|$. By taking an average to the noise, Eq. (19) becomes
$$\begin{aligned} \dot{\rho}\approx{-}i[H_\mathrm{eff},\rho]-i[H_{k},\langle\alpha\rho_{\alpha}\rangle], \end{aligned}$$
where $\rho =\langle \rho _{\alpha }\rangle$ is referenced from Ref. [71]. Refer to the definition of white noise from Novikov’s theorem [72], we have
$$\begin{aligned} \langle\alpha\rho_{\alpha}\rangle=\dfrac{1}{2} \langle\dfrac{\delta\rho_{\alpha}}{\delta\alpha(t')}\rangle|_{t'=t}={-}i\dfrac{\beta}{2}[H_{k},\rho]. \end{aligned}$$

Thus, the dynamics of the whole system (including the dissipation and noise terms) is

$$\begin{aligned} \dot{\rho}={-}i[H_\mathrm{eff},\rho]+\mathcal{L}\rho+\mathcal{N}\rho, \end{aligned}$$
where $\mathcal {N}\rho =-i\beta [H_{k},\langle \alpha \rho _{\alpha }\rangle ]=-\dfrac {\beta ^{2}}{2}[H_{k},[H_{k},\rho ]]$. The Lindblad operator $\mathcal {L}\rho$ is the same as that defined in Eq. (14).

In the following, to investigate the robustness against stochastic parameter fluctuation, we consider the amplitude noises caused by the JC interaction. Hence, the $\mathcal {N}\rho$ in this case can be rewritten as

$$\begin{aligned} \mathcal{N}\rho & =\dfrac{\beta^{2}}{2}[H_{k},[H_{k},\rho]]. & \\ \end{aligned}$$

According to Eq. (23), we numerically simulate the robustness of the scheme versus the stochastic parameter fluctuation. For the error Hamiltonian $H_{k}$, we select the deviation as $\beta \in [-1{\% },1{\% }]$, which is quite normal in real experiments [73,74]. As shown in Fig. 4, the influence caused by the stochastic parameter fluctuation is negligibly small and the fidelity only reduces $2{\% }$ at most. In this sense, the scheme is robust against stochastic noise errors caused by the JC interaction.

 figure: Fig. 4.

Fig. 4. Average fidelity of the operation versus time with different parameter fluctuations. The parameters used in the numerical simulation are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$, $\sqrt {N}G_{o}=2\pi \times 350 \,\mathrm {MHz}$, $N\sim 10^{12}$, $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$.

Download Full Size | PDF

5.2 Robustness to the additive white Gaussian noise

In real experiments, there may exist several kinds of noises, which disturb wave shapes of Rabi frequencies of pulses. It is worthwhile to study the influences caused by the noise. Here, we investigate the influences caused by the additive white Gaussian noise (AWGN), which is the most common noise in real experiments [75]. The Rabi frequencies under the influence of the AWGN of pulse amplitudes can be described as [76]

$$\begin{aligned} \Omega^{'}_{1}(t)=\Omega_{1}(t)+\mathrm{awgn}(\Omega_{1}(t),R_{SN1}),\\ \Omega^{'}_{2}(t)=\Omega_{2}(t)+\mathrm{awgn}(\Omega_{2}(t),R_{SN2}), \end{aligned}$$
where $\mathrm {awgn}(\Omega _{1}(t),R_{SN1})$ and $\mathrm {awgn}(\Omega _{2}(t),R_{SN2})$ are the functions to generate the AWGN. Note that the signal-to-noise ratio (SNR) indicates the ratio between the original Rabi frequencies and the scales of noise. Here we set the SNR as $R_{SN1}$ and $R_{SN2}$ for pulses $\Omega _{1}(t)$ and $\Omega _{2}(t)$, respectively.

We consider two different conditions based on Eq. (24). For each group of SNR, 50-time numerical simulations are performed, and the results for corresponding $R_{SN1}$ and $=R_{SN2}$ are plotted in Fig. 5. From Fig. 5, one can see that the scheme has great robustness against the AWGN. Even with a quite large $R_{SN1}$ and $R_{SN2}$, e.g., $R_{SN1}=R_{SN2}=5$, the fidelity of the target state can still reach $0.99$. Thus, the method is robust against the additive white Gaussian noise.

 figure: Fig. 5.

Fig. 5. Average fidelity of the operation versus simulation counts with SNR. (a) $R_{SN1}=R_{SN2}=5$, (b) $R_{SN1}=R_{SN2}=10$, (c) $R_{SN1}=R_{SN2}=20$, (d) $R_{SN1}=R_{SN2}=50$. The parameters used in the numerical simulation are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$, $\sqrt {N}G_{o}=2\pi \times 350 \,\mathrm {MHz}$, $N\sim 10^{12}$, $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$.

Download Full Size | PDF

6. Conclusion

We have proposed a method for long-distance transmission of arbitrary quantum states between two spatially separated microwave cavities. The method has these features and advantages. First, the local coherent microwave-optical conversion of photons can be implemented using a single-step operation. Second, only three steps are needed to complete the transmission of an arbitrary quantum states between two spatially separated microwave cavities. Third, the NV ensembles are unexcited during the entire process, thus the decoherence from the NV ensembles is greatly suppressed. Fourth, the two optical cavities and the receiver microwave cavity have no requirements for the initial states, and they do not need to be prepared in specific initial states. Last, the method is robust against the decay caused by environment, the parameter fluctuation, and the additive white Gaussian noise. Our numerical simulations demonstrate that arbitrary quantum states of a microwave photonic qubit can be high-fidelity transfered between two spatially separated microwave cavities. This method could be a good candidate to realize long-distance transmission of arbitrary quantum states between two spatially separated microwave cavities, which has important applications in long-distance networked quantum communication and quantum information processing.

Appendix A. Derivation of the effective Hamiltonian

Under the condition of large detuning, $|\Delta _{2}-\Delta _{1}|\gg \frac {G_{o,1}G_{m,1}}{2}(\frac {1}{\Delta _{1}}+\frac {1}{\Delta _{2}})$ and $|\Delta _{3}-\Delta _{1}|\gg \frac {\Omega _{1}G_{m,1}}{2}(\frac {1}{\Delta _{1}}+\frac {1}{\Delta _{3}})$, we keep the first term of the Hamiltonian (4) unchanged while using the second-order perturbation theory [77] to rewrite the second and third terms as

$$\begin{aligned} H_{e}=\sum_{m,n=1,2}\frac{1}{\omega_{mn}}[h^{\dagger}_{m},h_{n}]\mathrm{exp}(i[\omega_{m}-\omega_{n}]t), \end{aligned}$$
where $\omega _{mn}$ is the harmonic average of $\omega _{m}$ and $\omega _{n}$, with $\frac {1}{\omega _{mn}}=\frac {1}{2}(\frac {1}{\omega _{m}}+\frac {1}{\omega _{n}})$. In Eq. (25), one has:
$$\begin{aligned} & h_{1}=\sum_{j=1}^{N}G_{o,1}b_{1}^{\dagger}\sigma_{j,20}^{-},\,h_{1}^{\dagger}=\sum_{j=1}^{N}G_{o,1}b_{1}\sigma_{j,20}^+, & \\ & h_{2}=\sum_{j=1}^{N}\Omega_{1}\sigma_{j,21}^{-},\,h_{2}^{\dagger}=\sum_{j=1}^{N}\Omega_{1}\sigma_{j,21}^+, & \\ & \omega_{1}=\Delta_{2}-\Delta_{1},\,\omega_{2}=\Delta_{3}-\Delta_{1}, & \end{aligned}$$

