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

Adiabatic preparation of Multipartite GHZ states via Rydberg ground-state blockade

Open Access Open Access

Abstract

The multipartite GHZ states are useful resources for quantum information processing. Here we put forward a scalable way to adiabatically prepare the multipartite GHZ states in a chain of Rydberg atoms. Building on the ground-state blockade effect of Rydberg atoms and the stimulated Raman adiabatic passage (STIRAP), we suppress the adverse effect of the atomic spontaneous emission, and obtain a high fidelity of the multipartite GHZ states without requirements on the operational time. After investigating the feasibility of the proposal, we show a 3-qubit GHZ state can be generated in a wide range of relevant parameters and a fidelity above $98\%$ is achievable with the current experimental technologies.

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

1. Introduction

Entanglement is an elementary static resource of quantum mechanics. It provides the possibilities to test quantum mechanics against local hidden theory [1,2], and performs various applications in quantum information processing, such as superdense coding [3,4], quantum teleportation [57], quantum fingerprinting [8,9], and direct characterization of quantum dynamics [10]. While the entangled states of bipartite systems have been studied substantially [11,12], the multipartite entangled states are of more great interest for quantum information and quantum computation [1326], which can allow us to research the richer and more complex structures of a many-body quantum system [27]. However, the correlations they exhibit cannot be regarded as a trivial generalization of bipartite entanglement. As the number of particles for an entangled state exceeding two, the evaluation of the entanglement will be more difficult.

For the numerous multipartite entangled states, Greenberger- Horne-Zeilinger (GHZ) states [28] represent a paradigmatic example, which play an important role in fundamental tests of quantum mechanics. They not only manifest a conflict with local realism for non-statistical predictions of quantum mechanics [28], but also serve as a reference in quantum estimation theory yielding the Heisenberg scaling [13]. Therefore, the preparation of multipartite GHZ states is an indispensable part and many schemes have been proposed to create the multipartite GHZ states via a single- or multi-step process [2931]. Among the more powerful and spectacular strategies are adiabatic passage schemes, in which an interference-induced dark state provides a possibility for the efficient transfer of population from the source to target state. It has some advantages because the spontaneous emission from excited states can be suppressed and it is robust against experimental parameter errors. The most famous examples of adiabatic passages are the stimulated Raman adiabatic passage (STIRAP) [3234]. It is also an extensively used tool in quantum information processing tasks besides preparations of entanglement [3537].

Neutral atoms excited to high principal quantum number states are now emerging as a reliable tool for investigating the few- and many-body physics, since they can interact through strong and long-range dipole-dipole or van der Waals interaction [30]. After ingeniously integrating the interaction with other techniques, a lot of intriguing effects have been given rise to, where the two notable instances are Rydberg blockade [38] and antiblockade [39]. The former is capable of suppressing the simultaneous excitation for multiple Rydberg atoms. On the contrary, the latter will facilitate the simultaneous excitation of multiple Rydberg atoms and significantly inhibit the Rydberg blockade effect. Benefiting from these attractive effects, a variety of proposals were designed theoretically and experimentally for quantum computation [4050], entanglement generation [5163], quantum algorithms [64], quantum simulators [65], and quantum repeaters [66].

By virtue of a strong Rydberg antiblockade effect and a weak Raman transition, our group [67] further put forward another effect, the ground-state blockade of Rydberg atoms. Differing from the Rydberg blockade and antiblockade, this mechanism will prevent the double occupation of a certain ground state. It can maintain the nonlinear Rydberg-Rydberg interaction, and avoid the adverse effect of atomic spontaneous emission by restraining the evolution of the system into the subspace spanned by the ground states. We have applied the effect to obtain a Bell state and a $W$ state. Quite recently, Ostmann et al. [58] presented three different protocols with a multistep process in chains of Rydberg atoms, which can be used to prepare an antiferromagnetic GHZ state and a class of matrix product states. But the scheme is not good at resisting against atomic spontaneous emission and a concrete control on time of evolution is demanded. It inspires us to exploit the ground-state blockade to generate multipartite GHZ states in chains of Rydberg atoms.

In this paper, we propose a novel scheme to prepare a $N$-qubit GHZ state ($N=3,4,5\dots$) in a chain of four-level Rydberg atoms via the combination of the ground-state blockade effect and the STIRAP, where the atoms all consist of two ground states $|0\rangle ,|1\rangle$ (encoded quantum qubits), one excited state $|p\rangle$, and one Rydberg state $|r\rangle$. To intuitively explain the operational principle, we take the case of the three-qubit GHZ state $(|010\rangle +|101\rangle )/\sqrt {2}$ as an example. Due to the ground-state blockade effect, the initial state $|000\rangle$ will evolve into $(|010\rangle +|100\rangle )/\sqrt {2}$, which will further transform into the target state $(|010\rangle +|101\rangle )/\sqrt {2}$. First, there needs no accurate control over the operational time as the technique of STIRAP. Furthermore, the present scheme is robust against the atomic spontaneous emission and need not precisely tailored Rabi frequencies, Rydberg-mediated interaction and the detuning of the driving fields. The designs of other relevant parameters are also flexible. These will be demonstrated in the following in detail. Additionally, we thoroughly investigate the experimental feasibility with the state-of-the-art technology and a high fidelity above $98\%$ can be acquired.

2. Principle of the present scheme

To prepare the $N$-qubit GHZ state, our scheme needs a $(N-1)$-step process, which can be divided into two groups: Step 1 and the other steps (Step $n$, $n=2,3,\ldots ,N-1$). The concrete operations of the two groups and the corresponding atomic energy levels of $N$ identical Rydberg atoms have been shown in Fig. 1. For the Step 1, only the first two atoms will be addressed individually by three lasers. The ground states $|0\rangle$ and $|1\rangle$ of the first two atoms are dispersively coupled to the excited states $|p\rangle$ by the lasers with Rabi frequencies $\Omega _a(t)$ and $\Omega _b(t)$, detunings $-\Delta _p$, respectively. In order to avoid the precise termination of the operation, we integrate the STIRAP into our proposal by introducing the laser pulses in possession of Rabi frequencies,

$$\Omega_a(t)=\Omega_a\exp\left[-\frac{(t-t_c/2-\tau)^2}{T^2}\right],$$
$$\Omega_b(t)= \Omega_b\exp\left[-\frac{(t-t_c/2+\tau)^2}{T^2}\right],$$
where $T$, $t_c$ and $\tau$ are the pulse period, the pulse duration and the delay between the pulses, respectively. Moreover, the transitions $|1\rangle \leftrightarrow |r\rangle$ of the two atoms are driven by the lasers with Rabi frequencies $\Omega _c$, detunings $-\Delta _r$. In the interaction picture, the Hamiltonian can be written as
$$\begin{aligned}H^{1}_I(t)&=\sum_{j=1,2}\Omega_a(t)|p\rangle_j\langle0|e^{{-}i\Delta_pt}+\Omega_b(t)|p\rangle_j\langle1|e^{{-}i\Delta_pt}+\Omega_c|r\rangle_j\langle1|e^{{-}i\Delta_rt}+\textrm{H.c.}\\ &+\sum_{\alpha\neq\beta}U_{\alpha\beta}|rr\rangle_{\alpha\beta}\langle rr|, \end{aligned}$$
where $U_{\alpha \beta }$ is the Rydberg-mediated interaction of the $\alpha$-th atom and the $\beta$-th atom. Since the interaction originates from the dipole-dipole potential of the scale $C_3/r^3$ or the long-range van der Waals interaction proportional to $C_6/r^6$ ($r$ is the distance of two atoms and $C_{3,6}$ rely on the quantum numbers of the Rydberg state) [30,68], we can properly adjust the positions of atoms to ignore the Rydberg-mediated interactions except for those of the adjacent Rydberg atoms. Furthermore, while we operate the atoms of the Step 1 or the Step $n$, the corresponding next-nearest neighbor atoms will not be excited to the Rydberg state. Viewed from this perspective, we can ignore the next-nearest neighbor interaction. In addition, the final results for our work are independent of the Rydberg-mediated interaction type ($C_3/r^3$ or $C_6/r^6$) and only the strength of the interaction at a given fixed distance of these atoms is relevant.

 figure: Fig. 1.

