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

Heralded quantum multiplexing entanglement between stationary qubits via distribution of high-dimensional optical entanglement

Open Access Open Access

Abstract

Quantum entanglement between pairs of remote quantum memories (QMs) is a prerequisite for realizing many applications in quantum networks. Here, we present a heralded protocol for the parallel creation of quantum entanglement among multiple pairs of QMs placed in spatially separated nodes, where each QM, encoding a stationary qubit, couples to an optical cavity and deterministically interacts with single photons. Our protocol utilizes an entangled photon pair encoded in the high-dimensional time-bin degree of freedom to simultaneously entangle multiple QM pairs, and is efficient in terms of reducing the time consumption and photon loss during transmission. Furthermore, our approach can be extended to simultaneously support spatial-temporal multiplexing, as its success is heralded by the detection of single photons. These distinguishing features make our protocol particularly useful for long-distance quantum communication and large-scale quantum networks.

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

1. Introduction

Quantum entanglement is a unique phenomenon in quantum physics and enables many quantum technologies that surpass their classical counterparts [18]. It is worth noticing that large-scale quantum communication and distributed quantum computation require quantum entanglement between distant nodes in quantum networks [912]. However, quantum entanglement could only be created locally [13,14]. The propagation of photonic pulses becomes a prerequisite for creating nonlocal quantum entanglement between remote quantum nodes, owing to the stable and robust property of photons to transmit quantum information across long distances [15,16]. In practice, the transmission loss in optical fibers or free-space channels sets a limit on the efficiency and range of entanglement generation between distant nodes [17]. This bottleneck can be overcome by dividing the long entangled channel into multiple intermediate segments and then creating entanglement within each segment, a crucial element of quantum repeaters [1820].

In general, there are three methods for generating entanglement between spatially separated stationary quantum memories (QMs) [2123]. The first method entangles two remote QMs by creating a hybrid entangled state between a photonic pulse and a QM at each node, followed by the interference of the two photonic pulses at an intermediate node [2428]. The second approach generates hybrid entanglement between a QM and a photonic pulse, which is then sent to interact with another QM to entangle the two QMs [2931]. The third method utilizes an entangled photon source and sends two photons from each entangled pair to two separate nodes to entangle two remote QMs [3237]. In these approaches, a photonic pulse (single photon) or an entangled photon pair is transmitted over a distance equal to the separation between two nodes to entangle a pair of QMs. They require the transmission of many single photons or photon pairs, as most applications require multiple pairs of entangled QMs, such as quantum error correction [38,39] and multiqubit state teleportation [40].

Quantum multiplexing encodes a single photon with more information than a single qubit [41], and it has been introduced to entangle two pairs of QMs using only a single photonic pulse, encoding in both the polarization and the time-bin degrees of freedom (DOFs) [42,43]. This multiplexing approach takes advantage of the fact that a photonic pulse possesses multiple DOFs, enlarging the corresponding Hilbert space. It can also be extended to entangle additional pairs of QMs by encoding a photonic pulse with more DOFs [42]. Apart from qubits hyperencoded in multiple DOFs [41], qudits (i.e. high-dimensional quantum systems) can also form high-dimensional Hilbert spaces and are implementable in practical experiments. By introducing enhanced-capacity quantum multiplexing, which encodes a single photon in both polarization and high-dimensional time-bin DOFs, Xie et al.,[44] presented a general protocol for the parallel entanglement of multiple pairs of QMs using a single photon. Zhou et al., [45] showed that the efficiency of nonlocal multiqubit entanglement generation can be significantly improved. Meanwhile, parallel entanglement can be achieved using time-bin photonic qudits without the need for polarization encoding [46]. These protocols in principle entangle several QM qubits in one node with a high-dimensionally encoded single photon and then evolve the hybrid entanglement between QM qubits and the photonic qudit into entanglement between multiple pairs of QMs, after the photon reaches the other node and interacts sequentially with QMs in that location. Therefore, these protocols impose a constraint on the coherence time of QMs in the initial node and also constrain their ability to achieve spatial-temporal multiplexing [18]. Meanwhile, the polarization DOF of the photon is susceptible to polarization channel noise, leading to a reduction of hybrid entanglement. This, in turn, decreases the entanglement between QMs. A potential approach to address this issue is to employ advanced entanglement purification protocols [47,48].

Alternatively, it is worthwhile to generalize the original quantum multiplexing entangling protocol [42] to include optical entanglement, i.e., a quantum multiplexing version of the entangling method that makes use of entangled photon sources. This approach has the potential to efficiently create multiple entangled pairs of QMs in combination with the higher ability to spatial-temporal multiplexing with lower QM coherence times. In comparison to two-dimensional optical entanglement, high-dimensional entanglement offers more pronounced benefits [49,50], including robustness against noise [5155], heightened communication capacity [5658], and enhanced security in quantum cryptography [59,60]. It has been demonstrated that high-dimensional entanglement can be used to efficiently transmit information over long distances [61]. Moreover, there are multiple DOFs of a single photon that can be harnessed for qudit encoding [6264], such as spatial modes, frequency modes, and time bins. The time-bin photonic qudit is particularly suitable for long-distance transmission, owing to its resilience against decoherence and transmission noise, and it is easy to obtain high-dimensional time-bin entanglement [65,66].

In this article, we present a heralded scheme for the parallel creation of entanglement between multiple pairs of QMs by utilizing a pair of high-dimensional time-bin entangled photons. By distributing two photons from each entangled pair to two remote nodes and interacting each photon with the QMs in its respective node, the parallel entanglement between multiple pairs of nonlocal QMs can be created in a heralded way, signaled by the detection of one photon in each node. Therefore, our protocol has the potential to be expanded to simultaneously accommodate both spatial-temporal multiplexing and quantum multiplexing. In our protocol, each photon interacts with $n$ QMs to entangle $n$ pairs of QMs and travels half the distance between the nodes. This reduces the time required, when compared to protocols where photons traverse the entire channel and interact with all QMs. Furthermore, our protocol can potentially be extended to entangle additional pairs of QMs by increasing time-bin dimensions. These distinguishing features make our protocol efficient to entangle multiple pairs of nonlocal QMs and thus provide a viable route toward the construction of large-scale quantum networks.

2. Controlled-polarization flip unit based on the spin-photon interface

An interface between single photons and individual spins [67] is an essential platform for constructing quantum networks and distributed quantum computation [6870]. Such an interface can be implemented by coupling a four-level atom to a single-sided optical cavity, shown in Fig. 1, where the frequency of the cavity mode $\omega _c$ is nearly resonant with the transition from coupled ground state $\vert g\rangle$ to excited state $\vert \tilde {g}\rangle$ at a frequency $\omega _0$. Meanwhile, the transition $\vert e\rangle \leftrightarrow \vert \tilde {e}\rangle$ at a frequency $\omega _1$ decouples from the cavity mode, either due to a large detuning or polarization mismatch. A polarized photon with frequency $\omega \simeq \omega _c$ will be reflected by the atom-cavity system with a state-dependent reflection coefficient [7173]

$$r_k=1-\frac{2(1-i\Delta_k)}{(1-i\Delta_c)(1-i\Delta_k)+2C_k},$$
where the cooperativity is $C_k=g_k^2/\gamma _k\kappa$ for $k\in \{g,e\}$, $g_k$ represents the corresponding coupling rate between the cavity and the atom in state $\vert k\rangle$, $\kappa$ is the cavity decay rate, and $\gamma _k$ denotes the spontaneous decay rate. The detunings are defined as $\Delta _c=(\omega -\omega _{c})/\kappa$, $\Delta _g=(\omega -\omega _0)/\gamma _g$, and $\Delta _e=(\omega -\omega _1)/\gamma _e$. In principle, we have $r_g\simeq -r_e\simeq 1$ for $\Delta _c=\Delta _g=0$, $C_g\gg 1$, and $C_e/\Delta _e\simeq 0$.

A controlled-polarization flip unit can be constructed by using such an atom-cavity system and two circularly polarizing beam splitters (CPBSs) [4144,72], shown in Fig. 1(b). The CPBS reflects the right-circularly polarized photon $\vert R\rangle$ that matches the transition $\vert g\rangle \leftrightarrow \vert \tilde {g}\rangle$, and transmits left-circularly polarized photon $\vert L\rangle$. For a polarized photon in state $\vert D\rangle =\frac {1}{\sqrt {2}}(\vert L\rangle +\vert R\rangle )$ or $\vert A\rangle =\frac {1}{\sqrt {2}}(\vert L\rangle -\vert R\rangle )$ impinging into the input port of the unit, the output state of the combined system consisting of the photon and the atom can be described as follows:

$$\begin{aligned} &\vert D\rangle \vert g\rangle\longrightarrow\vert D\rangle \vert g\rangle,\qquad \vert D\rangle \vert e\rangle\longrightarrow\vert A\rangle \vert e\rangle,\\ &\vert A\rangle \vert g\rangle\longrightarrow\vert A\rangle \vert g\rangle,\qquad \vert A\rangle \vert e\rangle\longrightarrow\vert D\rangle \vert e\rangle, \end{aligned}$$
where the photon polarization is flipped only when the atom is in state $\vert e\rangle$. We refer to this unit as a QM below, and it constitutes an elementary building block for creating nonlocal entanglement.

 figure: Fig. 1.