There exist the following commutation relations:

$$\begin{aligned} & [h_{1}^{\dagger},h_{1}]=\sum_{j=1}^{N}b_{1}b_{1}^{\dagger}(|2\rangle_{j}\langle2|-|0\rangle_{j}\langle0|), & \\ & [h_{1}^{\dagger},h_{2}]=\sum_{j=1}^{N}-b_{1}|1\rangle_{j}\langle0|, & \\ & [h_{2}^{\dagger},h_{1}]=\sum_{j=1}^{N}-b_{1}^{\dagger}|0\rangle_{j}\langle1|, & \\ & [h_{2}^{\dagger},h_{2}]=\sum_{j=1}^{N}|2\rangle_{j}\langle2|-|1\rangle_{j}\langle1|. & \\ \end{aligned}$$

By substituting Eq. (27) into Eq. (25), the Hamiltonian $H_{e}$ can be rewritten as

$$\begin{aligned} H_{e}= & \sum_{j=1}^{N}[\frac{G_{o,1}^{2}}{\Delta_{2}}b_{1}b_{1}^{\dagger}|2\rangle_{j}\langle 2|-b_{1}^{\dagger}b_{1}|0\rangle_{j}\langle 0|] & \\ & +\sum_{j=1}^{N}\frac{\Omega_{1}^{2}}{\Delta_{3}}(|2\rangle_{j}\langle 2|-|1\rangle_{j}\langle 1|) & \\ & +\sum_{j=1}^{N}-G_{o,1}^{'}[b_{1}^{\dagger}\sigma_{j,10}^{-}e^{{-}i(\Delta_{2}-\Delta_{3})t}+b_{1}\sigma_{j,10}^+e^{i(\Delta_{2}-\Delta_{3})t}], & \end{aligned}$$
where $G_{o,1}^{'}=\frac {G_{o,1}\Omega _{1}}{2}(\frac {1}{\Delta _{2}}+\frac {1}{\Delta _{3}})$. In this case, the Hamiltonian (4) can be expressed as the following effective Hamiltonian
$$\begin{aligned} H_{\mathrm{eff}}= & \sum_{j=1}^{N}[\frac{G_{o,1}^{2}}{\Delta_{2}}b_{1}b_{1}^{\dagger}|2\rangle_{j}\langle 2|-b_{1}^{\dagger}b_{1}|0\rangle_{j}\langle 0|] & \\ & +\sum_{j=1}^{N}\frac{\Omega_{1}^{2}}{\Delta_{3}}(|2\rangle_{j}\langle 2|-|1\rangle_{j}\langle 1|) & \\ & +\sum_{j=1}^{N}G_{m,1}[a_{1}^{\dagger}\sigma_{j,10}^{-}e^{{-}i\Delta_{1}t}+a_{1}\sigma_{j,10}^+e^{i\Delta_{1}t}] & \\ & -\sum_{j=1}^{N}G_{o,1}^{'}[b_{1}^{\dagger}\sigma_{j,10}^{-}e^{{-}i(\Delta_{2}-\Delta_{3})t}+b_{1}\sigma_{j,10}^+e^{i(\Delta_{2}-\Delta_{3})t}]. & \end{aligned}$$

Here, the subscript 1 of $H_{\mathrm {eff}}$ represent part 1 in the hybrid system (Fig. 1). Note that the first and second lines of Eq. (29) describe Stark shifts, the third line of Eq. (29) represents the effective coupling between the $|0\rangle _{j}$ and $|1\rangle _{j}$ of the $j$th spin and the microwave cavity, and the last line of Eq. (29) represents the effective coupling between the $|0\rangle _{j}$ and $|1\rangle _{j}$ of the $j$th spin and the optical cavity.

Under the low excitation limit and for a large $N$, we could map the collective spin operators into boson operators by introducing the Holstein-Primakoff representation [78]:$\sum _{j=1}^{N}\sigma _{j,10}^+=c_{1}^{\dagger }\sqrt {N-c_{1}^{\dagger }c_{1}}\simeq \sqrt {N}c_{1}^{\dagger }$, $\sum _{j=1}^{N}\sigma _{j,10}^{-}=c_{1}\sqrt {N-c_{1}^{\dagger }c_{1}}\simeq \sqrt {N}c_{1}$, and $\sum _{j=1}^{N}\sigma _{j,z}=2c_{1}^{\dagger }c_{1}-N$. where the operators $c_{1}^{\dagger }$ and $c_{1}$ approximately obey the standard boson commutator $[c_{1},c_{1}^{\dagger }]\simeq 1$ and $\sum _{j=1}^{N}|0\rangle _{j}\langle 0|=\sum _{j=1}^{N}\frac {1}{2}(I-\sigma _{j,z})=N-c_{1}^{\dagger }c_{1}$.

When each spin is in the ground state, the first term in the bracket of the first line and all terms in the second line of Eq. (29) can be dropped off [52]. Therefore, Eq. (29) can be further rewritten as

$$\begin{aligned} \widetilde{H}_{\mathrm{eff,1}}= & -\frac{NG_{o,1}^{'}}{\Delta_{2}}b_{1}^{\dagger}b_{1}+\frac{G_{o,1 }^{2}}{\Delta_{2}}b_{1}^{\dagger}b_{1}c^{\dagger}_{1}c_{1} & \\ & +G_{2,1}[a_{1}^{\dagger}c_{1}e^{{-}i\Delta_{1}t}+a_{1}c_{1}^{\dagger}e^{i\Delta_{1}t}] & \\ & -G_{1,1}[b_{1}^{\dagger}c_{1}e^{{-}i(\Delta_{2}-\Delta_{3})t}+b_{1}c_{1}^{\dagger}e^{i(\Delta_{2}-\Delta_{3})t}]. & \end{aligned}$$
where $G_{1,1}=\sqrt {N}G_{o,1}^{'}=\frac {\sqrt {N}G_{o,1}\Omega _{1}}{2}(\frac {1}{\Delta _{2}}+\frac {1}{\Delta _{3}})$ and $G_{2,1}=\sqrt {N}G_{m,1}$.

