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

Nonlocal generalized quantum measurement of product observables with mixed entanglement

Open Access Open Access

Abstract

Nonlocal observables of spacelike separated quantum systems in combination with their measurements contribute greatly to quantum theory and its applications. We present a nonlocal generalized quantum measurement protocol for measuring product observables, assisted by a meter in a mixed entangled state rather than maximally or partially entangled pure states. By tuning the entanglement of the meter, measurement strength of arbitrary values can be achieved for nonlocal product observables, since measurement strength equals the concurrence of the meter. Furthermore, we present a specific scheme to measure the polarization of two nonlocal photons using linear optics. We refer to the polarization and spatial-mode degrees of freedom of the same photon pair as the system and the meter, respectively, which significantly simplifies the interaction between the system and the meter. This protocol can be useful for applications involving nonlocal product observables and nonlocal weak values, and for tests of quantum foundations in nonlocal scenarios.

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

1. Introduction

Quantum measurement plays an important role in quantum theory, and is essential for quantum information technologies. In 1932, von Neumann proposed a strong measurement protocol [1], which describes the interaction of the system and the measurement apparatus (or measurement meter) in a dynamical way. After a strong measurement, the quantum state of the system collapses randomly into an eigenstate of the measurement operator with a corresponding probability. In 1988, Aharonov et al. extended the von Neumann measurement to the weak coupling regime, and presented a weak measurement protocol [2] that reveals some information about the system without collapsing its state into an eigenstate. The weak measurement and weak value have become into a public spotlight [37] due to their various potential applications, such as small signal amplification [812], quantum state and geometric phase determination [1317], and the demonstration of nonclassical features of quantum mechanics [1821].

Nonlocal observables of spacelike separated quantum systems were a contentious topic due to the shortage of direct interaction between them [2224]. In 1931, Landau and Peierls [25] claimed that nonlocal variables cannot be measured instantly without violating relativistic causality. Subsequently, this conjecture was disapproved and nonlocal measurements of some types of observables became possible, given infinite entanglement, under the restriction of relativistic causality [2628].The nonlocal properties of some simple physical systems can, in principle, be measured at defined times without violating relativistic causality [2628]. Furthermore, nonlocal observables can be even measured instantaneously with finite entanglement consumption using verification measurements [2931].

Nonlocal product observables are a specific type of nonlocal observables and are essential for performing many quantum information tasks [3236], ranging from quantum nonlocal tests [37] and semicausal measurements [38] to stabilizer codes [39] and quantum thermal transistors [4042]. Nonlocal von Neumann measurements of these observables usually require a maximally entangled channel [28]. In 2016, Brodutch and Cohen proposed a protocol for generating von Neumann measurement Hamiltonian for nonlocal observables [43] and used it to perform both weak and strong measurements of nonlocal product observables via quantum erasure [4446]. In 2019, Li et al., demonstrated a von Neumann measurement of nonlocal product observables using linear optics and quantum erasure [47]. Meanwhile, strong von Neumann measurement of nonlocal variables [48] and direct measurement of nonlocal wave functions [49] have also been experimentally demonstrated using linear optics. Recently, Vadil and Edamatsu presented a generalized measurement protocol for nonlocal bipartite spin products [50]. They could efficiently perform nonlocal generalized measurements by tuning the entanglement of a meter in a partially entangled pure state to the desired measurement strength [50].

In this article, we propose a generalized protocol for measuring a product observable of a spacelike separated quantum system using a meter with only mixed entanglement. We also provide a specific scheme to perform a nonlocal measurement of a product Pauli observable $\sigma _{z}\otimes \sigma _{z}$ with linear optics. We use the multiple degrees of freedom (DoFs) of single photons [5155], and denote the polarization and spatial modes as the system and auxiliary (meter) qubits, respectively. The inevitable transmission noise can lead to the desired mixed entangled states in the spatial mode [56], while the noise, in contrast, is detrimental for previous nonlocal measurement protocols via quantum erasure [43,47] and for quantum communication protocols [51]. Moreover, our protocol can be used to perform nonlocal measurements with an arbitrary strength by tuning the meter entanglement [56]. Therefore, our protocol can be useful for various applications involving nonlocal product observables and nonlocal weak values, and for tests of quantum foundations in nonlocal scenarios.