Fig. 1. The total scheme preparing the $N$-qubit GHZ state composes of $N-1$ steps, which can be divided into two groups: Step 1 and Step $n~(n=2,3,\ldots ,N-1)$. For the Step 1, only the first two atoms will be addressed individually by three lasers, respectively. As for the Step $n$, we only coupled the $n$-th atom to one laser and the $(n+1)$-th atom to three lasers, respectively.

Download Full Size | PDF

After performing a rotation with respect to $\exp (-it\Delta _p|p\rangle \langle p|-it\Delta _r|r\rangle \langle r|)$, the Hamiltonian can be simplified as

$$\begin{aligned}H^{1}_I(t)&=\sum_{j=1,2}\Omega_a(t)|p\rangle_j\langle0|+\Omega_b(t)|p\rangle_j\langle1|+\Omega_c|r\rangle_j\langle1|+\textrm{H.c.}-\Delta_p|p\rangle_j\langle p|-\Delta_r|r\rangle_j\langle r|\\ &+\sum_{k=1}^{N-1}U_{k,k+1}|rr\rangle_{k,k+1}\langle rr|. \end{aligned}$$
Then we consider the large detuning limits $\Delta _p\gg \{\Omega _a,\Omega _b\}$ and the conditions of Rydberg antiblockade $\Delta _r\gg \Omega _c$ and $U_{12}=2\Delta _r-{2\Omega _c^2}/{\Delta _r}$, the Hamiltonian can be further reformulated into
$$\begin{aligned} H^{1}_I(t)&=\sqrt{2}\Omega_a(t)|\psi^{1}\rangle_{12}\langle00|+\Omega_b(t)|\psi^{1}\rangle_{12}\langle\psi^{0}|+\Omega_a(t)|\psi^{2}\rangle_{12}\langle\psi^{0}|+\sqrt{2}\Omega_b(t)|\psi^{2}\rangle_{12}\langle11|\\ &+\Omega|11\rangle_{12}\langle rr|+\textrm{H.c.}-\Delta_p|\psi^{1}\rangle_{12}\langle\psi^{1}|-\Delta_p|\psi^{2}\rangle_{12}\langle\psi^{2}|, \end{aligned}$$
where $|\psi ^0\rangle _{12}=(|01\rangle _{12}+|10\rangle _{12})/{\sqrt {2}},$ $|\psi ^1\rangle _{12}=(|0p\rangle _{12}+|p0\rangle _{12})/{\sqrt {2}},$ $|\psi ^2\rangle _{12}=(|1p\rangle _{12}+|p1\rangle _{12})/{\sqrt {2}},$ and $\Omega ={2\Omega _c^2}/{\Delta _r}.$ The Stark-shifts terms of the ground states have been omitted as they can be compensated by the auxiliary levels. Comparing the detuning of double excited states with that of single excited states, the former is so large that we also ignore the transitions of double excited states. In addition, the Rydberg-mediated interaction of the atoms are also neglected except for that between atoms 1 and 2, since only the first two atoms interact with lasers.

According to the principle of Rydberg ground-state blockade, we diagonalize the terms, $\Omega |11\rangle _{12}\langle rr|+\textrm {H.c.}$, to rewritten the Eq. (5) as,

$$\begin{aligned} H^{1}_I(t)&=\sqrt{2}\Omega_a(t)|\psi^{1}\rangle_{12}\langle00|+\Omega_b(t)|\psi^{1}\rangle_{12}\langle\psi^{0}|+\Omega_a(t)|\psi^{2}\rangle_{12}\langle\psi^{0}|+\Omega_b(t)|\psi^{2}\rangle_{12}(\langle+|+\langle-|)\\ &+\textrm{H.c.}+\Omega(|+\rangle_{12}\langle+|-|-\rangle_{12}\langle-|)-\Delta_p|\psi^{1}\rangle_{12}\langle\psi^{1}|-\Delta_p|\psi^{2}\rangle_{12}\langle\psi^{2}|, \end{aligned}$$
where $|\pm \rangle _{12}=(|rr\rangle _{12}\pm |11\rangle _{12})/\sqrt {2}$ are the eigenvectors of $\Omega |11\rangle _{12}\langle rr|+\textrm {H.c.}$ with respect to the eigenvalues $\pm \Omega$. In the region of $\Delta _r\gg \Omega _c$, the detuning $\Delta _p$ not only far exceeds the Rabi frequencies $\Omega _a$ and $\Omega _b$, but also largely outweighs the value of $\Omega =2\Omega _c^2/\Delta _r$. Therefore, capitalizing on adiabatic elimination of the $|\psi ^2\rangle _{12}$ in Eq. (6), two equivalent transitions $|+\rangle _{12}\leftrightarrow |\psi ^0\rangle _{12}$ and $|-\rangle _{12}\leftrightarrow |\psi ^0\rangle _{12}$ can be derived with the equivalent Rabi frequency $\Omega _a(t)\Omega _b(t)/\Delta _p$ and detuning $\Omega$. Once we set $\Omega \gg \max [\Omega _a(t)\Omega _b(t)/\Delta _p]$, the equivalent transitions can be also removed. Then a Hamiltonian of an equivalent three-level system can read as,
$$H^{1}_\textrm{eff}(t)=\sqrt{2}\Omega_a(t)|\psi^{1}\rangle_{12}\langle00|+\Omega_b(t)|\psi^{1}\rangle_{12}\langle\psi^{0}|+\textrm{H.c.}-\Delta_p|\psi^{1}\rangle_{12}\langle\psi^{1}|.$$
It reflects that the ground-state blockade effect emerges, i.e., the simultaneous occupation of ground states $|1\rangle$ for these atoms is prevented. Furthermore, the unique dark state of the equivalent system is
$$|\Phi\rangle_{12}=\cos[\Theta(t)]|00\rangle_{12}-\sin[\Theta(t)]|\psi^{0}\rangle_{12},$$
where $\Theta (t)=\arctan [\sqrt {2}\Omega _a(t)/\Omega _b(t)]$. According to the design of Eqs. (1) and (2), it is evident that the boundary conditions of the STIRAP are fulfilled
$$\lim_{t\rightarrow-\infty}\cos[\Theta(t)]=1,\lim_{t\rightarrow+\infty}\cos[\Theta(t)]=0.$$
Hence, the Step 1 will stabilize the system beginning with $|00..0\rangle _{12\cdots N}$ at the state $(|010\cdots 0\rangle _{12\cdots N}+|100\cdots 0\rangle _{12\cdots N})/\sqrt {2}$ as time goes on.