Fig. 1. (a) Level structure and transitions of a four-level atom. (b) Schematic of a controlled-polarization flip unit. CPBS represents a circular-polarization beam splitter that reflects the right-circularly polarized photon $\vert R\rangle$ and transmits left-circularly polarized photon $\vert L\rangle$.

Download Full Size | PDF

3. Heralded parallel protocol for entangling two pairs of QMs

In this section, we show that parallel entanglement between distant stationary qubits can be achieved by distributing high-dimensional optical entanglement, and present a specific scheme for entangling two pairs of QMs placed in two remote nodes, as shown in Fig. 2. The TBES represents a time-bin entangled source that prepares pairs of polarized photons entangled in four-dimensional time-bin DOF. OS$_i$ ($i=1, 2$) are optical switches that combine photon components in different spatial modes with different time bins into one spatial mode, and vice versa [74,75]. The H is a half-wave plate that flips the polarization of photons passing through it. The PBS is a polarizing beam splitter that reflects $A$-polarized photons and transmits $D$-polarized ones, while the PBS$'$ is a polarizing beam splitter that reflects $D$-polarized photons and transmits $A$-polarized ones. The two time-bin entangled photons prepared by the TBES are sent to nodes A and B, respectively. They experience parallel operation procedures in the two nodes, which can lead to entanglement between QM1 (QM3) and QM2 (QM4). The principle of this process can be detailed as follows.

 figure: Fig. 2.

Fig. 2. Schematics of generating two pairs of QMs via distribution of optical entanglement. (a) Step-by-step protocol with (b) circuit of PSE for photon state exchanges. Here QM$k$($k$=1,…,4) denotes a quantum memory. PC is a Pockel cell that flips the polarization of photons in the last two time-bin modes. OS$_1$ is an optical switch that directs photons in the time-bin states $\vert 0\rangle$ ($\vert 1\rangle$) through the upper (lower) path. OS$_2$ directs two spatial modes of a photon with different time bins into one spatial mode. $\mathcal {T}{(i)}$ introduces an optical delay of $it_\Delta$. H is a half-wave plate and flips the polarization. PBS is polarization beam splitter that transmits the $D$-polarized photon and reflects the $A$-polarized photon, while PBS$'$ transmits the $A$-polarized photon and reflects the $D$-polarized photon.

Download Full Size | PDF

Suppose QM$k$ ($k=1,\ldots,4$) is initialized to superposition state $\vert \Psi _{k}\rangle =(\vert g_k\rangle +\vert e_k\rangle )/\sqrt {2}$, and one pair of $D$-polarized photons p$1$p$2$ are initialized to the time-bin-entangled state $\vert \Psi _{P_0} \rangle =\frac {1}{2}\sum _{i=0}^{3}\mathcal {T}_1{(i)}\mathcal {T}_2{(i)}\vert DD\rangle$, where time operator $\mathcal {T}_{1(2)}{(i)}$ introduces a time delay of $it_{\Delta }$ for photons p$1$ and p$2$, which satisfies the relationship $\mathcal {T}_{1(2)}{(l+p)}=\mathcal {T}_{1(2)}{(l)}\mathcal {T}_{1(2)}{(p)}$ [65,66]. The two photons p$1$ and p$2$ are sent to node A and node B, respectively. The time-bin entangled state $\vert \Psi _{P_0} \rangle$ is converted into a hyper-entangled state $\vert \Psi _{P_1} \rangle =\frac {1}{\sqrt {2}}\sum _{i=0}^{1}\mathcal {T}_1{(i)}\mathcal {T}_2{(i)}\vert \phi _P\rangle$ with $\vert \phi _P\rangle =(|DD\rangle +|AA\rangle )/\sqrt {2}$ by respectively passing p$1$ and p$2$ through a converting unit consisting of a Pockel cell (PC) and an unbalanced polarizing Mach–Zehnder interferometer, shown in Fig. 2(a). Here PC flips the polarization of photons in the last two time-bin modes, and the two time delays of photons in state $\vert AA\rangle$ are compensated after they pass through the short arm of the interferometer.

The entangled photons p$1$ and p$2$ then interact with QM1 and QM2, respectively. Their polarization states are entangled with QM1 and QM2, since photon p$1$ (p$2$) flips its polarization when QM1 (QM2) is in state $|e\rangle$. Subsequently, the photon states are exchanged as $\mathcal {T}{(0)} |A\rangle \leftrightarrow \mathcal {T}{(1)}|D\rangle$ by passing the photons through the PSE circuit, consisting of OS$_{1,2}$, PBSs, PBS$'$s, H, and time delayers $\mathcal {T}{(1)}$. The combined state of QM1,2 and photons p$1$p$2$ then evolves into

$$\vert \Phi_1 \rangle=\frac{1}{\sqrt{2}}(\phi_T^{0,1}|\phi_{12}\rangle+\psi_T^{0,1}|\psi_{12}\rangle)|\phi_P\rangle,$$
where the polarization state of photons p$1$p$2$ is recovered for entangling the next QM pair, $|\phi _{12}\rangle =(\vert g_1g_2\rangle +\vert e_1e_2\rangle )/\sqrt {2}$ and $|\psi _{12}\rangle =(\vert g_1e_2\rangle +\vert e_1g_2\rangle )/\sqrt {2}$ represent two entangled states of QM1 and QM2. Operators $\phi ^{i,j}_{T}$ and $\psi ^{i,j}_{T}$ introduce delays for two photons in separable time-bin states and generate time-bin entangled states with
$$\begin{aligned} &\phi^{i,j}_{T}=\frac{1}{\sqrt{2}}[\mathcal{T}_1{(i)}\mathcal{T}_2{(i)}+\mathcal{T}_1{(j)}\mathcal{T}_2{(j)}],\\ &\psi^{i,j}_{T}=\frac{1}{\sqrt{2}}[\mathcal{T}_1{(i)}\mathcal{T}_2{(j)}+\mathcal{T}_1{(j)}\mathcal{T}_2{(i)}]. \end{aligned}$$

The two spatial modes of photons p$1$ and p$2$ are combined into one spatial mode by OS$_{2}$ at the output port of each PSE, which is then directed to interact with QM3 and QM4, respectively. The combined state of the two photons and four QMs evolves into a hybrid entangled state

$$\begin{aligned} \vert \Phi_2 \rangle&=\frac{1}{2}(\phi_T^{0,1}|\phi_{12}\rangle|\phi_{34}\rangle|\phi_P\rangle+\psi_T^{0,1}|\psi_{12}\rangle|\phi_{34}\rangle|\phi_P\rangle\\ &+\phi_T^{0,1}|\phi_{12}\rangle|\psi_{34}\rangle|\psi_P\rangle+\psi_T^{0,1}|\psi_{12}\rangle|\psi_{34}\rangle|\psi_P\rangle), \end{aligned}$$
with two entangled states of QM3 and QM4, $|\phi _{34}\rangle =(\vert g_3g_4\rangle +\vert e_3e_4\rangle )/\sqrt {2}$ and $|\psi _{34}\rangle =(\vert g_3e_4\rangle +\vert e_3g_4\rangle )/\sqrt {2}$, and two photon entangled states $\vert \psi _P\rangle =(|DA\rangle +|AD\rangle )/\sqrt {2}$. Within each pair, two QMs are entangled, and these entangled QMs further entangle with photons p$1$p$2$. Meanwhile, the QMs in one pair remain decoupled from the QMs in the other pair. (QM1 and QM2) and (QM3 and QM4) will be projected into a product state of two entangled states ($|\phi _{k,k+1}\rangle$ or $|\psi _{k,k+1}\rangle$, where $k=1,3$), based on the state of time-bin and polarization of p$1$p$2$, when photon p$1$ (p$2$) is measured in the polarization basis $\{\vert D\rangle$, $\vert A\rangle \}$ with time-resolved photon detectors. That is, two pairs of remote QMs can be prepared into entangled states in a heralded way in parallel by distributing one pair of four-dimensional time-bin entangled photons.

4. Heralded parallel protocol for entangling multiple pairs of QMs

Our protocol can be generalized to entangle more pairs of QMs placed in two remote nodes in parallel using one pair of photons entangled in the high-dimensional time-bin DOF. A parallel schematic for entangling three pairs of QMs, i.e., (QM1, QM2), (QM3, QM4), and (QM5, QM6) is shown in Fig. 3, where all QMs are all initialized to superposition states $\vert \Psi _{k}\rangle =(\vert g_k\rangle +\vert e_k\rangle )/\sqrt {2}$ ($k=1,\ldots,6$). The TBES produces a pair of $D$-polarized photons p$1$p$2$ that are entangled in the time-bin mode with $\vert \Psi _{P_2} \rangle =\frac {1}{2\sqrt {2}}\sum _{i=0}^{7}\mathcal {T}_1{(i)}\mathcal {T}_2{(i)}\vert DD\rangle$. Photons p$1$p$2$ are separated and distributed to nodes A and B, respectively. They pass through the same photonic circuit placed in the corresponding node and experience parallel operation procedures as detailed below.

 figure: Fig. 3.