2. Protocol for nonlocal generalized measurement of the product observable $\sigma _{\rm z}\otimes \sigma _{\rm z}$ with mixed entanglement

Consider a bipartite system consisting of two spacelike separated qubits, where Alice and Bob each have local access to one qubit. The mixed state of the system is arbitrary and can be described as

$$\rho_S= a_{1}|\Phi^+\rangle_{S}\langle\Phi^+| +a_{2}|\Psi^+\rangle_{S}\langle\Psi^+| +a_{3}|\Psi^-\rangle_{S}\langle\Psi^-| +a_{4}|\Phi^-\rangle_{S}\langle\Phi^-|,$$
where two-qubit Bell states of the system are
$$|{\Phi^{{\pm}}}\rangle_{S}=\frac{1}{\sqrt{2}}(|{00}\rangle_S\pm|{11}\rangle_S), \;\;\;\;\;\; |{\Psi^{{\pm}}}\rangle_{S}=\frac{1}{\sqrt{2}}(|{01}\rangle_S\pm|{10}\rangle_S),$$
and the coefficient $a_{i}$ represents the corresponding probability and is normalized with $\sum _{i=1}^{4}a_{i}=1$.

The meter used to measure the target system is initialized to a mixed entangled state of two orthogonal components and can be described as

$$\rho_{M}=F_1 |{\Phi^{+}}\rangle_{M}\langle{\Phi^{+}}|+F_2|{\Psi^{+}}\rangle_{M}\langle{ \Psi^{+}}|,$$
where the nonnegative coefficients $F_1$ and $F_2$ satisfy the normalized condition $F_1+F_2=1$, $|{\Phi ^{+}}\rangle _{M}$ and $|{\Psi ^{+}}\rangle _{M}$ are the Bell states of the meter with
$$|{\Phi^{{\pm}}}\rangle_{M}=\frac{1}{\sqrt{2}}(|{00}\rangle_M\pm|{11}\rangle_M), \;\;\;\;\;\; |{\Psi^{{\pm}}}\rangle_{M}=\frac{1}{\sqrt{2}}(|{01}\rangle_M\pm|{10}\rangle_M).$$

The coupling to the degenerate subspaces of a product observable can be achieved by controlled-NOT (CNOT) gate operations with the system and meter qubits as the control and target, respectively. After the coupling, a proper measurement on each meter qubit can read out the information about the nonlocal system and complete the nonlocal generalized measurement, shown in Fig. 1. Specifically, the nonlocal generalized measurement process can be described as follows:

 figure: Fig. 1.

Fig. 1. Schematics for nonlocal generalized quantum measurement of the product Pauli observable $\sigma _{\rm zA}\otimes \sigma _{\rm zB}$ using mixed entanglement. Here $\rho _S$ and $\rho _M$ represent the states of the system and the meter, respectively. Alice and Bob perform a controlled-NOT gate on their qubits before measuring their meters in $\sigma _{\rm zA}$ and $\sigma _{\rm zB}$ bases.

Download Full Size | PDF

(1) Alice and Bob perform CNOT gates on their qubits to couple the system to the meter, and evolve the composite system $\rho _S\otimes \rho _M$ into a symmetric state with respect to the meter subsystem, since the parity of the flip-operation number determines the meter state for the specified initial meter state.

(2) Alice and Bob then perform projective measurements on their meter qubits separately in the $\sigma _{\rm z}$ basis. For simplicity, we denote the eigenstates of $\sigma _{\rm z}$ with eigenvalues $+1$ and $-1$ as $|0\rangle$ and $|1\rangle$, respectively. Then, they can get the measurement result $+1$ or $-1$.

(3) Suppose Alice and Bob can share their measurement results. The global outcomes of the generalized quantum measurement of the product Pauli observable $\sigma _{\rm zA}\otimes \sigma _{\rm zB}$ can be calculated from the four different local outcomes, projecting the system into either the even parity or the odd parity subspace with the corresponding global measurement result $+1$ or $-1$, shown in Table 1.

Tables Icon

Table 1. Relationship between local measurement results and global results of the nonlocal product measurement.

If classical communication is not allowed between Alice and Bob, the protocol can success with a probability of $50\%$ by a modification as follows: one of the two parties Alice (Bob) can post-select the local result $+1$ and the other party can then read out the global result by measuring the corresponding local meter.