By applying the large-detuning condition $|\Delta _{1}|\gg G_{2,1}$ and $|\Delta _{2}-\Delta _{3}|\gg G_{1,1}$, the effective Hamiltonian (30) turns to

$$\begin{aligned} \widetilde{H}_{\mathrm{eff,1}}= & -\frac{G_{2,1}^{2}}{\Delta_{1}}a_{1}^{\dagger}a_{1} -(\frac{NG_{o,1}^{2}}{\Delta_{2}}+\frac{G_{1,1}^{2}}{\Delta_{2}-\Delta_{3}})b_{1}^{\dagger}b_{1} & \\ & +G_{\mathrm{eff,1}}[a_{1}^{\dagger}b_{1}e^{i(\Delta_{2}-\Delta_{3}-\Delta_{1})t}+a_{1}b_{1}^{\dagger}e^{{-}i(\Delta_{2}-\Delta_{3}-\Delta_{1})t}], \end{aligned}$$
where $G_{\mathrm {eff,1}}=\frac {G_{1,1}G_{2,1}}{2}(\frac {1}{\Delta _{1}}+\frac {1}{\Delta _{2}-\Delta _{3}})$ and we have used $[c_{1},c_{1}^{\dagger }]=1$ and $c_{1}^{\dagger }c_{1}|0\rangle _{c_{1}}=0$. Note that the collective spin mode of each NV ensemble is assumed to be in the vacuum state during the above derivation. Thus, the second term of Eq. (30) can be dropped off because of $c_{1}^{\dagger }c_{1}|0\rangle _{c_{1}}=0$. Here we divide the Hamiltonian $\widetilde {H}_{\mathrm {eff,1}}$ into two parts: the free Hamiltonian $H_{0,1}=-\frac {G_{2,1}^{2}}{\Delta _{1}}a_{1}^{\dagger }a_{1} -(\frac {NG_{o,1}^{2}}{\Delta _{2}}+\frac {G_{1,1}^{2}}{\Delta _{2}-\Delta _{3}})b_{1}^{\dagger }b_{1}$ and the interaction Hamiltonian $H_{I,1}=G_{\mathrm {eff,1}}[a_{1}^{\dagger }b_{1}e^{i(\Delta _{2}-\Delta _{1}-\Delta _{3})t}+a_{1}b_{1}^{\dagger }e^{-i(\Delta _{2}-\Delta _{1}-\Delta _{3})t}]$. The Hamiltonian $H_{0}$ describes Stark shifts, where the degree of freedom for the spin ensemble has been omitted because the spin ensemble is in the ground state $|0\rangle$. The Hamiltonian $H_{I}$ describes the effective coupling between the microwave cavity and the optical cavity in part 1 of the hybrid system with the effective coupling strength $G_{\mathrm {eff,1}}$. In a new interaction picture with respect to the Hamiltonian $H_{0,1}$, one obtains:
$$\begin{aligned} H_{\mathrm{eff,1}}= & e^{iH_{0,1}t}H_{I,1}e^{{-}iH_{0,1}t} & \\ = & G_{\mathrm{eff,1}}[e^{i\xi_{1}t}e^{i\xi_{2}t}a_{1}^{\dagger}b_{1}+H.c.], & \\ \end{aligned}$$
where $\xi _{1}=\Delta _{2}-\Delta _{1}-\Delta _{3}$ and $\xi _{2}=-\frac {G_{2,1}^{2}}{\Delta _{1}}+\frac {NG_{o,1}^{2}}{\Delta _{2}}+\frac {G_{1,1}^{2}}{\Delta _{2}-\Delta _{3}}$. Since the coupling strength $G_{2,1},\,G_{1,1}$ are the functions of the detuning $\Delta _{1}, \Delta _{2}$, and $\Delta _{3}$, one can adjust the detuning $\Delta _{1}, \Delta _{2}$, and $\Delta _{3}$ to realize $\xi _{1}+\xi _{2}=0$. In this case, Eq. (32) turns to
$$\begin{aligned} H_{\mathrm{eff,1}}= & G_{\mathrm{eff,1}}a_{1}^{\dagger}b_{1}+\mathrm{H.c.}, & \\ \end{aligned}$$
which is the effective Hamiltonian given in Eq. (5) of the main text.

Appendix B. Mathematical derivation of evolution

In the time interval $T_{1}<t<T_{2}$, the quantum states could be exchanged between the optical modes $b_{1}$ and $b_{2}$ as follows:

$$\begin{aligned} & e^{{-}iH_{\mathrm{fiber}}t}|\phi\rangle_{a_{1}}|\psi\rangle_{b_{1}}|\eta\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}e^{{-}iH_{\mathrm{fiber}}t}(|\psi\rangle_{b_{1}}|\eta\rangle_{b_{2}})|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}\bigotimes\sum_{m=0}^{\infty}\sum_{k=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}\frac{c_{k,b_{2}}}{\sqrt{k!}}e^{{-}iH_{\mathrm{fiber}}t}(b_{1}^{\dagger})^{m}(b_{2}^{\dagger})^{k}|0\rangle_{b_{1}}|0\rangle_{b_{2}}\bigotimes|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}\bigotimes\sum_{m=0}^{\infty}\sum_{k=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}\frac{c_{k,b_{2}}}{\sqrt{k!}}[e^{iH_{\mathrm{fiber}}t}(b_{1}^{\dagger})^{m}e^{{-}iH_{\mathrm{fiber}}t}][e^{iH_{\mathrm{fiber}}t}(b_{1}^{\dagger})^{k}e^{{-}iH_{\mathrm{fiber}}t}]|0\rangle_{b_{1}}|0\rangle_{b_{2}}\bigotimes|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}\bigotimes\sum_{k=0}^{\infty}\frac{c_{k,b_{2}}}{\sqrt{k!}}(ib_{1}^{\dagger})^{k}|0\rangle_{b_{1}} \sum_{m=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}(ib_{2}^{\dagger})^{m}|0\rangle_{b_{2}}\bigotimes|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}\bigotimes\sum_{k=0}^{\infty}e^{i\frac{\pi}{2}k}c_{k,b_{2}}|k\rangle_{b_{1}} \sum_{m=0}^{\infty}e^{i\frac{\pi}{2}m}c_{m,a_{1}}|m\rangle_{b_{2}}\bigotimes|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}|\psi\rangle_{b_{2}}|\xi\rangle_{a_{2}}, & \\ \end{aligned}$$
where we have used $b_{1}|0\rangle _{b_{1}}=0$, $b_{2}|0\rangle _{b_{2}}=0$, $(i)^{m}=e^{i\frac {\pi }{2}m}$, and $(i)^{k}=e^{i\frac {\pi }{2}k}$, and we have dropped the phase shifts $e^{i\frac {\pi }{2}m}$ and $e^{i\frac {\pi }{2}k}$, which can be corrected by the local rotations [16].