Fig. 3. Schematics of generating three pairs of QMs via distribution of optical entanglement. (a) Step-by-step protocol with (b) and (c) circuits of PSE$_1$ and PSE$_2$ for photon state exchanges. Here QM$k$($k$=1,…,6) denotes a quantum memory. PC$'$ is a Pockel cell that flips the polarization of photons in the last four time-bin modes. OS$_1$ (OS$_3$) is an optical switch that directs photons in the time-bin states $\vert 0\rangle$ and $\vert 2\rangle$ ($\vert 0\rangle$ and $\vert 1\rangle$) to the upper path and photons in states $\vert 1\rangle$ and $\vert 3\rangle$ ($\vert 2\rangle$ and $\vert 3\rangle$) to the lower path. OS$_2$ (OS$_4$) directs two spatial modes of a photon with different time bins into one spatial mode. H is a half-wave plate and flips the polarization. $\mathcal {T}{(i)}$ describes a time delayer, introducing an optical delay of $it_\Delta$. PBS is polarization beam splitter that transmits the D-polarized photon and reflects the A-polarized photon. PBS$'$ transmits the A-polarized photon and reflects the D-polarized photon.

Download Full Size | PDF

Upon arriving at nodes A and B, the state of photons p$1$p$2$ is transformed into a hyper-entangled state $\vert \Psi _{P_3} \rangle =\frac {1}{2}\sum _{i=0}^{3}\mathcal {T}_1{(i)}\mathcal {T}_2{(i)}\vert \phi _P\rangle$ utilizing a converting unit similar to the one used for entangling two pairs of QMs, while the PC′ now flips the polarization of photons in the last four time-bin modes and four time delays are compensated by passing them through the short arm of the interferometer. Photons p$1$ and p$2$ are then directed to interact with QM1 and QM2, resulting in their entanglement. Subsequently, PSE$_1$, shown in Fig. 3(b), engineers photon states and completes the exchange $\mathcal {T}{(0)} |A\rangle \leftrightarrow \mathcal {T}{(1)} |D\rangle$ and $\mathcal {T}{(2)} |A\rangle \leftrightarrow \mathcal {T}{(3)} |D\rangle$, by directing photons in time-bin states ($\vert 0\rangle$, $\vert 2\rangle$) through the upper path and those in states ($\vert 1\rangle$, $\vert 3\rangle$) through the lower path with OS$_1$. After two spatial modes with different time bins are combined into one spatial mode by OS$_2$ for each photon, the combined state of p$1$p$2$ and QM1,2 can be described as

$$\vert\Xi_1\rangle=\frac{1}{2}[(\phi_T^{0,1}+\phi_T^{2,3})|\phi_{12}\rangle+(\psi_T^{0,1}+\psi_T^{2,3})|\psi_{12}\rangle]|\phi_P\rangle,$$
where time-bin entangled operators $\phi _T^{0,1}$, $\phi _T^{2,3}$, $\psi _T^{0,1}$, and $\psi _T^{2,3}$ are defined in the same manner as in Eq. (4). QM1, QM2, and time-bin modes of p$1$p$2$ are entangled, while the polarization of p$1$p$2$ evolves into the original entangled state $|\phi _P\rangle$ for entangling the next QM pair.

The photons then interact with QM3 and QM4, and pass through PSE$_2$ that engineers photon states as $\mathcal {T}{(0)} |A\rangle \leftrightarrow \mathcal {T}{(2)} |D\rangle$ and $\mathcal {T}{(1)} |A\rangle \leftrightarrow \mathcal {T}{(3)} |D\rangle$, by directing photons in time-bin states ($\vert 0\rangle$, $\vert 1\rangle$) to the upper path and photons in states ($\vert 2\rangle$, $\vert 3\rangle$) to the lower path of PSE$_2$ with OS$_3$. The combined state of two photons and four QMs evolves into

$$\begin{aligned} \vert\Xi_2\rangle&=\frac{1}{2\sqrt{2}}[(\phi_T^{0,1}+\phi_T^{2,3})|\phi_{12}\rangle|\phi_{34}\rangle+(\psi_T^{0,1}+\psi_T^{2,3})|\psi_{12}\rangle|\phi_{34}\rangle\\ &+ (\psi_T^{0,2}+\psi_T^{1,3})|\phi_{12}\rangle|\psi_{34}\rangle+(\psi_T^{0,3}+\psi_T^{1,2})|\psi_{12}\rangle|\psi_{34}\rangle]|\phi_P\rangle, \end{aligned}$$
where two QMs within each pair are entangled, and these entangled QMs further entangle with time-bin modes of photons p$1$p$2$. Meanwhile, the QMs in one pair remain decoupled from the QMs in the other pair, and p$1$p$2$ also evolve into state $|\phi _P\rangle$ for entangling the next QM pair.

To entangle the third pair of QMs, photons p$1$ and p$2$ then interact with QM5 and QM6, respectively. The combined state of three pairs of QMs and two photons evolves into

$$\begin{aligned} \vert \Xi_3 \rangle &=\frac{1}{4}[(\phi_T^{0,1}+\phi_T^{2,3})|\phi_{12}\rangle|\phi_{34}\rangle+(\psi_T^{0,1}+\psi_T^{2,3})|\psi_{12}\rangle|\phi_{34}\rangle+(\psi_T^{0,2}+\psi_T^{1,3})|\phi_{12}\rangle|\psi_{34}\rangle\\ &+(\psi_T^{0,3}+\psi_T^{1,2})|\psi_{12}\rangle|\psi_{34}\rangle](|\phi_{56}\rangle|\phi_P\rangle+|\psi_{56}\rangle|\psi_P\rangle),\end{aligned}$$
where $|\phi _{56}\rangle =(\vert g_5g_6\rangle +\vert e_5e_6\rangle )/\sqrt {2}$ and $|\psi _{56}\rangle =(\vert g_5e_6\rangle +\vert e_5g_6\rangle )/\sqrt {2}$ are two entangled states of QM5 and QM6. $\vert \Xi _3 \rangle$ represents a hybrid entangled state of three pairs of entangled QMs and a pair of photons in the time-bin and polarization DOFs. The entangled state of QM1QM2 and QM3QM4 is determined by the time-bin state of photons p$1$p$2$, while the entangled state of QM5QM6 is determined by the polarization state of p$1$p$2$. Three pairs of QMs will be projected into a product state of three entangled states ($|\phi _{k,k+1}\rangle$ or $|\psi _{k,k+1}\rangle$, $k=1,3,5$), when each photon is measured in the basis consisting of two polarization states $\vert A\rangle$ and $\vert D\rangle$ in combination with four time-bin states $\vert 0\rangle$, $\vert 1\rangle$, $\vert 2\rangle$, and $\vert 3\rangle$. For instance, the three pairs of QMs will be prepared to state $|\phi _{12}\rangle |\phi _{34}\rangle |\phi _{56}\rangle$ when the result of photon measurement is $|AA\rangle$ or $|DD\rangle$ with the time-bin state $\vert ii\rangle$ ($i=0, 1, 2, 3$).

5. Physical implementation with practical atoms

Nitrogen-vacancy (NV) color center, a kind of isolated point defect in diamond, is a promising candidate serving as a quantum stationary system for implementing quantum networks [7678]. The electron spins of NV centers can be optically manipulated and read out with high fidelity [79,80], and can be used to initialize and manipulate nearby nuclear spins with long coherence time for quantum storage [81,82]. So far, NV centers have been used to perform various quantum information processing tasks, such as quantum teleportation over remote distance [83,84], quantum gates on electron spins and photons [85,86], and multinode quantum networks [87].

An NV$^-$ center can be treated as an approximate four-level system [71,72]. The electronic energy level structure of an NV$^-$ center consists of three subspaces: ground state manifold (GSM), excited state manifold (ESM), and meta-stable state manifold (MSM). A static magnetic field of approximately 20 mT along the NV$^-$ axis at low temperatures causes the level $|-1\rangle$ of the GSM to be far detuned by about 2.3 GHz above $\vert g\rangle =\vert 0\rangle$ and about 1.1 GHz below $\vert e\rangle =\vert +1\rangle$ [88]. The ESM comprises six basis states $\{M_{1,\ldots,6}\}$ of the excited-state Hamiltonian. The optical transitions between GSM and ESM, coupled to the cavity mode, can be described by an effective Hamiltonian [71,73,88,89]

$$H_{\rm eff}= \sum_{s=0,\pm1}\sum_{i=1}^6 \left[\omega_s|s\rangle\langle s| + (\omega_i-i\gamma_i)|M_i\rangle\langle M_i| + g_{s,i}(\hat{a}^\dagger\hat{\sigma}_{s,i}+\hat{a}\hat{\sigma}_{s,i}^\dagger)\right]+(\omega_c-i\kappa)\hat{a}^\dagger\hat{a},$$
where $\hat {a}^\dagger$ ($\hat {a}$) is the creation (annihilation) operator of the cavity mode, $\hat {\sigma }_{s,i}=|s\rangle \langle M_i|$ is the lowering operator of the NV$^-$ center, and $g_{s,i}$ for $s=0,\pm 1$ and $i=1,\ldots,6$ are the coupling constants between the cavity and transitions, $\omega _{c}$, $\omega _{s}$, and $\omega _i$ are the frequencies of the cavity mode, the GSM levels $|s\rangle$, and the ESM levels $|M_i\rangle$. $\kappa$ is the cavity decay rate, while $\gamma _i$ denotes the decay rate of state $|M_i\rangle$, including spontaneous radiation and non-radiative transitions involving the MSM [71].