While the Step 1 is finished, the other steps summarized as Step $n$ ($n=2,3,4\cdots N$-1) will be executed successively. As shown in Fig. 1, for the Step $n$, only the $n$-th and $(n+1)$-th atoms interact with external lasers. The $n$-th atom couples with a single laser (Rabi frequency $\Omega _c$, detuning $\Delta _r$) to realize the transition $|1\rangle _n\leftrightarrow |r\rangle _n$, and the manipulations of $(n+1)$-th atom are the same as those of the atoms in Step 1. The associated Hamiltonian is

$$\begin{aligned}H^{n}_I(t)&=\sum_{j=n,n+1}(\Omega_c|r\rangle_j\langle1|+\textrm{H.c.}-\Delta_r|r\rangle_j\langle r|)+\Omega_a(t)|p\rangle_{n+1}\langle0|+\Omega_b(t)|p\rangle_{n+1}\langle1|+\textrm{H.c.}\\ &-\Delta_p|p\rangle_{n+1}\langle p|+U_{n,n+1}|rr\rangle_{n,n+1}\langle rr|, \end{aligned}$$
where we have discarded the Rydberg-mediated interactions of atoms except for the $n$- and $(n+1)$-th atoms. Similar to the Step 1, the $H^{n}_I$ can be reduced into a three-level system with
$$H^{n}_\textrm{eff}(t)=\Omega_a(t)|0p\rangle_{n,n+1}\langle00|+\Omega_b(t)|0p\rangle_{n,n+1}\langle01|+\textrm{H.c.}-\Delta_p|0p\rangle_{n,n+1}\langle0p|.$$
The corresponding dark state is
$$|\Phi\rangle_{n,n+1}=\cos[\Theta'(t)]|00\rangle_{n,n+1}-\sin[\Theta'(t)]|01\rangle_{n,n+1}.$$
Here $\Theta '(t)=\arctan [\Omega _a(t)/\Omega _b(t)]$ and the boundary conditions of STIRAP is also satisfied.

Owing to the previous steps, the initial state of Step $n$ will be the state $|\psi _\textrm {even}\rangle$ ($n$ is an even number) or $|\psi _\textrm {odd}\rangle$ ($n$ is an odd number) with $|\psi _\textrm {even}\rangle =(|0101\cdots 01\rangle _{1234\cdots n-1}\otimes |00\rangle _{n,n+1}\otimes |0\cdots 0\rangle _{n+2\cdots N}+|1010\cdots 10\rangle _{1234\cdots n-1}\otimes |10\rangle _{n,n+1}\otimes |0\cdots 0\rangle _{n+2\cdots N})/\sqrt {2}$ or $|\psi _\textrm {odd}\rangle =(|0101\cdots 010\rangle _{1234\cdots n-3,n-2,n-1}\otimes |10\rangle _{n,n+1}\otimes |0\cdots 0\rangle _{n+2\cdots N}+|1010\cdots 101\rangle _{1234\cdots n-3,n-2,n-1}\otimes |00\rangle _{n,n+1}\otimes |0\cdots 0\rangle _{n+2\cdots N})/\sqrt {2}$. And there are only the $n$-th and $(n+1)$-th atoms in action for Step $n$. Therefore, referring to the Eq. (11), the state $(|01\rangle _{n,n+1}+|10\rangle _{n,n+1})/\sqrt {2}$ can be achieved from $(|00\rangle _{n,n+1}+|10\rangle _{n,n+1})/\sqrt {2}$. It is worth mentioning that there will be a relative phase between the terms of $|01\rangle _{n,n+1}$ and $|10\rangle _{n,n+1}$, which can be eliminated by a single qubit logical gate operated on the subspace $\{|0\rangle ,|1\rangle \}$.

In order to certify the feasibility of our proposal and the validity of the effective system, we show the shapes of pulses to prepare the $N$-qubit GHZ states, and plot the fidelity of the corresponding GHZ states governed by the original Hamiltonian and the effective Hamiltonian in Fig. 2. The aim of Figs. 2(a) and 2(c) is to prepare the 3-qubit GHZ state $|\varphi _2\rangle =(|010\rangle +|101\rangle )/\sqrt {2}$. We first apply the Step 1 to evolve the system from the initial state $|000\rangle$ to the state $|\varphi _1\rangle =(|010\rangle +|100\rangle )/\sqrt {2}$ at the timescale $\Omega _ct\in [0,6000]$. The fidelity of $|\varphi _1\rangle$ arrives at $99.65\%$ in the end of Step 1. In the Step 2 ($\Omega _ct\in (6000,14000]$), the state $|\varphi _1\rangle$ is transferred to the 3-qubit GHZ state $|\varphi _2\rangle$, and the target state owns a high fidelity of up to $99.57\%$. Similarly, in Figs. 2(b) and 2(d), we intend to prepare the 4-qubit GHZ state $|\phi _3\rangle =(|0101\rangle +|1010\rangle )/\sqrt {2}$. For the Step 1, the state $|\phi _1\rangle =(|0100\rangle +|1000\rangle )/\sqrt {2}$ comes from the initial state $|0000\rangle$ as $\Omega _ct\in [0,6000]$. Next, the Step 2 transfers the state $|\phi _1\rangle$ to the state $|\phi _2\rangle =(|0100\rangle +|1010\rangle )/\sqrt {2}$ at $\Omega _ct\in (6000,14000]$. Ultimately, the 4-qubit GHZ state $|\phi _3\rangle$ is successfully generated with a high fidelity $99.49\%$ in the Step 3 ($\Omega _ct\in (14000,22000]$). In addition, the empty circles of Figs. 2(c) and 2(d) are the fidelity of $|\varphi _2\rangle$ and $|\phi _3\rangle$ governed by the effective Hamiltonian. Their behaviors are identical to those of corresponding solid lines. Thus the feasibility of the original system and the validity of the reduced system are adequately proven. The tendencies of the original system can be accurately predicted by the reduced system.

 figure: Fig. 2.

Fig. 2. (a) The shapes of pulses to prepare the 3-qubit GHZ state. For Step 1 ($\Omega _c t\in [0,6000]$), we set $t_c=6000/\Omega _c$. For Step 2 ($\Omega _c t\in (6000,14000]$), we set $t_c=8000/\Omega _c$. (b) The shapes of pulses to prepare the 4-qubit GHZ state. For Step 1 ($\Omega _c t\in [0,6000]$), we set $t_c=6000/\Omega _c$. For Steps 2 and 3 ($\Omega _c t\in (6000,14000]$ and $(14000,22000]$), we set $t_c=8000/\Omega _c$. (c) and (d) are the fidelity of the $3$- and $4$-qubit states governed by the original Hamiltonian and the effective Hamiltonian, where the fidelity of state $\rho _i=|i\rangle \langle i|$ is defined as $F=\textrm {Tr}\sqrt {\rho _i^{1/2}\rho (t)\rho _i^{1/2}}$ and $\rho (t)$ is the density matrix of system at time $t$. The other relevant parameters are all chosen as: $\Omega _a=\Omega _b=\sqrt {0.05}\Omega _c$, $\Delta _p=20\Omega _c$, $\Delta _r=20\Omega _c$, $T=0.15t_c$, and $\tau =0.1t_c$.