Similarly, in the time interval $T_{2}<t<T_{3}$, the quantum states could be exchanged between the optical mode $b_{2}$ and the microwave mode $a_{2}$ as

$$\begin{aligned} & e^{{-}iH_{\mathrm{eff,2}}t}|\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}|\psi\rangle_{b_{2}}|\xi\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}e^{{-}iH_{\mathrm{eff,2}}t}(|\psi\rangle_{b_{2}}|\xi\rangle_{a_{2}}) & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}\bigotimes\sum_{m=0}^{\infty}\sum_{l=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}\frac{c_{l,a_{2}}}{\sqrt{l!}}e^{{-}iH_{\mathrm{eff,2}}t}(b_{2}^{\dagger})^{m}(a_{2}^{\dagger})^{l}|0\rangle_{b_{2}}|0\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}\bigotimes\sum_{m=0}^{\infty}\sum_{k=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}\frac{c_{l,a_{2}}}{\sqrt{l!}}[e^{iH_{\mathrm{eff,2}}t}(b_{2}^{\dagger})^{m}e^{{-}iH_{\mathrm{eff,2}}t}][e^{iH_{\mathrm{eff,2}}t}(a_{2}^{\dagger})^{l}e^{{-}iH_{\mathrm{eff,2}}t}]|0\rangle_{b_{2}}|0\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}\bigotimes\sum_{l=0}^{\infty}\frac{c_{l,a_{2}}}{\sqrt{l!}}(ib_{2}^{\dagger})^{l}|0\rangle_{b_{2}} \sum_{m=0}^{\infty}\frac{c_{m,a_{1}}}{\sqrt{m!}}(ia_{2}^{\dagger})^{m}|0\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}\bigotimes\sum_{l=0}^{\infty}e^{i\frac{\pi}{2}l}c_{l,a_{2}}|l\rangle_{b_{2}} \sum_{m=0}^{\infty}e^{i\frac{\pi}{2}m}c_{m,a_{1}}|m\rangle_{a_{2}} & \\ = & |\phi\rangle_{a_{1}}|\eta\rangle_{b_{1}}|\xi\rangle_{b_{2}}|\psi\rangle_{a_{2}}, & \\ \end{aligned}$$
where we have used $b_{2}|0\rangle _{b_{2}}=0$, $a_{2}|0\rangle _{a_{2}}=0$, $(i)^{m}=e^{i\frac {\pi }{2}m}$, and $(i)^{l}=e^{i\frac {\pi }{2}l}$, and we have dropped the phase shifts $e^{i\frac {\pi }{2}m}$ and $e^{i\frac {\pi }{2}l}$, which can be corrected by the local rotations [16].

Funding

National Natural Science Foundation of China (U21A20436, 11074062, 11374083, 11774076, 12004253, 12205069, 12364048); Innovation Program for Quantum Science and Technology (2021ZD0301705).

Disclosures

The authors declare no conflicts of interest.

Data availability

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

References

1. Z.-L. Xiang, X.-Y. Lü, T.-F. Li, et al., “Hybrid quantum circuit consisting of a superconducting flux qubit coupled to a spin ensemble and a transmission-line resonator,” Phys. Rev. B 87(14), 144516 (2013). [CrossRef]  

2. P.-B. Li, Z.-L. Xiang, P. Rabl, et al., “Hybrid quantum device with nitrogen-vacancy centers in diamond coupled to carbon nanotubes,” Phys. Rev. Lett. 117(1), 015502 (2016). [CrossRef]  

3. W. Song, W. Yang, J. An, et al., “Dissipation-assisted spin squeezing of nitrogen-vacancy centers coupled to a rectangular hollow metallic waveguide,” Opt. Express 25(16), 19226–19235 (2017). [CrossRef]  

4. T. Liu, J. Xu, Q.-P. Su, et al., “Two-photon process of a single two-level atom simultaneously absorbing or emitting two photons distributed in different cavities and its applications,” Phys. Rev. A 106(6), 063721 (2022). [CrossRef]  

5. F.-F. Du and Z.-R. Shi, “Robust hybrid hyper-controlled-not gates assisted by an input-output process of low-q cavities,” Opt. Express 27(13), 17493–17506 (2019). [CrossRef]  

6. O. V. Kibis, M. V. Boev, and V. M. Kovalev, “Optically induced hybrid bose–fermi system in quantum wells with different charge carriers,” Opt. Lett. 46(21), 5316–5319 (2021). [CrossRef]  

7. C.-G. Liao, X. Shang, H. Xie, et al., “Dissipation-driven entanglement between two microwave fields in a four-mode hybrid cavity optomechanical system,” Opt. Express 30(7), 10306–10316 (2022). [CrossRef]  

8. G.-Q. Zhang, W. Feng, W. Xiong, et al., “Generating bell states and n-partite w states of long-distance qubits in superconducting waveguide QED,” Phys. Rev. Appl. 20(4), 044014 (2023). [CrossRef]  

9. W. Kersten, N. de Zordo, O. Diekmann, et al., “Triggered superradiance and spin inversion storage in a hybrid quantum system,” Phys. Rev. Lett. 131(4), 043601 (2023). [CrossRef]  

10. W. Gordy, “Microwave spectroscopy,” Rev. Mod. Phys. 20(4), 668–717 (1948). [CrossRef]  

11. Y.-D. Wang and A. A. Clerk, “Using interference for high fidelity quantum state transfer in optomechanics,” Phys. Rev. Lett. 108(15), 153603 (2012). [CrossRef]  

12. Z.-L. Xiang, M. Zhang, L. Jiang, et al., “Intracity quantum communication via thermal microwave networks,” Phys. Rev. X 7(1), 011035 (2017). [CrossRef]  

13. M. Ruf, M. Weaver, S. van Dam, et al., “Resonant excitation and purcell enhancement of coherent nitrogen-vacancy centers coupled to a fabry-perot microcavity,” Phys. Rev. Appl. 15(2), 024049 (2021). [CrossRef]  

14. J.-F. Wang, F.-F. Yan, Q. Li, et al., “Coherent control of nitrogen-vacancy center spins in silicon carbide at room temperature,” Phys. Rev. Lett. 124(22), 223601 (2020). [CrossRef]  

15. B. Kim, H. Kurokawa, K. Sakai, et al., “Diamond optomechanical cavity with a color center for coherent microwave-to-optical quantum interfaces,” Phys. Rev. Appl. 20(4), 044037 (2023). [CrossRef]  

16. T. Liu, J.-L. Zhao, B.-Q. Guo, et al., “One-step implementation of a coherent conversion between microwave and optical cavities via an ensemble of nitrogen-vacancy centers,” Phys. Rev. A 103(2), 023706 (2021). [CrossRef]  

17. C. Zhao, Z. Yang, D. Wang, et al., “Quantum networks assisted by dark modes in optomagnonic systems,” Phys. Rev. A 108(4), 043703 (2023). [CrossRef]  