When only a $\sigma _-$ polarized photon is involved, we assume that $g_{-1,1}$, $g_{0,3}$, $g_{+1,5}$, and $g_{+1,6}$ are equal and labeled as $g$ for simplicity, while all other $g_{s,i}$ vanish due to the electric-dipole selection rules [88]. A $\sigma _-$ polarized photon impinging onto the NV$^-$-cavity system will be reflected with NV$^{-}$ state-dependent reflection coefficients. These coefficients can be obtained by solving the dynamic equations of the motion for the cavity field operator $\hat {a}$ and the NV$^{-}$ operators $\hat {\sigma }_{s,i}$ in combination with the input-output relation,

$$\begin{aligned} \frac{d\hat{a}}{dt}&=-(i\omega_c+\frac{\kappa}{2})\hat{a}-\sum_{i=1}^{6}g_{s,i}\hat{\sigma}_{s,i}-\sqrt{\kappa_c}\hat{a}_{in}+\hat{N},\\ \frac{d\hat{\sigma}_{s,i}}{dt}&=-(i\omega_{s,i}+\frac{\gamma_i}{2})\hat{\sigma}_{s,i}-g_{s,i}\hat{\sigma}^z_{s,i}\hat{a}+\hat{N}'_{s,i},\\ \hat{a}_{\rm out}&=\hat{a}_{\rm in}+\sqrt{\kappa_c}\hat{a}, \end{aligned}$$
where $\omega _{s,i}=\omega _i-\omega _s$ is the dipole transition frequency. $\kappa _{c}$ is the directional coupling rate between the cavity mode and the input and output mode. $\hat {a}_{\rm in}~(\hat {a}_{\rm out})$ represents the input (output) mode. $\hat {\sigma }^z_{s,i}=|M_i\rangle \langle M_i|-|s\rangle \langle s|$ for $s=0,\pm 1$ and $i=1,\ldots,6$ are NV$^-$ operators. $\hat {N}$ and $\hat {N}'_{s,i}$ represent noises and are used to preserve the desired commutation relations.

For an NV$^-$-cavity system with resonant frequency $\omega _{c}\approx \omega _{0,3}$ and the NV$^-$ center in the GSM state $|0\rangle$ or $|+1\rangle$, a $\sigma _-$ polarized photon with frequency $\omega \approx \omega _{0,3}$ can be scattered with state-dependent reflection coefficients [7173]

$$\begin{aligned} r_g &=1-\frac{2}{1-i\Delta_c+\frac{2 C_3}{1-i\Delta_3}},\\ r_e &=1-\frac{2}{1-i\Delta_c+\frac{2 C_5}{1-i\Delta_5}+\frac{2 C_6}{1-i\Delta_6}}, \end{aligned}$$
where the detunings are $\Delta _c=(\omega -\omega _{c})/\kappa$, $\Delta _3=(\omega -\omega _{0,3})/\gamma _3$, and $\Delta _{k}=(\omega -\omega _{+1,k})/\gamma _k$ for $k=5$ and $6$. The cooperativity parameters are $C_i=g^2/\gamma _i\kappa$ for $i=3, 5, 6$. For an NV$^-$-cavity system, we also have the desired state-dependent reflection coefficients $r_g\simeq -r_e\simeq 1$, when the parameters satisfy $\Delta _c=\Delta _3=0$, $C_3\gg 1$, and $C_5/\Delta _5\simeq C_6/\Delta _6\simeq 0$.

6. Performance of the heralded parallel protocol

In our heralded parallel protocol, one pair of high-dimensional time-bin entangled photons are utilized and their states are engineered by linear-optical elements and QMs. We have assumed an ideal case where the photon scattering process from the four-level atom-cavity system is deterministic with reflection coefficients $r_g=-r_e=1$. However, the finite cooperativity $C_i$ and normalized detunings including $\Delta _{i}$ and $\Delta _{c}$ in combination with additional excited states deviate the scattering from an ideal one, which decreases the efficiency and fidelity of the entanglement creation. Through proper parameter tuning, a practical NV$^-$-cavity system can actively suppress the detrimental effects of non-ideal scattering conditions on fidelity, although this may come at the cost of reduced efficiency. For instance, when parameters $C_3=110$, $\Delta _{c}=0.5$, $\Delta _{3}=0$, $\Delta _{5}=400$, $\Delta _{6}=550$, $(\gamma _{3}, \gamma _{5}, \gamma _{6})/2\pi =(6, 11, 13)$ MHz [71] are used, the reflection coefficients $r_g=-r_e\simeq 0.999$ can be achieved with a phase difference of $\pi$, leading to a unity fidelity but a decreased efficiency with $\eta _0=|r_{g}|^2$ for each single photon scattering.

The efficiency of our protocol for entangling $n$ pairs of QMs can be described as

$$\eta_n=\eta_{p}^2\eta_0^{2n}\eta_d^{2} \textrm{exp} (-\alpha L).$$
Here the exponential term represents the channel loss, $L$ is the distance between two nodes and $\alpha$ is the channel attenuation coefficient. $\eta _p$ is the PC efficiency which can reach $\eta _p\geq 0.99$ and $\eta _d$ is the efficiency of single-photon detector which can reach $\eta _d\geq 0.96$. Meanwhile, the dark count rate can be on the order of Hz, leading to a dark count probability of $10^{-8}$ per $10$ ns detection window [90]. The influence of the low dark count rate can be neglected in our calculation. For conventional schemes, requiring $n$ pairs of photons to entangle $n$ pairs of QMs, the efficiency is
$$\eta'_n=\eta_0^{2n}\eta_d^{2n}\textrm{exp}({-}n\alpha L).$$

The efficiencies $\eta _2$ and $\eta _3$ of our protocol for entangling $n=2$ and $n=3$ pairs of QMs as a function of the distance $L$ are shown in Fig. 4(a). Here, we assume that the attenuation rate of optical fiber channels is $0.2$ dB/km $(\alpha \simeq$1/22 km$^{-1})$ and single-photon detector efficiency is $\eta _d=0.96$.The corresponding efficiencies $\eta '_2$ and $\eta '_3$ of conventional schemes with the same parameters are also shown in Fig. 4(a). Meanwhile, the efficiency $\eta '_1$ for generating a single pair of entangled QMs is shown using the same parameters as a reference. When the QMs are located in the same node, that is $L\simeq 0$, we can entangle $n=2$ and $n=3$ pairs of QMs with the efficiencies of $\eta _2=0.768$ and $\eta _3=0.709$, which exceed the corresponding efficiencies $\eta '_2=0.723$ and $\eta '_3=0.614$ of the previous schemes. When QM pairs are separated by a medium distance $L=25$ km, we can entangle $n=3$ pairs of QMs with the efficiency $\eta _{3}=0.228$ which is approximately three times higher than that for preparing two pairs of QM entanglement with two photon pairs with $\eta '_2=0.074$, while approximately $83{\%}$ of that for generating a single pair of entangled QMs. Theoretically, our protocol demonstrates a further advantage in efficiency as either the distance between two nodes or the size of the time-bin dimension increases.

 figure: Fig. 4.

Fig. 4. (a) Efficiency of entanglement creation for $n$ pairs of QMs versus distance $L$. (b) Efficiency enhancement versus distance $L$ for spatial multiplexing with $x=5$ entangled photon pairs. Here we set $\eta _0=0.96$, $\eta _d=0.96$, and $\alpha \simeq 1/22$ km$^{-1}$.

Download Full Size | PDF

Spatial multiplexing can further increase the efficiency of entanglement creation by transmitting several photon pairs in parallel [38]. Suppose there are $x$ high-dimensional entangled photon pairs in each transmission round and there are $y$ receiver units in each quantum node [36]. Each receiver unit is composed of photon-state engineering blocks and $m\geq 2$ QMs, shown in Figs. 2 and 3. When a unit $y_i$ receives a photon $x_j$, heralded by a successful measurement of the photon, the photon $x_{j+1}$ is directed to the next unit $y_{i+1}$, otherwise $x_{j+1}$ is directed to the unit $y_{i}$ after it has been re-initialized. Finally, two nodes exchange information about the received photons and project two units receiving photons from the same photon pair into the target entangled states [36].

For simplicity, we consider the case where entanglement has been established between all units of two quantum nodes, corresponding to the situation that $y$ pairs of entangled photons have arrived at their respective nodes, interacted with the receiver units, and clicked $2y$ photon detectors. The efficiencies of our scheme ($P_1$) and conventional schemes ($P_2$) are [36]

$${P_l=\sum_{i=ly}^{lx}p_l^{ly}{p'_l}^{i-ly}C^{ly-1}_{i-1}.}$$

Here $C^a_b=\frac {b!}{a!(b-a)!}$ is the binomial coefficient. $p_1=\eta _2$ ($p_2=\eta '_1$) is the efficiency of the parallel (conventional) scheme without spatial multiplexing, in which a pair of four-dimensional (two-dimensional) entangled photons have been successfully detected by two single-photon detectors. $p'_l=(1-\sqrt {p_l})^2$ for $l=1$ and $2$ is the probability that both photons are lost. Note that the efficiency $P_2$ corresponds to the scenario in which $2y$ pairs of QMs are entangled with $2x$ two-dimensional entangled photon pairs using conventional schemes.