Download Full Size | PDF

3. Analysis on the influences of relevant parameters

As is well known, the heart of the Rydberg antiblockade is a rigorous relation between the strength of Rydberg-mediated interaction and the detuning of the driving fields [39]. It obviously increases the experimental complexity. However, the Rydberg antiblockade effect of our scheme is easier to achieve. Because there is no rigorous requirement on the strength of Rydberg-mediated interaction and the detuning of the driving fields. In the previous derivation, we assume $U_{n,n+1}=U_{12}=2\Delta _r-\Omega$ for simplicity, which is not necessary. In actual fact, there can be a difference $\delta$, i.e.,$U_{n,n+1}=U_{12}=2\Delta _r-\Omega +\delta$. The Hamiltonian of Eq. (5) can be represented as

$$\begin{aligned} H^{1}_I(t)&=\sqrt{2}\Omega_a(t)|\psi^{1}\rangle_{12}\langle00|+\Omega_b(t)|\psi^{1}\rangle_{12}\langle\psi^{0}|+\Omega_a(t)|\psi^{2}\rangle_{12}\langle\psi^{0}|+\sqrt{2}\Omega_b(t)|\psi^{2}\rangle_{12}\langle11|\\ &+\Omega|11\rangle_{12}\langle rr|+\textrm{H.c.}-\Delta_p|\psi^{1}\rangle_{12}\langle\psi^{1}|-\Delta_p|\psi^{2}\rangle_{12}\langle\psi^{2}|+\delta|rr\rangle_{12}\langle rr|. \end{aligned}$$
After diagonalizing the terms $\Omega |11\rangle _{12}\langle rr|+\textrm {H.c.}+\delta |rr\rangle _{12}\langle rr|$, we can obtain that
$$\begin{aligned}H^{1}_I(t)&=\sqrt{2}\Omega_a(t)|\psi^{1}\rangle_{12}\langle00|+\Omega_b(t)|\psi^{1}\rangle_{12}\langle\psi^{0}|+\Omega_a(t)|\psi^{2}\rangle_{12}\langle\psi^{0}|+\Omega_b(t)|\psi^{2}\rangle_{12}(\sin\theta\langle\tilde+|\\ &-\cos\theta\langle\tilde-|)+\textrm{H.c.}+\tilde\Omega_+|\tilde+\rangle_{12}\langle\tilde+|+\tilde\Omega_-|\tilde-\rangle_{12}\langle\tilde-|-\Delta_p|\psi^{1}\rangle_{12}\langle\psi^{1}|-\Delta_p|\psi^{2}\rangle_{12}\langle\psi^{2}|.\end{aligned}$$
Here we have abbreviated the eigenvectors $|\tilde \pm \rangle =\cos \theta |rr\rangle \pm \sin \theta |11\rangle$ of $\Omega |11\rangle _{12}\langle rr|+\textrm {H.c.}+\delta |rr\rangle _{12}\langle rr|$ homologous with the eigenvalues $\tilde \Omega _{\pm }=(\delta \pm \sqrt {\delta ^2+4\Omega ^2})/2$, and parameterized $\arctan [2\Omega /(\delta +\sqrt {\delta ^2+4\Omega ^2})]$ as $\theta$. Distinctly, once the conditions $|\tilde \Omega _+|\gg \max \{|\Omega _a(t)\Omega _b(t)\sin \theta /\Delta _p|\}$ and $|\tilde \Omega _-|\gg \max \{|\Omega _a(t)\Omega _b(t)\cos \theta /\Delta _p|\}$ are satisfied, the effective Hamiltonian of Eq. (7) will emerge again. And the Eq. (11) can be treated in the similar method.

In Fig. 3, we investigate the dynamical evolution of the fidelity for the 3-qubit GHZ state with different $\delta$. When the $\delta$ is $0.01\Omega _c$, $0.03\Omega _c$, and $0.06\Omega _c$, the ratios of $\tilde \Omega _+$ to $\Omega _a\Omega _b\sin \theta /\Delta _p$ and $\tilde \Omega _-$ to $\Omega _a\Omega _b\cos \theta /\Delta _p$ are respectively equal to $43.14,~50.33,~63.68$ and $37.13,~32.15,~26.23$. Consequently, the fidelities of the target state are all as great as $99.57\%$ at $\Omega _ct=14000$. And the appearances fully affirm the experimental flexibility of our scheme.

 figure: Fig. 3.

Fig. 3. The dynamical evolution of the fidelity for the 3-qubit GHZ state with different $\delta$. For the Step 1 and Step 2, we set $t_c=6000/\Omega _c$ and $t_c=8000/\Omega _c$, respectively. The other relevant parameters are $\Omega _{a}=\Omega _b=\sqrt {2}\Omega _c$, $\Delta _p=800\Omega _c$, $\Delta _r=20\Omega _c$, $T=0.15t_c$, and $\tau =0.1t_c$.

Download Full Size | PDF

In Fig. 4, we exhibit the fidelity of the 3-qubit GHZ state as functions of time with different $\Delta _r$. While the $\Delta _r$ is respectively chosen as $10\Omega _c$, $100\Omega _c$, and $400\Omega _c$, the fidelity of the target state will arrive at $99.05\%$, $99.76\%$, and $98.22\%$ in succession, which means the quality of the target state is not a linear growth with the values of $\Delta _r$. It can be interpreted by the limiting conditions $\Delta _r\gg \Omega _c$ and $\Omega =2\Omega _c^2/\Delta _r\gg \max [\Omega _a(t)\Omega _b(t)/\Delta _p]$, where a little $\Delta _r$ or a large one will destroy the former or the latter. Although there may be an optimal value of $\Delta _r$ for the preparation of GHZ states, it still has a wide range for $\Delta _r$ to ensure the target state with a high fidelity above $99\%$.

 figure: Fig. 4.

Fig. 4. The dynamical evolution of the fidelity for the 3-qubit GHZ state with different $\Delta _r$. For the Step 1 and Step 2, we set $t_c=6000/\Omega _c$ and $t_c=8000/\Omega _c$, respectively. The other relevant parameters are $\Omega _{a}=\Omega _b=\sqrt {2}\Omega _c$, $\Delta _p=800\Omega _c$, $\delta =0$, $T=0.15t_c$, and $\tau =0.1t_c$.

Download Full Size | PDF

Besides the outstanding flexibility, our scheme is also immune to the atomic spontaneous emission, because the adiabatic elimination of the excited states can be perfectly executed by the suitable regulation on $\Delta _p$, $\Omega _a$, and $\Omega _b$.

The original master equation of the $j$-th step ($j=1,2,\ldots ,N-1$) can be written as

$$\dot\rho={-}i[H_I^{j}(t),\rho]+\mathcal{L}_j\rho+\mathcal{L}_{j+1}\rho,$$
where
$$\mathcal{L}_j\rho = \sum_{k=1}^3L_{jk}\rho L_{jk}^\dagger-\frac{1}{2}(L_{jk}^\dagger L_{jk}\rho+\rho L_{jk}^\dagger L_{jk}),$$
$$L_{j1}=\sqrt{\frac{\gamma_p}{2}}|0\rangle_j\langle p|,$$
$$L_{j2}= \sqrt{\frac{\gamma_p}{2}}|1\rangle_j\langle p|,$$
$$L_{j3}= \sqrt{\gamma}|1\rangle_j\langle r|. $$
Here we consider the decay rates of the atomic spontaneous emission from exited states and Rydberg states to the ground states are respectively equal to ${\gamma _p/2}$ and ${\gamma }$. And it is worthy mentioning that the Rydberg states decaying to different states will hardly have effect on the final results, since it is not populated during the whole process.