3. Nonlocal generalized measurement with linear optics and mixed entanglement

Photons interacting weakly with their environments are regarded as the best physical system for long-distance transmission of quantum states. Various nonlocal operations on single photon pairs can be realized at room temperature with only linear optics, such as entanglement purification and concentration [5764], multiple quantum entanglement preparation [6567], and quantum nonlocal measurements [4850,68]. Moreover, photons with multiple DoFs, such as polarization, spatial mode, time-bin, angular momentum, and frequency, enable high-capacity and compact quantum information processing [5155], one-step quantum secure direct communication (QSDC) [6971], and efficient quantum secret sharing [72]. By encoding a photon in both polarization and spatial modes, we present a nonlocal generalized protocol to measure the polarization of two spacelike separated photons, the spatial mode of which performs as a meter in a mixed entangled state. In principle, other DoFs of single photons, such as time-bin and frequency, can also be used to implement a nonlocal generalized measurement of product observables.

The physical implementation of our protocol with linear optics is shown in Fig. 2. A spontaneous parametric down-conversion (SPDC) source, consisting of a type-II $\beta$-barium borate (BBO) crystal and a mirror, generates photon pairs that are entangled in both the spatial mode and the polarization DoFs, which can be referred to as a hyperentangled state [5153], described as follows:

$$|{\Psi}\rangle_{1}=\frac{1}{2}(|a_1b_1\rangle+|a_2b_2\rangle)\otimes(|HH\rangle+|VV\rangle). {}$$

Here $a_{1}$ and $b_{1}$ ($a_{2}$ and $b_{2}$) are the upper (lower) spatial modes of photons $A$ and $B$ sent to Alice and Bob, respectively. $|ij\rangle =|i\rangle _A\otimes |j\rangle _B$ with $i,j\in \{V,H\}$ represents two polarized photons, the first (second) of which is sent to Alice (Bob). For simplicity, we shall use $|ij\rangle$ to describe the polarization here after if there is no confusion.

 figure: Fig. 2.

Fig. 2. Schematics of nonlocal generalized measurement on bipartite spin products with mixed entanglement in a linear optical system. The photon pairs are generated by an SPDC source. The photon polarization is used as the system, and the spatial mode is used as the meter. The beam splitter (BS) implements a Hadamard operation on the spatial mode. Polarization beam splitter (PBS) transmits $H$-polarized photons and reflects $V$-polarized ones, and thus can be referred to as a CNOT gate on the polarization (control) and spatial mode (target) of the same photon. HWP and QWP are, respectively, half-wave plate and quarter-wave plate, which are combined to implement polarization rotations. The upper (lower) modes are denoted as $a_{1}$ and $b_{1}$ ($a_{2}$ and $b_{2}$) for photons sent to Alice and Bob, respectively.

Download Full Size | PDF

The photon $A$ ($B$) is transmitted to Alice (Bob) through a multicore fiber [56]. The channel noise environment will inevitably disturb the photon-pair state. For a multicore fiber channel, the dominant noise for the spatial modes is dephasing, since the crosstalk between the cores is about -$45$ dB$/100$ km [73], and the dominant noise for the polarization is random rotation due to thermal fluctuation, vibration, and birefringence effects of the fiber channel. Therefore, the state of the photon pair after transmission can be evolved into a mixed state [56]

$$\begin{aligned} \rho_{1}&=\rho_{s1}\otimes\rho_{p1}\\ &= (F_{1}|{\Phi^{+}}\rangle_{s}\langle{\Phi^{+}}|+F_{2}|{\Phi^{-}}\rangle_{s}\langle{ \Phi^{-}}|) \otimes (a_{1}|\Phi^+\rangle_p\langle\Phi^+| \\ &+ a_{2}|\Psi^+\rangle_p\langle\Psi^+| +a_{3}|\Psi^-\rangle_p\langle\Psi^-| +a_{4}|\Phi^-\rangle_p\langle\Phi^-|). \end{aligned}$$

The spatial mode of the photon pair is in a mixed entangled state $\rho _{s1}$ composed of $|{\Phi ^{+}}\rangle _{s}$ and $|{\Phi ^{-}}\rangle _{s}$. The spatial-mode entangled states of a photon pair can be described as

$$|{\Phi^{{\pm}}}\rangle_{s}=\frac{1}{\sqrt{2}}(|{a_1b_1}\rangle_M\pm|{a_2b_2}\rangle_M), \;\;\;\;\;\; |{\Psi^{{\pm}}}\rangle_{s}=\frac{1}{\sqrt{2}}(|{a_1b_2}\rangle_M\pm|{a_2b_1}\rangle_M).$$