18. L. Luo, L. Ye, H. He, et al., “Experimental observation of the topological interface state in a one-dimensional waterborne acoustic waveguide,” Phys. Rev. Appl. 20(4), 044008 (2023). [CrossRef]  

19. K. Stannigel, P. Rabl, A. S. Sørensen, et al., “Optomechanical transducers for long-distance quantum communication,” Phys. Rev. Lett. 105(22), 220501 (2010). [CrossRef]  

20. S. Barzanjeh, M. Abdi, G. J. Milburn, et al., “Reversible optical-to-microwave quantum interface,” Phys. Rev. Lett. 109(13), 130503 (2012). [CrossRef]  

21. N. Lauk, N. Sinclair, S. Barzanjeh, et al., “Perspectives on quantum transduction,” Quantum Sci. Technol. 5(2), 020501 (2020). [CrossRef]  

22. K. Dodge, Y. Liu, A. R. Klots, et al., “Hardware implementation of quantum stabilizers in superconducting circuits,” Phys. Rev. Lett. 131(15), 150602 (2023). [CrossRef]  

23. L. Tian, “Adiabatic state conversion and pulse transmission in optomechanical systems,” Phys. Rev. Lett. 108(15), 153604 (2012). [CrossRef]  

24. I. Lekavicius, D. A. Golter, T. Oo, et al., “Transfer of phase information between microwave and optical fields via an electron spin,” Phys. Rev. Lett. 119(6), 063601 (2017). [CrossRef]  

25. X.-L. Hei, P.-B. Li, X.-F. Pan, et al., “Enhanced tripartite interactions in spin-magnon-mechanical hybrid systems,” Phys. Rev. Lett. 130(7), 073602 (2023). [CrossRef]  

26. M. Tsang, “Cavity quantum electro-optics,” Phys. Rev. A 81(6), 063837 (2010). [CrossRef]  

27. R. Hisatomi, A. Osada, Y. Tabuchi, et al., “Bidirectional conversion between microwave and light via ferromagnetic magnons,” Phys. Rev. B 93(17), 174427 (2016). [CrossRef]  

28. C. Gonzalez-Ballestero, T. van der Sar, and O. Romero-Isart, “Towards a quantum interface between spin waves and paramagnetic spin baths,” Phys. Rev. B 105(7), 075410 (2022). [CrossRef]  

29. C. Zhong, Z. Wang, C. Zou, et al., “Proposal for heralded generation and detection of entangled microwave–optical-photon pairs,” Phys. Rev. Lett. 124(1), 010511 (2020). [CrossRef]  

30. R. W. Andrews, R. W. Peterson, T. P. Purdy, et al., “Bidirectional and efficient conversion between microwave and optical light,” Nat. Phys. 10(4), 321–326 (2014). [CrossRef]  

31. M. Mirhosseini, A. Sipahigil, M. Kalaee, et al., “Superconducting qubit to optical photon transduction,” Nature 588(7839), 599–603 (2020). [CrossRef]  

32. J. Han, T. Vogt, C. Gross, et al., “Coherent microwave-to-optical conversion via six-wave mixing in rydberg atoms,” Phys. Rev. Lett. 120(9), 093201 (2018). [CrossRef]  

33. J. G. Bartholomew, J. Rochman, T. Xie, et al., “On-chip coherent microwave-to-optical transduction mediated by ytterbium in yvo4,” Nat. Commun. 11(1), 3266 (2020). [CrossRef]  

34. A. P. Higginbotham, P. S. Burns, M. D. Urmey, et al., “Harnessing electro-optic correlations in an efficient mechanical converter,” Nat. Phys. 14(10), 1038–1042 (2018). [CrossRef]  

35. X.-X. Li, B. Li, and P.-B. Li, “Simulation of topological phases with color center arrays in phononic crystals,” Phys. Rev. Res. 2(1), 013121 (2020). [CrossRef]  

36. Y. Kubo, F. R. Ong, P. Bertet, et al., “Strong coupling of a spin ensemble to a superconducting resonator,” Phys. Rev. Lett. 105(14), 140502 (2010). [CrossRef]  

37. D. I. Schuster, A. P. Sears, E. Ginossar, et al., “High-cooperativity coupling of electron-spin ensembles to superconducting cavities,” Phys. Rev. Lett. 105(14), 140501 (2010). [CrossRef]  

38. B. Julsgaard, C. Grezes, P. Bertet, et al., “Quantum memory for microwave photons in an inhomogeneously broadened spin ensemble,” Phys. Rev. Lett. 110(25), 250503 (2013). [CrossRef]  

39. S. Schietinger, T. Schröder, and O. Benson, “One-by-one coupling of single defect centers in nanodiamonds to high-q modes of an optical microresonator,” Nano Lett. 8(11), 3911–3915 (2008). [CrossRef]  

40. N. Bar-Gill, L. M. Pham, A. Jarmola, et al., “Solid-state electronic spin coherence time approaching one second,” Nat. Commun. 4(1), 1743 (2013). [CrossRef]  

41. W. Zhang, J. Zhang, J. Wang, et al., “Depth-dependent decoherence caused by surface and external spins for NV centers in diamond,” Phys. Rev. B 96(23), 235443 (2017). [CrossRef]  

42. G. He, B. Ye, R. Gong, et al., “Quasi-floquet prethermalization in a disordered dipolar spin ensemble in diamond,” Phys. Rev. Lett. 131(13), 130401 (2023). [CrossRef]  

43. W. Xiong, J. Chen, B. Fang, et al., “Strong tunable spin-spin interaction in a weakly coupled nitrogen vacancy spin-cavity electromechanical system,” Phys. Rev. B 103(17), 174106 (2021). [CrossRef]  

44. T. Xie, Z. Zhao, M. Guo, et al., “Identity test of single NV centers in diamond at hz-precision level,” Phys. Rev. Lett. 127(5), 053601 (2021). [CrossRef]  

45. L. Hanlon, M. Olney-Fraser, M. W. Doherty, et al., “Enhancement of spin-to-charge conversion of diamond NV centers at ambient conditions using surface electrodes,” Phys. Rev. B 107(13), 134111 (2023). [CrossRef]  

46. Q. Ai, P.-B. Li, W. Qin, et al., “The NV metamaterial: Tunable quantum hyperbolic metamaterial using nitrogen vacancy centers in diamond,” Phys. Rev. B 104(1), 014109 (2021). [CrossRef]  

47. G. Q. Yan, S. Li, T. Yamamoto, et al., “Electric-field-induced coherent control of nitrogen-vacancy centers,” Phys. Rev. Appl. 18(6), 064031 (2022). [CrossRef]  

48. P.-B. Li and F. Nori, “Hybrid quantum system with Nitrogen-Vacancy Centers in diamond coupled to surface-phonon polaritons in piezomagnetic superlattices,” Phys. Rev. Appl. 10(2), 024011 (2018). [CrossRef]  