In the Fig. 5, we illustrate the fidelity of the 3-qubit GHZ state by the Eq. (15) with the different $\Delta _p$ and a fixed ratio $\Omega _a\Omega _b/\Delta _p$. When $\Delta _p$ is selected as $20\Omega _c,$ $200\Omega _c$, $400\Omega _c$, and $1600\Omega _c$ in turn, the corresponding fidelity of the 3-qubit GHZ state is equal to $74.96\%$, $96.10\%$, $97.55\%$ and $98.45\%$ at $\Omega _ct=14000$. These performances soundly indicate the robustness of the scheme against atomic spontaneous emission is strengthened via increasing $\Delta _p$ (for a fixed ratio $\Omega _a\Omega _b/\Delta _p$). It is also in accordance with the character of the ground-state blockade effect [67].

 figure: Fig. 5.

Fig. 5. The fidelity of the $3$-qubit GHZ state governed by the original master equation. The decay rates are $\gamma _p=3\Omega _c$ and $\gamma =0.001\Omega _c$. The other relevant parameters are the same as those of Fig. 2(c).

Download Full Size | PDF

In an experiment, one of the key elements for our scheme is the single site addressability of the Rydberg atoms. Referring to [58,6973], arraying a group of Rydberg atoms into various geometries has been an available experimental technology. Particularly, Nogrette et al. [69] demonstrated single-atom trapping in two-dimensional arrays of microtraps with arbitrary geometries. And Endres et al. [72] took advantage of atom-by-atom assembly to implement a platform for the deterministic preparation of regular one-dimensional arrays of individually controlled cold atoms. In addition, based on the techniques of optical tweezer arrays with Rydberg atoms, Ostmann et al. [58] also devised GHZ and matrix product states engineering and quantum state transport in a one-dimensional geometry. Another core technology of our scheme is the ground-state blockade effect of Rydberg atoms, the experimental feasibility of which has been fleshed out in more detail in [67]. And the Rabi laser frequencies of $\Omega _a$ and $\Omega _b$ can be altered continuously between $2\pi \times (0,100)$ MHz [61,74]. As for the transition $|1\rangle \leftrightarrow |r\rangle$, it can be realized by two indirect transitions with Rabi frequencies $\Omega _1, \Omega _2$ and a common detuning $\Delta$ [53,75,76], and the $\Omega _c$ can be equivalent to $\Omega _1\Omega _2/\Delta$. Consequently, the desired value of $\Omega _c$ can be obtained by experimentally tuning $\Delta$ at will. In [50,67,74,77,78], we can acquire a group of experimental parameters for the atomic decay rates in the ${}^{87}$Rb atom with $(\gamma _e,\gamma )=2\pi \times (3,0.001)$ MHz, where the excited state and the Rydberg state respectively correspond to the $5p_{3/2}$ atomic state and the $97d_{5/2}$ Rydberg state. According to the decay rates and the other experimental parameters $(\Omega _c,\Omega _a,\Omega _b,\Delta _p,\Delta _r)=2\pi \times (1,1.4,1.4,800,20)$ MHz, we numerically calculate the fidelity of the 3-qubit GHZ state, which reaches $98.42\%$, where the $t_c$ for the Step 1 and Step 2 are $6000/\Omega _c$ and $8000/\Omega _c$, $T$ is $0.15t_c$, and $\tau$ is $0.1t_c$. All of the analyses reliably confirm the experimental feasibility of our scheme.

4. Summary

In summary, we elaborately achieve a scheme to adiabatically prepare multipartite GHZ state in a chain of Rydberg atoms. Employing the ground-state blockade of Rydberg atoms and the technique of the STIRAP, the scheme simultaneously possesses the robustness against atomic spontaneous emission and the flexibility for experimental operations. Moreover, there is no rigorous requirement on the relevant parameters and they can be varied in wide ranges during the course of an experiment. We reasonably investigate the influences of relevant parameters, scientifically explore the feasibility of the scheme by the current experimental parameters, and finally obtain a high fidelity of 3-qubit GHZ state above $98\%$. We believe our scheme supplies a new prospect with regard to the generation of entangled states.

Funding

National Natural Science Foundation of China (NSFC) (11774047).

References

1. A. Einstein, B. Podolsky, and N. Rosen, “Can quantum-mechanical description of physical reality be considered complete?” Phys. Rev. 47(10), 777–780 (1935). [CrossRef]  

2. J. S. Bell, Speakable and Unspeakable in Quantum Mechanics (Cambridge University, 1988).

3. C. H. Bennett and S. J. Wiesner, “Communication via one- and two-particle operators on einstein-podolsky-rosen states,” Phys. Rev. Lett. 69(20), 2881–2884 (1992). [CrossRef]  

4. K. Mattle, H. Weinfurter, P. G. Kwiat, and A. Zeilinger, “Dense coding in experimental quantum communication,” Phys. Rev. Lett. 76(25), 4656–4659 (1996). [CrossRef]  

5. C. H. Bennett, G. Brassard, C. Crépeau, R. Jozsa, A. Peres, and W. K. Wootters, “Teleporting an unknown quantum state via dual classical and einstein-podolsky-rosen channels,” Phys. Rev. Lett. 70(13), 1895–1899 (1993). [CrossRef]  

6. D. Bouwmeester, J. W. Pan, K. Mattle, M. Eibl, H. Weinfurter, and A. Zeilinger, “Experimental quantum teleportation,” Nature 390(6660), 575–579 (1997). [CrossRef]  

7. D. Boschi, S. Branca, F. De Martini, L. Hardy, and S. Popescu, “Experimental realization of teleporting an unknown pure quantum state via dual classical and einstein-podolsky-rosen channels,” Phys. Rev. Lett. 80(6), 1121–1125 (1998). [CrossRef]  

8. H. Buhrman, R. Cleve, J. Watrous, and R. de Wolf, “Quantum fingerprinting,” Phys. Rev. Lett. 87(16), 167902 (2001). [CrossRef]  

9. R. T. Horn, S. A. Babichev, K.-P. Marzlin, A. I. Lvovsky, and B. C. Sanders, “Single-qubit optical quantum fingerprinting,” Phys. Rev. Lett. 95(15), 150502 (2005). [CrossRef]  

10. M. Mohseni and D. A. Lidar, “Direct characterization of quantum dynamics,” Phys. Rev. Lett. 97(17), 170501 (2006). [CrossRef]  

11. M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University, 2000).

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

13. V. Giovannetti, S. Lloyd, and L. Maccone, “Quantum-enhanced measurements: Beating the standard quantum limit,” Science 306(5700), 1330–1336 (2004). [CrossRef]  

14. C.-P. Yang and S. Han, “Preparation of greenberger-horne-zeilinger entangled states with multiple superconducting quantum-interference device qubits or atoms in cavity qed,” Phys. Rev. A 70(6), 062323 (2004). [CrossRef]  