Here the parameter $F_{1}$ ($F_2$) represents the probability that the spatial mode of the photon pair is in state $|{\Phi ^{+}} \rangle _{s}$ ($|{\Phi ^{-}}\rangle _{s}$). Therefore, $F_1$ can also be referred to as the spatial mode fidelity [74] of the transmission with respect to the initial state $|{\Phi ^{+}} \rangle _{s}$, and $F_2$ can be referred to as the probability with a phase-flip error. The polarization of the photon pair evolves into a mixed state $\rho _{p1}$, in which the channel noise introduces bit flip, phase flip, and phase-bit flip for the polarization DoF with probabilities $a_2$, $a_3$, and $a_4$, respectively. The polarization Bell states can be described as

$$|{\Phi^{{\pm}}}\rangle_{p}=\frac{1}{\sqrt{2}}(|{HH}\rangle_S\pm|{VV}\rangle_S), \;\;\;\;\;\; |{\Psi^{{\pm}}}\rangle_{p}=\frac{1}{\sqrt{2}}(|{HV}\rangle_S\pm|{VH}\rangle_S),$$

The mixed spatial-mode entanglement can not be directly used to perform nonlocal measurement on the product operator $\sigma ^p_{\rm zA}\otimes \sigma ^p_{\rm zB}$ with $\sigma ^p_{\rm zA}=|H\rangle _A\langle H|-|V\rangle _A\langle V|$ and $\sigma ^p_{\rm zB}=|H\rangle _B\langle H|-|V\rangle _B\langle V|$, since the polarization parities of $|{\Phi ^{+}}\rangle _{s}$ and $|{\Phi ^{-}}\rangle _{s}$ are both even and can not be distinguished by local measurements in the $\sigma ^s_{\rm zA}\otimes \sigma ^s_{\rm zB}$ basis with $\sigma ^s_{\rm zA}=|a_1\rangle \langle a_1|-|a_2\rangle \langle a_2|$ and $\sigma ^s_{\rm zB}=|b_1\rangle \langle b_1|-|b_2\rangle \langle b_2|$. Fortunately, a pair of polarization-independent beam splitters (BSs) interfering the lower and upper modes of each photon can transform the phase-flip state into bit-flip state and complete the following transformations of the spatial modes: $|{\Phi ^{+}}\rangle _{s}\stackrel {\rm BS}\longrightarrow |{\Phi ^{+}}\rangle _{s}$, $|{\Phi ^{-}}\rangle _{s} \stackrel {\rm BS}\longrightarrow |{\Psi ^{+}}\rangle _{s}$, when the BSs apply a Hadamard operation on the spatial mode DoF with the transformation matrix $\frac {1}{\sqrt {2}}\!\left ( \begin {smallmatrix} 1 & 1 \\ 1 & -1 \end {smallmatrix} \right )$. Subsequently, a set of half-wave plate (HWP) and quarter-wave plate (QWP) placed at each output port of two BS can evolve the photon pair polarization into another state

$$\rho_{p2}= c_{1}|\Phi^+\rangle_p\langle\Phi^+| +c_{2}|\Psi^+\rangle_p\langle\Psi^+| +c_{3}|\Psi^-\rangle_p\langle\Psi^-| +c_{4}|\Phi^-\rangle_p\langle\Phi^-|,$$
where the coefficients $c_i$ satisfy the normalization condition $\sum _{i=1}^{4}c_{i}=1$.

So far, the state of the polarization (system) and the spatial mode (meter) of two spacelike separated photons has been generated, and it can be described as