For a specific example of spatial multiplexing, with $x=5$ and $y=2$, the efficiency of entangling at least one pair of receiver units is

$$P'_1=P_1+C_5^1 p_1 {p'_1}^4+4p_1\sqrt{p_1p'_1}\sum_{j=0}^{3}\sum_{i=0}^{j}{p'_1}^{\frac{j+i}{2}}C_{i+1}^{1}f(j),$$
where the auxiliary function $f$ is given by $f(j)=\sqrt {p_1}$ for $j\in \{0,1,2\}$ and $f(3)=1$. In parallel entangling schemes that utilize a four-dimensional single photon, the efficiency of entangling at least one pair of receiver units is enhanced when employing spatial multiplexing with two receiver units in each node, resulting in an efficiency denoted as
$$P_3=2 P_s (1-P_s)+P_s^2,$$
where $P_s=\eta _0^4\eta _d\textrm {exp} (-\alpha L)$ is the corresponding efficiency without spatial multiplexing. Generally, the efficiency of parallel schemes utilizing single photons is always slightly higher than that of our protocol for each transmission that aims to entangle the same number of QM pairs, owing to the finite efficiency $\eta _d$ of single-photon detectors. When considering spatial multiplexing, our protocol can obtain a larger multiplexing capacity for a network with a limited number of QMs at each node and substantial communication delays between two neighboring nodes. For such a situation, our protocol can achieve a higher efficiency than that of parallel schemes utilizing single photons.

To show the enhanced efficiency achieved by spatial multiplexing, we introduce the parameters

$$\begin{aligned} E_1^{2y}&=P_1/\eta_2^y,\\ E_2^{2y}&=P_1/P_2,\\ {E}_{3}^{4}&=P'_1/P_3, \end{aligned}$$
where $\eta _2^y$ is the efficiency of entangling $y$ receiver units using our parallel protocol without spatial multiplexing. As shown in Fig. 4(b), the efficiency of our protocol can be enhanced through spatial multiplexing and surpasses that achieved with conventional schemes. For instance, the enhancement can reach $E_1^{4}=3.4$ and $E_2^{4}=27.9$ for $y=2$, $x=5$, and $L=50$ km. Similarly, when compared to parallel schemes utilizing a four-dimensional single photon and two receiver units, our scheme using four-dimensional entangled photon pairs exhibits a higher efficiency, with ${E}_{3}^{4}=1.7$ for $L=50$ km.

In our protocol, optical switches are used to manipulate the time-bin states of photons. So far, we assume that the OS is ideal with the unity efficiency. In practice, the efficiency $\eta _s$ of the OS is always less than unity due to the intersection loss, and our efficiency should be multiplied by a parameter $\eta ^m_s$, where we assume the efficiency of all OSs are identical and $m$ is the number of OSs used in our parallel protocol. To show the potential enhancement of our parallel protocol, we introduce an efficiency enhancement $\eta ^{\rm enh}_n$ with ideal OSs and a threshold OS efficiency $\eta _n^{\rm thr}$, shown in Fig. 5, as follows:

$$\begin{aligned}&\eta_n^{\rm enh}=\eta_n/\eta'_n,\\ &\eta_n^{\rm thr}=\sqrt[m]{\eta'_n/\eta_n}. \end{aligned}$$

 figure: Fig. 5.

Fig. 5. Efficiency enhancement of entanglement creation and threshold efficiency of optical switches for the parameters assumed in Fig. 4(a).

Download Full Size | PDF

For the specific case with $n=2$ and $n=3$, we have $m=4$ and $m=8$, shown in Fig. 2 and Fig. 3. Interestingly, we have the coincidence that $\eta _2^{\rm thr}\simeq \eta _3^{\rm thr}=\sqrt {\eta _d/\eta _{p}}\exp ({-0.25\alpha L})$, which is independent of the single-photon scattering probability $\eta _0$. Our protocol can tolerate a lower OS efficiency threshold for a longer distance $L$. For instance, we have $\eta _{3}^{\rm thr}=0.739$ for $L=25$ km and $\eta _3^{\rm thr}=0.557$ for $L=50$ km.

To show the robustness against variations in detunings and cooperativities, we simulate the average fidelity of the parallel protocol for entangling two pairs of QMs, which is defined as

$$F=\dfrac{1}{16}\sum_{i\in\{M\}}|\langle\Phi^{(i)}_{\rm de}|\Phi^{(i)}_{\rm un}\rangle|^2,$$
where $\{M\}$ consists of $16$ results of two individual photon measurements, shown in Eq. (5). $|\Phi ^{(i)}_{\rm un}\rangle$ is the state of QMs when generally unbalanced reflections shown in Eq. (11) are adopted and the $i$-th result is obtained, and $|\Phi ^{(i)}_{\rm de}\rangle$ is the corresponding desired state with ideal reflections.

The average fidelity $F$ as a function of $C_3$ is shown in Fig. 6 for five different detunings $\Delta _3$, conditioned on $\Delta _c=0.5$, $\Delta _{5}=400$ and $\Delta _{6}=550$. As anticipated, when $\Delta _3=0$ and $C_3=110$, the average fidelity approaches unity. In addition, for $C_3\geq 80$ and $\Delta _3\in [-10,10]$, the average fidelity retains a large value with $F\geq 0.95$ and the average fidelity of each QM pair is $F_{1}=\sqrt {F}\geq 0.97$. The fidelity $F_{1}$ corresponds to the moment when QMs become entangled after single-photon measurements, and it is independent of the distance $L$ between the two nodes. In contrast, in multiplexing protocols using single photons [42,44], the QMs in the sender node must store the quantum states during photon transmission. The corresponding fidelity $F'_{1}$ of each entangled QM pair, at the moment when the QMs become entangled using ideal spin-photon interfaces, can be given by $F'_{1}=[1+\exp (-L/cT_2)]/2$ [42], where $c$ represents the speed of light in fiber channels and $T_2$ denotes the QM coherence time. Therefore, our protocol can achieve a higher fidelity in entangling QMs for distances $L>10.5$ km, assuming $T_2=1$ ms [42]. Subsequently, a confirmation signal is transmitted across the entire channel for both protocols before further operations can proceed.

 figure: Fig. 6.

Fig. 6. Average fidelity of the protocol for entangling two pairs of QMs in parallel. Here $\Delta _c=0.5$, $\Delta _{5}=400$, $\Delta _{6}=550$, $(\gamma _{3}, \gamma _{5}, \gamma _{6})/2\pi =(6, 11, 13)$ MHz.

Download Full Size | PDF

7. Discussion and summary

Our protocol for entanglement creation between multiple pairs of QMs is achieved by the distribution of a single pair of high-dimensional entangled photons. Two photons experience identical operational procedures in parallel, including the interaction with the QMs, which convert high-dimensional time-bin entanglement into hybrid entanglement between QMs and photons, and then into parallel entanglement between QM pairs. Our protocol thus presents a specific connection between high-dimensional optical time-bin entanglement and two-dimensional stationary entanglement [52]. The QMs are projected into an anticipated product state, comprising one Bell state for each QM pair, which is heralded by the detection of two photons in a proper basis. By the distribution of high-dimensional optical entanglement, the limitation of exponential scaling channel loss on entangling distant QMs is significantly suppressed. This leads to a noteworthy improvement in efficiency when compared to that achieved by entangling each distant QM pair with a single pair of entangled photons [3235]. In fact, the efficiency of our protocol for generating multiple pairs of entangled QMs becomes comparable to that of generating one pair of QMs, when utilizing nearly ideal photon-spin interfaces and optical switches.

The attenuation rate of $0.2$ dB/km in Sec. 6 is suitable for telecom wavelengths. To convert the telecom wavelength to the wavelengths where the NV center operates, a quantum frequency converter should be inserted in each node before interacting the photons with the NV centers. This will decrease the efficiency of our protocol by $\eta _{\rm con}^2$, while the efficiency of a conventional scheme, which involves entangling $n$ QM pairs with $n$ entangled photon pairs, will decrease by $\eta _{\rm con}^{2n}$. Here, $\eta _{\rm con}$ represents the converter efficiency, and a quantum frequency converter with $\eta _{\rm con}=57{\%}$ has been recently used to demonstrate atom-telecom-photon entanglement [91].

The previous parallel protocols [4246] utilize a high-dimensional single photon, sequentially interacting with each QM, whereas our protocol uses an entangled high-dimensional photon pair, interacts each photon with half of all QMs, and works in a parallel style rather than a cascaded one. Considering spatial multiplexing, the previous parallel protocols make efficient use of QMs and promptly announce QM entanglement after photon measurements [38], while our parallel protocol optimizes the utilization of the entangled photon source and QMs experience reduced dephasing during a given number of attempts [37]. Moreover, with a limited number of QMs, the capability of our protocol for spatial multiplexing surpasses the number of QMs within each node [36]. The optimization of entanglement generation rates, achieved by combining high-dimensional entangled photon pairs or high-dimensional single photons with spatial multiplexing, requires further investigation for the construction of large-scale quantum networks in the future. Additionally, our protocol can be extended to simultaneously entangle multiple qubits in parallel across several quantum nodes [9296].