15. O. Gühne, G. Tóth, and H. J. Briegel, “Multipartite entanglement in spin chains,” New J. Phys. 7, 229 (2005). [CrossRef]  

16. C.-P. Yang, “Preparation of $n$-qubit greenberger-horne-zeilinger entangled states in cavity qed: An approach with tolerance to nonidentical qubit-cavity coupling constants,” Phys. Rev. A 83(6), 062302 (2011). [CrossRef]  

17. M. Hofmann, A. Osterloh, and O. Gühne, “Scaling of genuine multiparticle entanglement close to a quantum phase transition,” Phys. Rev. B 89(13), 134101 (2014). [CrossRef]  

18. J. Stasińska, B. Rogers, M. Paternostro, G. De Chiara, and A. Sanpera, “Long-range multipartite entanglement close to a first-order quantum phase transition,” Phys. Rev. A 89(3), 032330 (2014). [CrossRef]  

19. C.-P. Yang, Q.-P. Su, S.-B. Zheng, and F. Nori, “Entangling superconducting qubits in a multi-cavity system,” New J. Phys. 18(1), 013025 (2016). [CrossRef]  

20. C. Song, K. Xu, W. Liu, C.-p. Yang, S.-B. Zheng, H. Deng, Q. Xie, K. Huang, Q. Guo, L. Zhang, P. Zhang, D. Xu, D. Zheng, X. Zhu, H. Wang, Y.-A. Chen, C.-Y. Lu, S. Han, and J.-W. Pan, “10-qubit entanglement and parallel logic operations with a superconducting circuit,” Phys. Rev. Lett. 119(18), 180511 (2017). [CrossRef]  

21. L. Pezzè, M. Gabbrielli, L. Lepori, and A. Smerzi, “Multipartite entanglement in topological quantum phases,” Phys. Rev. Lett. 119(25), 250401 (2017). [CrossRef]  

22. Q.-P. Su, H.-H. Zhu, L. Yu, Y. Zhang, S.-J. Xiong, J.-M. Liu, and C.-P. Yang, “Generating double noon states of photons in circuit qed,” Phys. Rev. A 95(2), 022339 (2017). [CrossRef]  

23. D. Sauerwein, N. R. Wallach, G. Gour, and B. Kraus, “Transformations among pure multipartite entangled states via local operations are almost never possible,” Phys. Rev. X 8(3), 031020 (2018). [CrossRef]  

24. C.-P. Yang and Z.-F. Zheng, “Deterministic generation of greenberger-horne-zeilinger entangled states of cat-state qubits in circuit qed,” Opt. Lett. 43(20), 5126–5129 (2018). [CrossRef]  

25. P. Contreras-Tejada, C. Palazuelos, and J. I. de Vicente, “Resource theory of entanglement with a unique multipartite maximally entangled state,” Phys. Rev. Lett. 122(12), 120503 (2019). [CrossRef]  

26. G. K. Naik, R. Singh, and S. K. Mishra, “Controlled generation of genuine multipartite entanglement in floquet ising spin models,” Phys. Rev. A 99(3), 032321 (2019). [CrossRef]  

27. O. Gühne and G. Tóth, “Entanglement detection,” Phys. Rep. 474(1-6), 1–75 (2009). [CrossRef]  

28. M. H. D. M. Greenberger and A. Zeilinger, Bell’s Theorem, Quantum Theory, and Conceptions of the Universe (Kluwer Academic, Dordrecht, 1989), pp. 69–72.

29. M. Müller, I. Lesanovsky, H. Weimer, H. P. Büchler, and P. Zoller, “Mesoscopic rydberg gate based on electromagnetically induced transparency,” Phys. Rev. Lett. 102(17), 170502 (2009). [CrossRef]  

30. M. Saffman, T. G. Walker, and K. Mølmer, “Quantum information with rydberg atoms,” Rev. Mod. Phys. 82(3), 2313–2363 (2010). [CrossRef]  

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

32. J. R. Kuklinski, U. Gaubatz, F. T. Hioe, and K. Bergmann, “Adiabatic population transfer in a three-level system driven by delayed laser pulses,” Phys. Rev. A 40(11), 6741–6744 (1989). [CrossRef]  

33. U. Gaubatz, P. Rudecki, S. Schiemann, and K. Bergmann, “Population transfer between molecular vibrational levels by stimulated raman scattering with partially overlapping laser fields. a new concept and experimental results,” J. Chem. Phys. 92(9), 5363–5376 (1990). [CrossRef]  

34. K. Bergmann, H. Theuer, and B. W. Shore, “Coherent population transfer among quantum states of atoms and molecules,” Rev. Mod. Phys. 70(3), 1003–1025 (1998). [CrossRef]  

35. Z. Kis and F. Renzoni, “Qubit rotation by stimulated raman adiabatic passage,” Phys. Rev. A 65(3), 032318 (2002). [CrossRef]  

36. S.-B. Zheng, “Nongeometric conditional phase shift via adiabatic evolution of dark eigenstates: A new approach to quantum computation,” Phys. Rev. Lett. 95(8), 080502 (2005). [CrossRef]  

37. N. V. Vitanov, A. A. Rangelov, B. W. Shore, and K. Bergmann, “Stimulated raman adiabatic passage in physics, chemistry, and beyond,” Rev. Mod. Phys. 89(1), 015006 (2017). [CrossRef]  

38. D. Jaksch, J. I. Cirac, P. Zoller, S. L. Rolston, R. Côté, and M. D. Lukin, “Fast quantum gates for neutral atoms,” Phys. Rev. Lett. 85(10), 2208–2211 (2000). [CrossRef]  

39. C. Ates, T. Pohl, T. Pattard, and J. M. Rost, “Antiblockade in rydberg excitation of an ultracold lattice gas,” Phys. Rev. Lett. 98(2), 023002 (2007). [CrossRef]  

40. D. Møller, L. B. Madsen, and K. Mølmer, “Quantum gates and multiparticle entanglement by rydberg excitation blockade and adiabatic passage,” Phys. Rev. Lett. 100(17), 170504 (2008). [CrossRef]  

41. T. Amthor, C. Giese, C. S. Hofmann, and M. Weidemüller, “Evidence of antiblockade in an ultracold rydberg gas,” Phys. Rev. Lett. 104(1), 013001 (2010). [CrossRef]  

42. L. Isenhower, E. Urban, X. L. Zhang, A. T. Gill, T. Henage, T. A. Johnson, T. G. Walker, and M. Saffman, “Demonstration of a neutral atom controlled-not quantum gate,” Phys. Rev. Lett. 104(1), 010503 (2010). [CrossRef]  

43. D. Barredo, S. Ravets, H. Labuhn, L. Béguin, A. Vernier, F. Nogrette, T. Lahaye, and A. Browaeys, “Demonstration of a strong rydberg blockade in three-atom systems with anisotropic interactions,” Phys. Rev. Lett. 112(18), 183002 (2014). [CrossRef]  

44. D. Petrosyan and K. Mølmer, “Binding potentials and interaction gates between microwave-dressed rydberg atoms,” Phys. Rev. Lett. 113(12), 123003 (2014). [CrossRef]  

45. D. D. B. Rao and K. Mølmer, “Robust rydberg-interaction gates with adiabatic passage,” Phys. Rev. A 89(3), 030301 (2014). [CrossRef]  