49. Y. Zhang, Q. Wu, S.-L. Su, et al., “Cavity quantum electrodynamics effects with nitrogen vacancy center spins coupled to room temperature microwave resonators,” Phys. Rev. Lett. 128(25), 253601 (2022). [CrossRef]  

50. P. Neumann, R. Kolesov, V. Jacques, et al., “Excited-state spectroscopy of single NV defects in diamond using optically detected magnetic resonance,” New J. Phys. 11(1), 013017 (2009). [CrossRef]  

51. J. F. Barry, J. M. Schloss, E. Bauch, et al., “Sensitivity optimization for NV-diamond magnetometry,” Rev. Mod. Phys. 92(1), 015004 (2020). [CrossRef]  

52. B. Li, P.-B. Li, Y. Zhou, et al., “Quantum microwave-optical interface with nitrogen-vacancy centers in diamond,” Phys. Rev. A 96(3), 032342 (2017). [CrossRef]  

53. J. H. Wesenberg, A. Ardavan, G. A. D. Briggs, et al., “Quantum computing with an electron spin ensemble,” Phys. Rev. Lett. 103(7), 070502 (2009). [CrossRef]  

54. Y. Kubo, C. Grezes, A. Dewes, et al., “Hybrid quantum circuit with a superconducting qubit coupled to a spin ensemble,” Phys. Rev. Lett. 107(22), 220501 (2011). [CrossRef]  

55. S. D. Bennett, N. Y. Yao, J. Otterbach, et al., “Phonon-induced spin-spin interactions in diamond nanostructures: Application to spin squeezing,” Phys. Rev. Lett. 110(15), 156402 (2013). [CrossRef]  

56. I. Jakobi, S. A. Momenzadeh, F. F. de Oliveira, et al., “Efficient creation of dipolar coupled nitrogen-vacancy spin qubits in diamond,” J. Phys.: Conf. Ser. 752(1), 012001 (2016). [CrossRef]  

57. S. Putz, D. Krimer, R. Amsüss, et al., “Protecting a spin ensemble against decoherence in the strong-coupling regime of cavity QED,” Nat. Phys. 10(10), 720–724 (2014). [CrossRef]  

58. 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]  

59. T. Pellizzari, “Quantum networking with optical fibres,” Phys. Rev. Lett. 79(26), 5242–5245 (1997). [CrossRef]  

60. A. Serafini, S. Mancini, and S. Bose, “Distributed quantum computation via optical fibers,” Phys. Rev. Lett. 96(1), 010503 (2006). [CrossRef]  

61. D.-W. Luo, X.-F. Qian, and T. Yu, “Nonlocal magnon entanglement generation in coupled hybrid cavity systems,” Opt. Lett. 46(5), 1073–1076 (2021). [CrossRef]  

62. B. E. Saleh and M. C. Teich, Fundamentals of photonics (John Wiley & sons, 2019).

63. P. Qing, J. Gong, X. Lin, et al., “A simple approach to fiber-based tunable microcavity with high coupling efficiency,” Appl. Phys. Lett. 114(2), 021106 (2019). [CrossRef]  

64. F. De Martini, G. Di Giuseppe, and M. Marrocco, “Single-mode generation of quantum photon states by excited single molecules in a microcavity trap,” Phys. Rev. Lett. 76(6), 900–903 (1996). [CrossRef]  

65. F. De Martini, O. Jedrkiewicz, and P. Mataloni, “Generation of quantum photon states in an active microcavity trap,” J. Mod. Opt. 44(11-12), 2053–2066 (1997). [CrossRef]  

66. Y. xi Liu, L. F. Wei, and F. Nori, “Generation of nonclassical photon states using a superconducting qubit in a microcavity,” Europhys. Lett. 67(6), 941–947 (2004). [CrossRef]  

67. R. Amsüss, C. Koller, T. Nöbauer, et al., “Cavity QED with magnetically coupled collective spin states,” Phys. Rev. Lett. 107(6), 060502 (2011). [CrossRef]  

68. L. J. Zou, D. Marcos, S. Diehl, et al., “Implementation of the dicke lattice model in hybrid quantum system arrays,” Phys. Rev. Lett. 113(2), 023603 (2014). [CrossRef]  

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

70. Y. H. Chen, Z. C. Shi, J. Song, et al., “Accelerated and noise-resistant generation of high-fidelity steady-state entanglement with rydberg atoms,” Phys. Rev. A 97(3), 032328 (2018). [CrossRef]  

71. A. Ruschhaupt, X. Chen, D. Alonso, et al., “Optimally robust shortcuts to population inversion in two-level quantum systems,” New J. Phys. 14(9), 093040 (2012). [CrossRef]  

72. H. R. Miller, D. C. Ravenel, and W. S. Wilson, “Periodic phenomena in the adams-novikov spectral sequence,” Ann. Math. 106(3), 469–516 (1977). [CrossRef]  

73. A. Jayashankar, M. D. H. Long, H. K. Ng, et al., “Achieving fault tolerance against amplitude-damping noise,” Phys. Rev. Res. 4(2), 023034 (2022). [CrossRef]  

74. T. Tanaka, S. Uno, T. Onodera, et al., “Noisy quantum amplitude estimation without noise estimation,” Phys. Rev. A 105(1), 012411 (2022). [CrossRef]  

75. P. Bergmans, “A simple converse for broadcast channels with additive white gaussian noise (corresp.),” IEEE Trans. Inf. Theory 20(2), 279–280 (1974). [CrossRef]  

76. A. Mukherjee, S. Mandal, D. Ghosh, et al., “Influence of additive white gaussian noise on the oeo output,” IEEE J. Quantum Electron. 57(1), 1–10 (2021). [CrossRef]  

77. D. F. James and J. Jerke, “Effective hamiltonian theory and its applications in quantum information,” Can. J. Phys. 85(6), 625–632 (2007). [CrossRef]  

78. K. Hammerer, A. S. Sørensen, and E. S. Polzik, “Quantum interface between light and atomic ensembles,” Rev. Mod. Phys. 82(2), 1041–1093 (2010). [CrossRef]  