In summary, we have proposed a heralded protocol for entangling multiple pairs of QMs located in remote nodes with one pair of high-dimensional entangled photons. This protocol exploits quantum multiplexing of the time-bin entanglement and can entangle additional pairs of QMs by increasing the dimensions of the time-bin DOF, leading to a substantial enhancement of entanglement efficiency. Our protocol is heralded by the detection of one photon in each node, and thus can be extended to simultaneously support spatial-temporal multiplexing. These distinct features make our protocol particularly useful for long-distance quantum communication and large-scale quantum networks.

Funding

National Natural Science Foundation of China (12004029, 11904171, 62221004); Fundamental Research Funds for the Central Universities (30922010807).

Acknowledgment

The authors would like to thank Prof. Qing Ai and Prof. Zhangqi Yin for helpful discussions.

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. M. Ruf, N. H. Wan, H. Choi, et al., “Quantum networks based on color centers in diamond,” J. Appl. Phys. 130(7), 070901 (2021). [CrossRef]  

2. S. Wehner, D. Elkouss, and R. Hanson, “Quantum internet: A vision for the road ahead,” Science 362(6412), eaam9288 (2018). [CrossRef]  

3. I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simulation,” Rev. Mod. Phys. 86(1), 153–185 (2014). [CrossRef]  

4. F.-G. Deng, G.-L. Long, and X.-S. Liu, “Two-step quantum direct communication protocol using the Einstein-Podolsky-Rosen pair block,” Phys. Rev. A 68(4), 042317 (2003). [CrossRef]  

5. T. Li and G.-L. Long, “Quantum secure direct communication based on single-photon Bell-state measurement,” New J. Phys. 22(6), 063017 (2020). [CrossRef]  

6. J. Wu, G.-L. Long, and M. Hayashi, “Quantum secure direct communication with private dense coding using a general preshared quantum state,” Phys. Rev. Appl. 17(6), 064011 (2022). [CrossRef]  

7. Y.-B. Sheng, L. Zhou, and G.-L. Long, “One-step quantum secure direct communication,” Sci. Bull. 67(4), 367–374 (2022). [CrossRef]  

8. X.-Y. Luo, Y. Yu, J.-L. Liu, et al., “Postselected entanglement between two atomic ensembles separated by 12.5 km,” Phys. Rev. Lett. 129(5), 050503 (2022). [CrossRef]  

9. F. Xu, X. Ma, Q. Zhang, et al., “Secure quantum key distribution with realistic devices,” Rev. Mod. Phys. 92(2), 025002 (2020). [CrossRef]  

10. Z. Gao, T. Li, and Z. Li, “Long-distance measurement-device-independent quantum secure direct communication,” EPL 125(4), 40004 (2019). [CrossRef]  

11. F. Massa, A. Moqanaki, Ä. Baumeler, et al., “Experimental two-way communication with one photon,” Adv. Quantum Technol. 2(11), 1900050 (2019). [CrossRef]  

12. Z. Zhou, Y. Sheng, P. Niu, et al., “Measurement-device-independent quantum secure direct communication,” Sci.China Phys. Mech. Astron. 63(3), 230362 (2020). [CrossRef]  

13. D. Collins and S. Popescu, “Classical analog of entanglement,” Phys. Rev. A 65(3), 032321 (2002). [CrossRef]  

14. W. Qin, A. Miranowicz, H. Jing, et al., “Generating long-lived macroscopically distinct superposition states in atomic ensembles,” Phys. Rev. Lett. 127(9), 093602 (2021). [CrossRef]  

15. P. Kok, W. J. Munro, K. Nemoto, et al., “Linear optical quantum computing with photonic qubits,” Rev. Mod. Phys. 79(1), 135–174 (2007). [CrossRef]  

16. J.-W. Pan, Z.-B. Chen, C.-Y. Lu, et al., “Multiphoton entanglement and interferometry,” Rev. Mod. Phys. 84(2), 777–838 (2012). [CrossRef]  

17. N. Gisin, G. Ribordy, W. Tittel, et al., “Quantum cryptography,” Rev. Mod. Phys. 74(1), 145–195 (2002). [CrossRef]  

18. N. Sangouard, C. Simon, H. de Riedmatten, et al., “Quantum repeaters based on atomic ensembles and linear optics,” Rev. Mod. Phys. 83(1), 33–80 (2011). [CrossRef]  

19. T.-J. Wang, S.-Y. Song, and G.-L. Long, “Quantum repeater based on spatial entanglement of photons and quantum-dot spins in optical microcavities,” Phys. Rev. A 85(6), 062311 (2012). [CrossRef]  

20. Y.-B. Sheng, L. Zhou, and G.-L. Long, “Hybrid entanglement purification for quantum repeaters,” Phys. Rev. A 88(2), 022302 (2013). [CrossRef]  

21. T. Northup and R. Blatt, “Quantum information transfer using photons,” Nat. Photonics 8(5), 356–363 (2014). [CrossRef]  

22. W. J. Munro, K. Azuma, K. Tamaki, et al., “Inside quantum repeaters,” IEEE J. Sel. Top. Quantum Electron. 21(3), 78–90 (2015). [CrossRef]  

23. J. Borregaard, A. S. Sørensen, and P. Lodahl, “Quantum networks with deterministic spin-photon interfaces,” Adv. Quantum Technol. 2(5-6), 1800091 (2019). [CrossRef]  

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

25. C.-S. Yu, X. X. Yi, H.-S. Song, et al., “Robust preparation of Greenberger-Horne-Zeilinger and W states of three distant atoms,” Phys. Rev. A 75(4), 044301 (2007). [CrossRef]  

26. T. Li and F.-G. Deng, “Error-rejecting quantum computing with solid-state spins assisted by low-Q optical microcavities,” Phys. Rev. A 94(6), 062310 (2016). [CrossRef]  

27. D. L. Hurst, K. B. Joanesarson, J. Iles-Smith, et al., “Generating maximal entanglement between spectrally distinct solid-state emitters,” Phys. Rev. Lett. 123(2), 023603 (2019). [CrossRef]  

28. E. Callus and P. Kok, “Cumulative generation of maximal entanglement between spectrally distinct qubits using squeezed light,” Phys. Rev. A 104(5), 052407 (2021). [CrossRef]  

29. E. Togan, Y. Chu, A. S. Trifonov, et al., “Quantum entanglement between an optical photon and a solid-state spin qubit,” Nature 466(7307), 730–734 (2010). [CrossRef]  

30. C. Wang, Y. Zhang, and G.-S. Jin, “Entanglement purification and concentration of electron-spin entangled states using quantum–dot spins in optical microcavities,” Phys. Rev. A 84(3), 032307 (2011). [CrossRef]  

31. T. Li, A. Miranowicz, X. Hu, et al., “Quantum memory and gates using a $\mathrm {\Lambda }$-type quantum emitter coupled to a chiral waveguide,” Phys. Rev. A 97(6), 062318 (2018). [CrossRef]  

32. T. Li, G.-J. Yang, and F.-G. Deng, “Heralded quantum repeater for a quantum communication network based on quantum dots embedded in optical microcavities,” Phys. Rev. A 93(1), 012302 (2016). [CrossRef]  

33. G. Wang and G.-L. Long, “Entanglement purification for memory nodes in a quantum network,” Sci. China Phys. Mech. Astron. 63(2), 220311 (2020). [CrossRef]  

34. X. Liu, J. Hu, Z.-F. Li, et al., “Heralded entanglement distribution between two absorptive quantum memories,” Nature 594(7861), 41–45 (2021). [CrossRef]  

35. D. Lago-Rivera, S. Grandi, J. V. Rakonjac, et al., “Telecom-heralded entanglement between multimode solid-state quantum memories,” Nature 594(7861), 37–40 (2021). [CrossRef]  

36. C. Jones, D. Kim, M. T. Rakher, et al., “Design and analysis of communication protocols for quantum repeater networks,” New J. Phys. 18(8), 083015 (2016). [CrossRef]  

37. P. van Loock, W. Alt, C. Becher, et al., “Extending quantum links: Modules for fiber- and memory-based quantum repeaters,” Adv. Quantum Technol. 3(11), 1900141 (2020). [CrossRef]  

38. W. Munro, K. Harrison, A. Stephens, et al., “From quantum multiplexing to high-performance quantum networking,” Nat. Photonics 4(11), 792–796 (2010). [CrossRef]  

39. S. J. Devitt, W. J. Munro, and K. Nemoto, “Quantum error correction for beginners,” Rep. Prog. Phys. 76(7), 076001 (2013). [CrossRef]  

40. S. Pirandola, J. Eisert, C. Weedbrook, et al., “Advances in quantum teleportation,” Nat. Photonics 9(10), 641–652 (2015). [CrossRef]  

41. F.-G. Deng, B.-C. Ren, and X.-H. Li, “Quantum hyperentanglement and its applications in quantum information processing,” Sci. Bull. 62(1), 46–68 (2017). [CrossRef]  

42. N. Lo Piparo, W. J. Munro, and K. Nemoto, “Quantum multiplexing,” Phys. Rev. A 99(2), 022337 (2019). [CrossRef]  

43. N. Lo Piparo, M. Hanks, C. Gravel, et al., “Resource reduction for distributed quantum information processing using quantum multiplexed photons,” Phys. Rev. Lett. 124(21), 210503 (2020). [CrossRef]  