46. S.-L. Su, E. Liang, S. Zhang, J.-J. Wen, L.-L. Sun, Z. Jin, and A.-D. Zhu, “One-step implementation of the rydberg-rydberg-interaction gate,” Phys. Rev. A 93(1), 012306 (2016). [CrossRef]  

47. X.-F. Shi, “Rydberg quantum gates free from blockade error,” Phys. Rev. Appl. 7(6), 064017 (2017). [CrossRef]  

48. S.-L. Su, Y. Gao, E. Liang, and S. Zhang, “Fast rydberg antiblockade regime and its applications in quantum logic gates,” Phys. Rev. A 95(2), 022319 (2017). [CrossRef]  

49. X.-F. Shi, “Deutsch, toffoli, and cnot gates via rydberg blockade of neutral atoms,” Phys. Rev. Appl. 9(5), 051001 (2018). [CrossRef]  

50. S. L. Su, H. Z. Shen, E. Liang, and S. Zhang, “One-step construction of the multiple-qubit rydberg controlled-phase gate,” Phys. Rev. A 98(3), 032306 (2018). [CrossRef]  

51. M. Saffman and K. Mølmer, “Efficient multiparticle entanglement via asymmetric rydberg blockade,” Phys. Rev. Lett. 102(24), 240502 (2009). [CrossRef]  

52. X. L. Zhang, L. Isenhower, A. T. Gill, T. G. Walker, and M. Saffman, “Deterministic entanglement of two neutral atoms via rydberg blockade,” Phys. Rev. A 82(3), 030306 (2010). [CrossRef]  

53. T. Wilk, A. Gaëtan, C. Evellin, J. Wolters, Y. Miroshnychenko, P. Grangier, and A. Browaeys, “Entanglement of two individual neutral atoms using rydberg blockade,” Phys. Rev. Lett. 104(1), 010502 (2010). [CrossRef]  

54. D. D. B. Rao and K. Mølmer, “Dark entangled steady states of interacting rydberg atoms,” Phys. Rev. Lett. 111(3), 033606 (2013). [CrossRef]  

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

56. S. Möbius, M. Genkin, A. Eisfeld, S. Wüster, and J. M. Rost, “Entangling distant atom clouds through rydberg dressing,” Phys. Rev. A 87(5), 051602 (2013). [CrossRef]  

57. S.-L. Su, Q. Guo, H.-F. Wang, and S. Zhang, “Simplified scheme for entanglement preparation with rydberg pumping via dissipation,” Phys. Rev. A 92(2), 022328 (2015). [CrossRef]  

58. M. Ostmann, J. Minář, M. Marcuzzi, E. Levi, and I. Lesanovsky, “Non-adiabatic quantum state preparation and quantum state transport in chains of rydberg atoms,” New J. Phys. 19(12), 123015 (2017). [CrossRef]  

59. S.-L. Su, Y. Tian, H. Z. Shen, H. Zang, E. Liang, and S. Zhang, “Applications of the modified rydberg antiblockade regime with simultaneous driving,” Phys. Rev. A 96(4), 042335 (2017). [CrossRef]  

60. J. Song, C. Li, Z.-J. Zhang, Y.-Y. Jiang, and Y. Xia, “Implementing stabilizer codes in noisy environments,” Phys. Rev. A 96(3), 032336 (2017). [CrossRef]  

61. 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]  

62. I. I. Beterov, G. N. Hamzina, E. A. Yakshina, D. B. Tretyakov, V. M. Entin, and I. I. Ryabtsev, “Adiabatic passage of radio-frequency-assisted förster resonances in rydberg atoms for two-qubit gates and the generation of bell states,” Phys. Rev. A 97(3), 032701 (2018). [CrossRef]  

63. 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]  

64. A.-X. Chen, “Implementation of deutsch-jozsa algorithm and determination of value of function via rydberg blockade,” Opt. Express 19(3), 2037–2045 (2011). [CrossRef]  

65. H. Weimer, M. Muller, I. Lesanovsky, P. Zoller, and H. P. Buchler, “A rydberg quantum simulator,” Nat. Phys. 6(5), 382–388 (2010). [CrossRef]  

66. Y. Han, B. He, K. Heshami, C.-Z. Li, and C. Simon, “Quantum repeaters based on rydberg-blockade-coupled atomic ensembles,” Phys. Rev. A 81(5), 052311 (2010). [CrossRef]  

67. X. Q. Shao, D. X. Li, Y. Q. Ji, J. H. Wu, and X. X. Yi, “Ground-state blockade of rydberg atoms and application in entanglement generation,” Phys. Rev. A 96(1), 012328 (2017). [CrossRef]  

68. T. G. Walker and M. Saffman, “Consequences of zeeman degeneracy for the van der waals blockade between rydberg atoms,” Phys. Rev. A 77(3), 032723 (2008). [CrossRef]  

69. F. Nogrette, H. Labuhn, S. Ravets, D. Barredo, L. Béguin, A. Vernier, T. Lahaye, and A. Browaeys, “Single-atom trapping in holographic 2d arrays of microtraps with arbitrary geometries,” Phys. Rev. X 4(2), 021034 (2014). [CrossRef]  

70. D. W. Schönleber, A. Eisfeld, M. Genkin, S. Whitlock, and S. Wüster, “Quantum simulation of energy transport with embedded rydberg aggregates,” Phys. Rev. Lett. 114(12), 123005 (2015). [CrossRef]  

71. D. Barredo, S. de Léséleuc, V. Lienhard, T. Lahaye, and A. Browaeys, “An atom-by-atom assembler of defect-free arbitrary two-dimensional atomic arrays,” Science 354(6315), 1021–1023 (2016). [CrossRef]  

72. M. Endres, H. Bernien, A. Keesling, H. Levine, E. R. Anschuetz, A. Krajenbrink, C. Senko, V. Vuletic, M. Greiner, and M. D. Lukin, “Atom-by-atom assembly of defect-free one-dimensional cold atom arrays,” Science 354(6315), 1024–1027 (2016). [CrossRef]  

73. D. W. Schönleber, C. D. B. Bentley, and A. Eisfeld, “Engineering thermal reservoirs for ultracold dipole-dipole-interacting rydberg atoms,” New J. Phys. 20(1), 013011 (2018). [CrossRef]  

74. X.-F. Zhang, Q. Sun, Y.-C. Wen, W.-M. Liu, S. Eggert, and A.-C. Ji, “Rydberg polaritons in a cavity: A superradiant solid,” Phys. Rev. Lett. 110(9), 090402 (2013). [CrossRef]  

75. A. Gaëtan, Y. Miroshnychenko, T. Wilk, A. Chotia, M. Viteau, D. Comparat, P. Pillet, A. Browaeys, and P. Grangier, “Observation of collective excitation of two individual atoms in the rydberg blockade regime,” Nat. Phys. 5(2), 115–118 (2009). [CrossRef]  

76. 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]  

77. F. Brennecke, T. Donner, S. Ritter, T. Bourdel, M. Köhl, and T. Esslinger, “Cavity qed with a bose-einstein condensate,” Nature 450(7167), 268–271 (2007). [CrossRef]  

78. A. Grankin, E. Brion, E. Bimbard, R. Boddeda, I. Usmani, A. Ourjoumtsev, and P. Grangier, “Quantum statistics of light transmitted through an intracavity rydberg medium,” New J. Phys. 16(4), 043020 (2014). [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 (5)