Data availability

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

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Schematic of a hybrid system. For each part, an optical cavity with an embedded NV ensemble is placed above a superconducting coplanar waveguide (CPW) cavity. The two optical cavities are linked by an optical fiber.
Fig. 2.
Fig. 2. (a) Atomic structure of an NV center. A single substitutional nitrogen atom (N) is accompanied by a vacancy (V) at a nearest neighbor lattice position. (b) Simplified energy-level diagram of an NV center. The energy-level structure consists of a ground state $^{3}A$, an excited state $^{3}E$, and a metastable state $^{1}A$. Both $^{3}A$ and $^{3}E$ are spin triplet states while the metastable $^{1}A$ is a spin singlet state. Here $D_{gs}$ and $D_{es}$ correspond, respectively, to the zero-field splitting between $m_{s}=0$ and $m_{s}=\pm 1$ in the triplet ground state $^{3}A$ and in the triplet excited state $^{3}E$. (c) The interactions in each part of the hybrid system. For each part, a microwave cavity couples the $|0\rangle _{j}\leftrightarrow |1\rangle _{j}$ transition, an optical cavity couples the $|0\rangle _{j}\leftrightarrow |2\rangle _{j}$, while a classical field drives the $|1\rangle _{j}\leftrightarrow |2\rangle _{j}$ transition of spin $j$ in each NV ensemble, with a Rabi frequency $\Omega$.
Fig. 3.
Fig. 3. Average fidelity of the operation. The parameters used in the numerical simulation are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$, $\sqrt {N}G_{o}=2\pi \times 350 \,\mathrm {MHz}$, $N\sim 10^{12}$, $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$.
Fig. 4.
Fig. 4. Average fidelity of the operation versus time with different parameter fluctuations. The parameters used in the numerical simulation are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$, $\sqrt {N}G_{o}=2\pi \times 350 \,\mathrm {MHz}$, $N\sim 10^{12}$, $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$.
Fig. 5.
Fig. 5. Average fidelity of the operation versus simulation counts with SNR. (a) $R_{SN1}=R_{SN2}=5$, (b) $R_{SN1}=R_{SN2}=10$, (c) $R_{SN1}=R_{SN2}=20$, (d) $R_{SN1}=R_{SN2}=50$. The parameters used in the numerical simulation are $\sqrt {N}G_{m}=2\pi \times 10\,\mathrm {MHz}$, $\sqrt {N}G_{o}=2\pi \times 350 \,\mathrm {MHz}$, $N\sim 10^{12}$, $\Delta _{1}=2\pi \times 10\,\mathrm {MHz}$, $\Delta _{2}=2\pi \times 1\,\mathrm {GHz}$, $\Delta _{3}=2\pi \times 500\,\mathrm {MHz}$, $\Omega =2\pi \times 100\,\mathrm {MHz}$, $\kappa _{a_{1(2)}}=2\pi \times 3\,\mathrm {KHz}$, $\kappa _{b_{1(2)}}=2\pi \times 100\,\mathrm {KHz}$, and $\gamma _{c_{1(2)}}=2\pi \times 10\,\mathrm {KHz}$.

Equations (35)