44. Z. Xie, Y. Liu, X. Mo, et al., “Quantum entanglement creation for distant quantum memories via time-bin multiplexing,” Phys. Rev. A 104(6), 062409 (2021). [CrossRef]  

45. H. Zhou, T. Li, and K. Xia, “Parallel and heralded multiqubit entanglement generation for quantum networks,” Phys. Rev. A 107(2), 022428 (2023). [CrossRef]  

46. Y. Zheng, H. Sharma, and J. Borregaard, “Entanglement distribution with minimal memory requirements using time-bin photonic qudits,” PRX Quantum 3(4), 040319 (2022). [CrossRef]  

47. Y.-B. Sheng and F.-G. Deng, “Deterministic entanglement purification and complete nonlocal Bell-state analysis with hyperentanglement,” Phys. Rev. A 81(3), 032307 (2010). [CrossRef]  

48. P.-S. Yan, L. Zhou, W. Zhong, et al., “Advances in quantum entanglement purification,” Sci. China-Phys. Mech. Astron. 66(5), 250301 (2023). [CrossRef]  

49. M. Erhard, M. Krenn, and A. Zeilinger, “Advances in high-dimensional quantum entanglement,” Nat. Rev. Phys. 2(7), 365–381 (2020). [CrossRef]  

50. W.-Q. Liu, H.-R. Wei, and L.-C. Kwek, “Low-cost Fredkin gate with auxiliary space,” Phys. Rev. Appl. 14(5), 054057 (2020). [CrossRef]  

51. M. Doda, M. Huber, G. Murta, et al., “Quantum key distribution overcoming extreme noise: Simultaneous subspace coding using high-dimensional entanglement,” Phys. Rev. Appl. 15(3), 034003 (2021). [CrossRef]  

52. N. T. Islam, C. C. W. Lim, C. Cahall, et al., “Scalable high-rate, high-dimensional time-bin encoding quantum key distribution,” Quantum Sci. Technol. 4(3), 035008 (2019). [CrossRef]  

53. Z.-Y. Zhang and J.-M. Liu, “Creation of high-dimensional entanglement of polar molecules via optimal control fields,” Phys. Rev. A 105(2), 023113 (2022). [CrossRef]  

54. R. Qu, Y. Wang, M. An, et al., “Retrieving high-dimensional quantum steering from a noisy environment with n measurement settings,” Phys. Rev. Lett. 128(24), 240402 (2022). [CrossRef]  

55. M. Rambach, M. Qaryan, M. Kewming, et al., “Robust and efficient high-dimensional quantum state tomography,” Phys. Rev. Lett. 126(10), 100402 (2021). [CrossRef]  

56. Y. Wang, S. Ru, F. Wang, et al., “Experimental demonstration of efficient high-dimensional quantum gates with orbital angular momentum,” Quantum Sci. Technol. 7(1), 015016 (2022). [CrossRef]  

57. E. Vashukevich, E. Bashmakova, T. Y. Golubeva, et al., “High-fidelity quantum gates for OAM single qudits on quantum memory,” Laser Phys. Lett. 19(2), 025202 (2022). [CrossRef]  

58. W.-T. Fang, Y.-H. Li, Z.-Y. Zhou, et al., “On-chip generation of time-and wavelength-division multiplexed multiple time-bin entanglement,” Opt. Express 26(10), 12912–12921 (2018). [CrossRef]  

59. I. Vagniluca, B. Da Lio, D. Rusca, et al., “Efficient time-bin encoding for practical high-dimensional quantum key distribution,” Phys. Rev. Appl. 14(1), 014051 (2020). [CrossRef]  

60. N. J. Cerf, M. Bourennane, A. Karlsson, et al., “Security of quantum key distribution using $\mathit {d}$-level systems,” Phys. Rev. Lett. 88(12), 127902 (2002). [CrossRef]  

61. B. Da Lio, D. Bacco, D. Cozzolino, et al., “Stable transmission of high-dimensional quantum states over a 2-km multicore fiber,” IEEE J. Sel. Top. Quantum Electron. 26(4), 1–8 (2020). [CrossRef]  

62. A. Mair, A. Vaziri, G. Weihs, et al., “Entanglement of the orbital angular momentum states of photons,” Nature 412(6844), 313–316 (2001). [CrossRef]  

63. C. Wang, F.-G. Deng, Y.-S. Li, et al., “Quantum secure direct communication with high-dimension quantum superdense coding,” Phys. Rev. A 71(4), 044305 (2005). [CrossRef]  

64. C. Xu, S. Huang, Q. Yu, et al., “Manipulating the orbital-angular-momentum correlation of entangled two-photon states in three-dimensional nonlinear photonic crystals,” Phys. Rev. A 104(6), 063716 (2021). [CrossRef]  

65. H. De Riedmatten, I. Marcikic, H. Zbinden, et al., “Creating high dimensional entanglement using mode-locked lasers,” Quantum Inf. Comput. 2(6), 425–433 (2002). [CrossRef]  

66. D. Stucki, H. Zbinden, and N. Gisin, “A Fabry-Perot-like two-photon interferometer for high-dimensional time-bin entanglement,” J. Mod. Opt. 52(18), 2637–2648 (2005). [CrossRef]  

67. W. Qin, A. Miranowicz, P.-B. Li, et al., “Exponentially enhanced light-matter interaction, cooperativities, and steady-state entanglement using parametric amplification,” Phys. Rev. Lett. 120(9), 093601 (2018). [CrossRef]  

68. J. I. Cirac, A. K. Ekert, S. F. Huelga, et al., “Distributed quantum computation over noisy channels,” Phys. Rev. A 59(6), 4249–4254 (1999). [CrossRef]  

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

70. W. Qin, X. Wang, A. Miranowicz, et al., “Heralded quantum controlled-phase gates with dissipative dynamics in macroscopically distant resonators,” Phys. Rev. A 96(1), 012315 (2017). [CrossRef]  

71. K. Nemoto, M. Trupke, S. J. Devitt, et al., “Photonic architecture for scalable quantum information processing in diamond,” Phys. Rev. X 4(3), 031022 (2014). [CrossRef]  

72. N. Lo Piparo, M. Razavi, and W. J. Munro, “Measurement-device-independent quantum key distribution with nitrogen vacancy centers in diamond,” Phys. Rev. A 95(2), 022338 (2017). [CrossRef]  

73. A. Reiserer and G. Rempe, “Cavity-based quantum networks with single atoms and optical photons,” Rev. Mod. Phys. 87(4), 1379–1418 (2015). [CrossRef]  

74. J. Borregaard, H. Pichler, T. Schröder, et al., “One-way quantum repeater based on near-deterministic photon-emitter interfaces,” Phys. Rev. X 10(2), 021071 (2020). [CrossRef]  

75. C. Wang, M. Zhang, X. Chen, et al., “Integrated lithium niobate electro-optic modulators operating at CMOS-compatible voltages,” Nature 562(7725), 101–104 (2018). [CrossRef]  

76. P. Lodahl, “Quantum-dot based photonic quantum networks,” Quantum Sci. Technol. 3(1), 013001 (2018). [CrossRef]  

77. A. J. Heinrich, W. D. Oliver, L. M. Vandersypen, et al., “Quantum-coherent nanoscience,” Nat. Nanotechnol. 16(12), 1318–1329 (2021). [CrossRef]  

78. S.-H. Wei, B. Jing, X.-Y. Zhang, et al., “Towards real-world quantum networks: a review,” Laser Photonics Rev. 16(3), 2100219 (2022). [CrossRef]  

79. G. D. Fuchs, V. V. Dobrovitski, D. M. Toyli, et al., “Gigahertz dynamics of a strongly driven single quantum spin,” Science 326(5959), 1520–1522 (2009). [CrossRef]  

80. F. Dolde, V. Bergholm, Y. Wang, et al., “High-fidelity spin entanglement using optimal control,” Nat. Commun. 5(1), 3371 (2014). [CrossRef]  

81. P. Neumann, J. Beck, M. Steiner, et al., “Single-shot readout of a single nuclear spin,” Science 329(5991), 542–544 (2010). [CrossRef]  

82. L. Robledo, L. Childress, H. Bernien, et al., “High-fidelity projective read-out of a solid-state spin quantum register,” Nature 477(7366), 574–578 (2011). [CrossRef]  

83. H. Bernien, B. Hensen, W. Pfaff, et al., “Heralded entanglement between solid-state qubits separated by three metres,” Nature 497(7447), 86–90 (2013). [CrossRef]  

84. W. Pfaff, B. J. Hensen, H. Bernien, et al., “Unconditional quantum teleportation between distant solid-state quantum bits,” Science 345(6196), 532–535 (2014). [CrossRef]  

85. H.-R. Wei and F.-G. Deng, “Compact quantum gates on electron-spin qubits assisted by diamond nitrogen-vacancy centers inside cavities,” Phys. Rev. A 88(4), 042323 (2013). [CrossRef]  

86. M. Li and M. Zhang, “Robust universal photonic quantum gates operable with imperfect processes involved in diamond nitrogen-vacancy centers inside low-Q single-sided cavities,” Opt. Express 26(25), 33129–33141 (2018). [CrossRef]  