Fig. 1.
Fig. 1. The total scheme preparing the $N$-qubit GHZ state composes of $N-1$ steps, which can be divided into two groups: Step 1 and Step $n~(n=2,3,\ldots ,N-1)$. For the Step 1, only the first two atoms will be addressed individually by three lasers, respectively. As for the Step $n$, we only coupled the $n$-th atom to one laser and the $(n+1)$-th atom to three lasers, respectively.
Fig. 2.
Fig. 2. (a) The shapes of pulses to prepare the 3-qubit GHZ state. For Step 1 ($\Omega _c t\in [0,6000]$), we set $t_c=6000/\Omega _c$. For Step 2 ($\Omega _c t\in (6000,14000]$), we set $t_c=8000/\Omega _c$. (b) The shapes of pulses to prepare the 4-qubit GHZ state. For Step 1 ($\Omega _c t\in [0,6000]$), we set $t_c=6000/\Omega _c$. For Steps 2 and 3 ($\Omega _c t\in (6000,14000]$ and $(14000,22000]$), we set $t_c=8000/\Omega _c$. (c) and (d) are the fidelity of the $3$- and $4$-qubit states governed by the original Hamiltonian and the effective Hamiltonian, where the fidelity of state $\rho _i=|i\rangle \langle i|$ is defined as $F=\textrm {Tr}\sqrt {\rho _i^{1/2}\rho (t)\rho _i^{1/2}}$ and $\rho (t)$ is the density matrix of system at time $t$. The other relevant parameters are all chosen as: $\Omega _a=\Omega _b=\sqrt {0.05}\Omega _c$, $\Delta _p=20\Omega _c$, $\Delta _r=20\Omega _c$, $T=0.15t_c$, and $\tau =0.1t_c$.
Fig. 3.
Fig. 3. The dynamical evolution of the fidelity for the 3-qubit GHZ state with different $\delta$. For the Step 1 and Step 2, we set $t_c=6000/\Omega _c$ and $t_c=8000/\Omega _c$, respectively. The other relevant parameters are $\Omega _{a}=\Omega _b=\sqrt {2}\Omega _c$, $\Delta _p=800\Omega _c$, $\Delta _r=20\Omega _c$, $T=0.15t_c$, and $\tau =0.1t_c$.
Fig. 4.
Fig. 4. The dynamical evolution of the fidelity for the 3-qubit GHZ state with different $\Delta _r$. For the Step 1 and Step 2, we set $t_c=6000/\Omega _c$ and $t_c=8000/\Omega _c$, respectively. The other relevant parameters are $\Omega _{a}=\Omega _b=\sqrt {2}\Omega _c$, $\Delta _p=800\Omega _c$, $\delta =0$, $T=0.15t_c$, and $\tau =0.1t_c$.
Fig. 5.
Fig. 5. The fidelity of the $3$-qubit GHZ state governed by the original master equation. The decay rates are $\gamma _p=3\Omega _c$ and $\gamma =0.001\Omega _c$. The other relevant parameters are the same as those of Fig. 2(c).

Equations (19)

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

Ω a ( t ) = Ω a exp [ ( t t c / 2 τ ) 2 T 2 ] ,
Ω b ( t ) = Ω b exp [ ( t t c / 2 + τ ) 2 T 2 ] ,
H I 1 ( t ) = j = 1 , 2 Ω a ( t ) | p j 0 | e i Δ p t + Ω b ( t ) | p j 1 | e i Δ p t + Ω c | r j 1 | e i Δ r t + H.c. + α β U α β | r r α β r r | ,
H I 1 ( t ) = j = 1 , 2 Ω a ( t ) | p j 0 | + Ω b ( t ) | p j 1 | + Ω c | r j 1 | + H.c. Δ p | p j p | Δ r | r j r | + k = 1 N 1 U k , k + 1 | r r k , k + 1 r r | .
H I 1 ( t ) = 2 Ω a ( t ) | ψ 1 12 00 | + Ω b ( t ) | ψ 1 12 ψ 0 | + Ω a ( t ) | ψ 2 12 ψ 0 | + 2 Ω b ( t ) | ψ 2 12 11 | + Ω | 11 12 r r | + H.c. Δ p | ψ 1 12 ψ 1 | Δ p | ψ 2 12 ψ 2 | ,
H I 1 ( t ) = 2 Ω a ( t ) | ψ 1 12 00 | + Ω b ( t ) | ψ 1 12 ψ 0 | + Ω a ( t ) | ψ 2 12 ψ 0 | + Ω b ( t ) | ψ 2 12 ( + | + | ) + H.c. + Ω ( | + 12 + | | 12 | ) Δ p | ψ 1 12 ψ 1 | Δ p | ψ 2 12 ψ 2 | ,
H eff 1 ( t ) = 2 Ω a ( t ) | ψ 1 12 00 | + Ω b ( t ) | ψ 1 12 ψ 0 | + H.c. Δ p | ψ 1 12 ψ 1 | .
| Φ 12 = cos [ Θ ( t ) ] | 00 12 sin [ Θ ( t ) ] | ψ 0 12 ,
lim t cos [ Θ ( t ) ] = 1 , lim t + cos [ Θ ( t ) ] = 0.
H I n ( t ) = j = n , n + 1 ( Ω c | r j 1 | + H.c. Δ r | r j r | ) + Ω a ( t ) | p n + 1 0 | + Ω b ( t ) | p n + 1 1 | + H.c. Δ p | p n + 1 p | + U n , n + 1 | r r n , n + 1 r r | ,
H eff n ( t ) = Ω a ( t ) | 0 p n , n + 1 00 | + Ω b ( t ) | 0 p n , n + 1 01 | + H.c. Δ p | 0 p n , n + 1 0 p | .
| Φ n , n + 1 = cos [ Θ ( t ) ] | 00 n , n + 1 sin [ Θ ( t ) ] | 01 n , n + 1 .
H I 1 ( t ) = 2 Ω a ( t ) | ψ 1 12 00 | + Ω b ( t ) | ψ 1 12 ψ 0 | + Ω a ( t ) | ψ 2 12 ψ 0 | + 2 Ω b ( t ) | ψ 2 12 11 | + Ω | 11 12 r r | + H.c. Δ p | ψ 1 12 ψ 1 | Δ p | ψ 2 12 ψ 2 | + δ | r r 12 r r | .
H I 1 ( t ) = 2 Ω a ( t ) | ψ 1 12 00 | + Ω b ( t ) | ψ 1 12 ψ 0 | + Ω a ( t ) | ψ 2 12 ψ 0 | + Ω b ( t ) | ψ 2 12 ( sin θ + ~ | cos θ ~ | ) + H.c. + Ω ~ + | + ~ 12 + ~ | + Ω ~ | ~ 12 ~ | Δ p | ψ 1 12 ψ 1 | Δ p | ψ 2 12 ψ 2 | .
ρ ˙ = i [ H I j ( t ) , ρ ] + L j ρ + L j + 1 ρ ,
L j ρ = k = 1 3 L j k ρ L j k 1 2 ( L j k L j k ρ + ρ L j k L j k ) ,
L j 1 = γ p 2 | 0 j p | ,
L j 2 = γ p 2 | 1 j p | ,
L j 3 = γ | 1 j r | .
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.