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

G m = j = 1 N | G m , j ( r j ) | 2 / N , G o = j = 1 N | G o , j ( r j ) | 2 / N
H I , 1 = j = 1 N G m , 1 ( a 1 σ j , 10 e i Δ j , 1 t + a 1 σ j , 10 + e i Δ j , 1 t ) + j = 1 N G o , 1 ( b 1 σ j , 20 e i Δ j , 2 t + b 1 σ j , 20 + e i Δ j , 2 t ) + Ω 1 ( σ j , 21 e i Δ j , 3 t + σ j , 21 + e i Δ j , 3 t ) ,
σ j , 10 + = | 1 j 0 | , σ j , 10 = | 0 j 1 | , σ j , 20 + = | 2 j 0 | , σ j , 20 = | 0 j 2 | , σ j , 21 + = | 2 j 1 | , σ j , 21 = | 1 j 2 | ,
H I , 1 = j = 1 N G m , 1 ( a 1 σ j , 10 e i Δ 1 t + a 1 σ j , 10 + e i Δ 1 t ) + j = 1 N G o , 1 ( b 1 σ j , 20 e i Δ 2 t + b 1 σ j , 20 + e i Δ 2 t ) + Ω 1 ( σ j , 21 e i Δ 3 t + σ j , 21 + e i Δ 3 t ) .
H e f f , 1 = G e f f , 1 a 1 b 1 + H . c . .
H e f f , 2 = G e f f , 2 a 2 b 2 + H . c . ,
H e f f = H e f f , 1 + H e f f , 2 + H f i b e r , H e f f , 1 = G e f f , 1 a 1 b 1 + H . c . , H e f f , 2 = G e f f , 2 a 2 b 2 + H . c . , H f i b e r = J e f f b 1 b 2 + H . c . .
H e f f = { G e f f , 1 a 1 b 1 + H . c . 0 < t < T 1 , J e f f b 1 b 2 + H . c . T 1 < t < T 2 , G e f f , 2 a 2 b 2 + H . c . T 2 < t < T 3 .
a 1 ( t ) = c o s ( G e f f , 1 t ) a 1 + i s i n ( G e f f , 1 t ) b 1 , b 1 ( t ) = c o s ( G e f f , 1 t ) b 1 + i s i n ( G e f f , 1 t ) a 1 .
a 1 ( t ) = i b 1 , b 1 ( t ) = i a 1 .
e i H e f f , 1 t | ψ a 1 | ϕ b 1 | η b 2 | ξ a 2 = e i H e f f , 1 t ( | ψ a 1 | ϕ b 1 ) | η b 2 | ξ a 2 = m = 0 n = 0 c m , a 1 m ! c n , b 1 n ! e i H e f f , 1 t ( a 1 ) m ( b 1 ) n | 0 a 1 | 0 b 1 | η b 2 | ξ a 2 = m = 0 n = 0 c m , a 1 m ! c n , b 1 n ! [ e i H e f f , 1 t ( a 1 ) m e i H e f f , 1 t ] [ e i H e f f , 1 t ( b 1 ) n e i H e f f , 1 t ] | 0 a 1 | 0 b 1 | η b 2 | ξ a 2 = n = 0 c n , b 1 n ! ( i a 1 ) n | 0 a 1 m = 0 c m , a 1 m ! ( i b 1 ) m | 0 b 1 | η b 2 | ξ a 2 = n = 0 e i π 2 n c n , b 1 | n a 1 m = 0 e i π 2 m c m , a 1 | m b 1 | η b 2 | ξ a 2 = | ϕ a 1 | ψ b 1 | η b 2 | ξ a 2 ,
e i H f i b e r t | ϕ a 1 | ψ b 1 | η b 2 | ξ a 2 = | ϕ a 1 e i H f i b e r t ( | ψ b 1 | η b 2 ) | ξ a 2 = | ϕ a 1 | η b 1 | ψ b 2 | ξ a 2 .
e i H e f f , 2 t | ϕ a 1 | η b 1 | ψ b 2 | ξ a 2 = | ϕ a 1 | η b 1 e i H e f f , 2 t ( | ψ b 2 | ξ a 2 ) = | ϕ a 1 | η b 1 | ξ b 2 | ψ a 2 .
ρ ˙ = i [ H e f f , ρ ] + P , P = κ a 1 L ( a 1 ) + κ b 1 L ( b 1 ) + γ c 1 L ( c 1 ) + κ a 2 L ( a 2 ) + κ b 2 L ( b 2 ) + γ c 2 L ( c 2 ) , L ( o ) = 2 o ρ o o o ρ ρ o o ,
| ψ 0 = | ψ a 1 | 0 b 1 | 0 b 2 | 0 a 2 .
| S = | 0 a 1 | 0 b 1 | 0 b 2 | ψ a 2 .
F = 1 4 π 2 0 2 π 0 2 π S | ρ | S d θ d ϕ ,
ϕ ˙ α ( t ) = i ( H e f f + β H k α ( t ) ) ϕ α ( t ) , α ( t ) = W t t , H k = i = 1 , 2 G e f f , i a i b i + H . c . ,
ρ ˙ α = i [ H e f f , ρ α ] i [ H k , α ρ α ] ,
ρ ˙ i [ H e f f , ρ ] i [ H k , α ρ α ] ,
α ρ α = 1 2 δ ρ α δ α ( t ) | t = t = i β 2 [ H k , ρ ] .
ρ ˙ = i [ H e f f , ρ ] + L ρ + N ρ ,
N ρ = β 2 2 [ H k , [ H k , ρ ] ] .
Ω 1 ( t ) = Ω 1 ( t ) + a w g n ( Ω 1 ( t ) , R S N 1 ) , Ω 2 ( t ) = Ω 2 ( t ) + a w g n ( Ω 2 ( t ) , R S N 2 ) ,
H e = m , n = 1 , 2 1 ω m n [ h m , h n ] e x p ( i [ ω m ω n ] t ) ,
h 1 = j = 1 N G o , 1 b 1 σ j , 20 , h 1 = j = 1 N G o , 1 b 1 σ j , 20 + , h 2 = j = 1 N Ω 1 σ j , 21 , h 2 = j = 1 N Ω 1 σ j , 21 + , ω 1 = Δ 2 Δ 1 , ω 2 = Δ 3 Δ 1 ,
[ h 1 , h 1 ] = j = 1 N b 1 b 1 ( | 2 j 2 | | 0 j 0 | ) , [ h 1 , h 2 ] = j = 1 N b 1 | 1 j 0 | , [ h 2 , h 1 ] = j = 1 N b 1 | 0 j 1 | , [ h 2 , h 2 ] = j = 1 N | 2 j 2 | | 1 j 1 | .
H e = j = 1 N [ G o , 1 2 Δ 2 b 1 b 1 | 2 j 2 | b 1 b 1 | 0 j 0 | ] + j = 1 N Ω 1 2 Δ 3 ( | 2 j 2 | | 1 j 1 | ) + j = 1 N G o , 1 [ b 1 σ j , 10 e i ( Δ 2 Δ 3 ) t + b 1 σ j , 10 + e i ( Δ 2 Δ 3 ) t ] ,
H e f f = j = 1 N [ G o , 1 2 Δ 2 b 1 b 1 | 2 j 2 | b 1 b 1 | 0 j 0 | ] + j = 1 N Ω 1 2 Δ 3 ( | 2 j 2 | | 1 j 1 | ) + j = 1 N G m , 1 [ a 1 σ j , 10 e i Δ 1 t + a 1 σ j , 10 + e i Δ 1 t ] j = 1 N G o , 1 [ b 1 σ j , 10 e i ( Δ 2 Δ 3 ) t + b 1 σ j , 10 + e i ( Δ 2 Δ 3 ) t ] .
H ~ e f f , 1 = N G o , 1 Δ 2 b 1 b 1 + G o , 1 2 Δ 2 b 1 b 1 c 1 c 1 + G 2 , 1 [ a 1 c 1 e i Δ 1 t + a 1 c 1 e i Δ 1 t ] G 1 , 1 [ b 1 c 1 e i ( Δ 2 Δ 3 ) t + b 1 c 1 e i ( Δ 2 Δ 3 ) t ] .
H ~ e f f , 1 = G 2 , 1 2 Δ 1 a 1 a 1 ( N G o , 1 2 Δ 2 + G 1 , 1 2 Δ 2 Δ 3 ) b 1 b 1 + G e f f , 1 [ a 1 b 1 e i ( Δ 2 Δ 3 Δ 1 ) t + a 1 b 1 e i ( Δ 2 Δ 3 Δ 1 ) t ] ,
H e f f , 1 = e i H 0 , 1 t H I , 1 e i H 0 , 1 t = G e f f , 1 [ e i ξ 1 t e i ξ 2 t a 1 b 1 + H . c . ] ,
H e f f , 1 = G e f f , 1 a 1 b 1 + H . c . ,
e i H f i b e r t | ϕ a 1 | ψ b 1 | η b 2 | ξ a 2 = | ϕ a 1 e i H f i b e r t ( | ψ b 1 | η b 2 ) | ξ a 2 = | ϕ a 1 m = 0 k = 0 c m , a 1 m ! c k , b 2 k ! e i H f i b e r t ( b 1 ) m ( b 2 ) k | 0 b 1 | 0 b 2 | ξ a 2 = | ϕ a 1 m = 0 k = 0 c m , a 1 m ! c k , b 2 k ! [ e i H f i b e r t ( b 1 ) m e i H f i b e r t ] [ e i H f i b e r t ( b 1 ) k e i H f i b e r t ] | 0 b 1 | 0 b 2 | ξ a 2 = | ϕ a 1 k = 0 c k , b 2 k ! ( i b 1 ) k | 0 b 1 m = 0 c m , a 1 m ! ( i b 2 ) m | 0 b 2 | ξ a 2 = | ϕ a 1 k = 0 e i π 2 k c k , b 2 | k b 1 m = 0 e i π 2 m c m , a 1 | m b 2 | ξ a 2 = | ϕ a 1 | η b 1 | ψ b 2 | ξ a 2 ,
e i H e f f , 2 t | ϕ a 1 | η b 1 | ψ b 2 | ξ a 2 = | ϕ a 1 | η b 1 e i H e f f , 2 t ( | ψ b 2 | ξ a 2 ) = | ϕ a 1 | η b 1 m = 0 l = 0 c m , a 1 m ! c l , a 2 l ! e i H e f f , 2 t ( b 2 ) m ( a 2 ) l | 0 b 2 | 0 a 2 = | ϕ a 1 | η b 1 m = 0 k = 0 c m , a 1 m ! c l , a 2 l ! [ e i H e f f , 2 t ( b 2 ) m e i H e f f , 2 t ] [ e i H e f f , 2 t ( a 2 ) l e i H e f f , 2 t ] | 0 b 2 | 0 a 2 = | ϕ a 1 | η b 1 l = 0 c l , a 2 l ! ( i b 2 ) l | 0 b 2 m = 0 c m , a 1 m ! ( i a 2 ) m | 0 a 2 = | ϕ a 1 | η b 1 l = 0 e i π 2 l c l , a 2 | l b 2 m = 0 e i π 2 m c m , a 1 | m a 2 = | ϕ a 1 | η b 1 | ξ b 2 | ψ a 2 ,
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.