87. M. Pompili, S. L. N. Hermans, S. Baier, et al., “Realization of a multinode quantum network of remote solid-state qubits,” Science 372(6539), 259–264 (2021). [CrossRef]  

88. M. Hanks, M. Trupke, J. Schmiedmayer, et al., “High-fidelity spin measurement on the nitrogen-vacancy center,” New J. Phys. 19(10), 103002 (2017). [CrossRef]  

89. 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]  

90. L. You, “Superconducting nanowire single-photon detectors for quantum information,” Nanophotonics 9(9), 2673–2692 (2020). [CrossRef]  

91. T. van Leent, M. Bock, R. Garthoff, et al., ““Long-distance distribution of atom-photon entanglement at telecom wavelength,” Phys. Rev. Lett. 124(1), 010510 (2020). [CrossRef]  

92. Y. Zhou, B. Li, X.-X. Li, et al., “Preparing multiparticle entangled states of nitrogen-vacancy centers via adiabatic ground-state transitions,” Phys. Rev. A 98(5), 052346 (2018). [CrossRef]  

93. Y.-F. Qiao, H.-Z. Li, X.-L. Dong, et al., “Phononic-waveguide-assisted steady-state entanglement of silicon-vacancy centers,” Phys. Rev. A 101(4), 042313 (2020). [CrossRef]  

94. Y. Zhong, H.-S. Chang, A. Bienfait, et al., “Deterministic multi-qubit entanglement in a quantum network,” Nature 590(7847), 571–575 (2021). [CrossRef]  

95. T.-J. Wang, G.-Q. Yang, and C. Wang, “Control power of high-dimensional controlled teleportation,” Phys. Rev. A 101(1), 012323 (2020). [CrossRef]  

96. P.-J. Stas, Y. Q. Huan, B. Machielse, et al., “Robust multi-qubit quantum network node with integrated error detection,” Science 378(6619), 557–560 (2022). [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 (6)

Fig. 1.
Fig. 1. (a) Level structure and transitions of a four-level atom. (b) Schematic of a controlled-polarization flip unit. CPBS represents a circular-polarization beam splitter that reflects the right-circularly polarized photon $\vert R\rangle$ and transmits left-circularly polarized photon $\vert L\rangle$ .
Fig. 2.
Fig. 2. Schematics of generating two pairs of QMs via distribution of optical entanglement. (a) Step-by-step protocol with (b) circuit of PSE for photon state exchanges. Here QM $k$ ( $k$ =1,…,4) denotes a quantum memory. PC is a Pockel cell that flips the polarization of photons in the last two time-bin modes. OS $_1$ is an optical switch that directs photons in the time-bin states $\vert 0\rangle$ ( $\vert 1\rangle$ ) through the upper (lower) path. OS $_2$ directs two spatial modes of a photon with different time bins into one spatial mode. $\mathcal {T}{(i)}$ introduces an optical delay of $it_\Delta$ . H is a half-wave plate and flips the polarization. PBS is polarization beam splitter that transmits the $D$ -polarized photon and reflects the $A$ -polarized photon, while PBS $'$ transmits the $A$ -polarized photon and reflects the $D$ -polarized photon.
Fig. 3.
Fig. 3. Schematics of generating three pairs of QMs via distribution of optical entanglement. (a) Step-by-step protocol with (b) and (c) circuits of PSE $_1$ and PSE $_2$ for photon state exchanges. Here QM $k$ ( $k$ =1,…,6) denotes a quantum memory. PC $'$ is a Pockel cell that flips the polarization of photons in the last four time-bin modes. OS $_1$ (OS $_3$ ) is an optical switch that directs photons in the time-bin states $\vert 0\rangle$ and $\vert 2\rangle$ ( $\vert 0\rangle$ and $\vert 1\rangle$ ) to the upper path and photons in states $\vert 1\rangle$ and $\vert 3\rangle$ ( $\vert 2\rangle$ and $\vert 3\rangle$ ) to the lower path. OS $_2$ (OS $_4$ ) directs two spatial modes of a photon with different time bins into one spatial mode. H is a half-wave plate and flips the polarization. $\mathcal {T}{(i)}$ describes a time delayer, introducing an optical delay of $it_\Delta$ . PBS is polarization beam splitter that transmits the D-polarized photon and reflects the A-polarized photon. PBS $'$ transmits the A-polarized photon and reflects the D-polarized photon.
Fig. 4.
Fig. 4. (a) Efficiency of entanglement creation for $n$ pairs of QMs versus distance $L$ . (b) Efficiency enhancement versus distance $L$ for spatial multiplexing with $x=5$ entangled photon pairs. Here we set $\eta _0=0.96$ , $\eta _d=0.96$ , and $\alpha \simeq 1/22$ km $^{-1}$ .
Fig. 5.
Fig. 5. Efficiency enhancement of entanglement creation and threshold efficiency of optical switches for the parameters assumed in Fig. 4(a).
Fig. 6.
Fig. 6. Average fidelity of the protocol for entangling two pairs of QMs in parallel. Here $\Delta _c=0.5$ , $\Delta _{5}=400$ , $\Delta _{6}=550$ , $(\gamma _{3}, \gamma _{5}, \gamma _{6})/2\pi =(6, 11, 13)$ MHz.

Equations (19)

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

r k = 1 2 ( 1 i Δ k ) ( 1 i Δ c ) ( 1 i Δ k ) + 2 C k ,
| D | g | D | g , | D | e | A | e , | A | g | A | g , | A | e | D | e ,
| Φ 1 = 1 2 ( ϕ T 0 , 1 | ϕ 12 + ψ T 0 , 1 | ψ 12 ) | ϕ P ,
ϕ T i , j = 1 2 [ T 1 ( i ) T 2 ( i ) + T 1 ( j ) T 2 ( j ) ] , ψ T i , j = 1 2 [ T 1 ( i ) T 2 ( j ) + T 1 ( j ) T 2 ( i ) ] .
| Φ 2 = 1 2 ( ϕ T 0 , 1 | ϕ 12 | ϕ 34 | ϕ P + ψ T 0 , 1 | ψ 12 | ϕ 34 | ϕ P + ϕ T 0 , 1 | ϕ 12 | ψ 34 | ψ P + ψ T 0 , 1 | ψ 12 | ψ 34 | ψ P ) ,
| Ξ 1 = 1 2 [ ( ϕ T 0 , 1 + ϕ T 2 , 3 ) | ϕ 12 + ( ψ T 0 , 1 + ψ T 2 , 3 ) | ψ 12 ] | ϕ P ,
| Ξ 2 = 1 2 2 [ ( ϕ T 0 , 1 + ϕ T 2 , 3 ) | ϕ 12 | ϕ 34 + ( ψ T 0 , 1 + ψ T 2 , 3 ) | ψ 12 | ϕ 34 + ( ψ T 0 , 2 + ψ T 1 , 3 ) | ϕ 12 | ψ 34 + ( ψ T 0 , 3 + ψ T 1 , 2 ) | ψ 12 | ψ 34 ] | ϕ P ,
| Ξ 3 = 1 4 [ ( ϕ T 0 , 1 + ϕ T 2 , 3 ) | ϕ 12 | ϕ 34 + ( ψ T 0 , 1 + ψ T 2 , 3 ) | ψ 12 | ϕ 34 + ( ψ T 0 , 2 + ψ T 1 , 3 ) | ϕ 12 | ψ 34 + ( ψ T 0 , 3 + ψ T 1 , 2 ) | ψ 12 | ψ 34 ] ( | ϕ 56 | ϕ P + | ψ 56 | ψ P ) ,
H e f f = s = 0 , ± 1 i = 1 6 [ ω s | s s | + ( ω i i γ i ) | M i M i | + g s , i ( a ^ σ ^ s , i + a ^ σ ^ s , i ) ] + ( ω c i κ ) a ^ a ^ ,
d a ^ d t = ( i ω c + κ 2 ) a ^ i = 1 6 g s , i σ ^ s , i κ c a ^ i n + N ^ , d σ ^ s , i d t = ( i ω s , i + γ i 2 ) σ ^ s , i g s , i σ ^ s , i z a ^ + N ^ s , i , a ^ o u t = a ^ i n + κ c a ^ ,
r g = 1 2 1 i Δ c + 2 C 3 1 i Δ 3 , r e = 1 2 1 i Δ c + 2 C 5 1 i Δ 5 + 2 C 6 1 i Δ 6 ,
η n = η p 2 η 0 2 n η d 2 exp ( α L ) .
η n = η 0 2 n η d 2 n exp ( n α L ) .
P l = i = l y l x p l l y p l i l y C i 1 l y 1 .
P 1 = P 1 + C 5 1 p 1 p 1 4 + 4 p 1 p 1 p 1 j = 0 3 i = 0 j p 1 j + i 2 C i + 1 1 f ( j ) ,
P 3 = 2 P s ( 1 P s ) + P s 2 ,
E 1 2 y = P 1 / η 2 y , E 2 2 y = P 1 / P 2 , E 3 4 = P 1 / P 3 ,
η n e n h = η n / η n , η n t h r = η n / η n m .
F = 1 16 i { M } | Φ d e ( i ) | Φ u n ( i ) | 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.