$$\rho_{2}=\rho_{s2}\otimes\rho_{p2},$$
where $\rho _{s2}$ represents the spatial mode state with $\rho _{s2}= F_{1}|{\Phi ^{+}}\rangle _{s}\langle {\Phi ^{+}}|+F_{2}|{\Psi ^{+}}\rangle _{s}\langle { \Psi ^{+}}|$. In fact, additional bit-flip noise can be added with any proportion to the spatial modes with a proper loading setup [56], which can be used to change $F_1$ to a desired value for performing nonlocal generalized measurement with different strengths. The photons then impinge on two polarizing beam splitters (PBSs) that transmit H-polarized photons and reflect V-polarized photons. Each PBS completes a CNOT gate on each photon in a deterministic way with the polarization and spatial-mode qubits as the control and target, respectively. The composite state of two photons evolves into
$$\rho_{3}= F_1\rho_{sp1}+F_2\rho_{sp2},$$
where
$$\begin{aligned}\rho_{sp1}&= (c_{1}|\Phi^+\rangle_p\langle\Phi^+| +c_{4}|\Phi^-\rangle_p\langle\Phi^-|)\otimes|\Phi^+\rangle_{s}\langle{\Phi^+}|\\ &+(c_{2}|\Psi^+\rangle_p\langle\Psi^+| +c_{3}|\Psi^-\rangle_p\langle\Psi^-|)\otimes|\Psi^+\rangle_{s}\langle{\Psi^+}|,\\ \rho_{sp2}&= (c_{1}|\Phi^+\rangle_p\langle\Phi^+| +c_{4}|\Phi^-\rangle_p\langle\Phi^-|)\otimes|\Psi^+\rangle_{s}\langle{\Psi^+}|\\ &+(c_{2}|\Psi^+\rangle_p\langle\Psi^+| +c_{3}|\Psi^-\rangle_p\langle\Psi^-|)\otimes|\Phi^+\rangle_{s}\langle{\Phi^+}|. \end{aligned}$$

We can complete the nonlocal generalized measurement of the product Pauli observable $\sigma ^p_{\rm zA}\otimes \sigma ^p_{\rm zB}$ by performing project measurement on the spatial modes of photons A and B, which can be achieved by either nondestructive measurements or post-selecting measurements. For a proof-of-principle demonstration, post-selecting measurements can be used, in a manner similar to what has been suggested for quantum entanglement distillation [59]. Finally, a full tomography of the two-photon polarization state can be achieved by a combination of a QWP, a HWP, a PBS and two single-photon detectors in each spatial mode [47].

The results of the nonlocal generalized measurement are shown in Table 2. The local result of the projective measurement of spatial modes for Alice and Bob, heralded by the click of one single-photon detector $D_a$ ($a=1,2,3,4$) and by the click of one detector $D_b$ ($b=5,6,7,8$), can be either the upper mode denoted by $+1$ or the lower mode denoted by $-1$. The global result of nonlocal generalized measurement of the product observable $\sigma ^p_{\rm zA}\otimes \sigma ^p_{\rm zB}$ will be $+1$ when two local measurement results of the meter qubits are identical, otherwise the global result will be $-1$.

Tables Icon

Table 2. Relationship between local measurement results and the global results of the nonlocal generalized product measurement of $\sigma ^p_{\rm zA}\otimes \sigma ^p_{\rm zB}$ on two polarized photons.

After the projective measurement of spatial modes (meter qubits), the final state of the polarization (system qubits) can be described by applying the following four measurement operators on the polarization DoF of the photon pair:

$$\begin{aligned}M_{+{+}}&=\sqrt{\frac{F_{1}}{2}}\left(\Pi_{HH}+\Pi_{VV}\right)+\sqrt{\frac{F_{2}}{2}}\left(\Pi_{HV}+\Pi_{VH}\right),\\ M_{+{-}}&=\sqrt{\frac{F_{2}}{2}}\left(\Pi_{HH}+\Pi_{VV}\right)+\sqrt{\frac{F_{1}}{2}}\left(\Pi_{HV}+\Pi_{VH}\right),\\ M_{-{+}}&=\sqrt{\frac{F_{2}}{2}}\left(\Pi_{HH}+\Pi_{VV}\right)+\sqrt{\frac{F_{1}}{2}}\left(\Pi_{HV}+\Pi_{VH}\right),\\ M_{-{-}}&=\sqrt{\frac{F_{1}}{2}}\left(\Pi_{HH}+\Pi_{VV}\right)+\sqrt{\frac{F_{2}}{2}}\left(\Pi_{HV}+\Pi_{VH}\right), \end{aligned}$$
where the subscript $+$ ($-$) represents that the local measurement result of Alice and Bob is $+1$ ($-1$). The operator $\Pi _{ij}$ with $i,j\in \{H,V\}$ applied on the polarization represents a projector onto state $|ij\rangle _p$. The nonlocal Pauli product $\sigma ^p_{\rm zA}\otimes \sigma ^p_{\rm zB}$ is thus a degenerate observable [43], since we have $M_{++}= M_{--}$ and $M_{+-}=M_{-+}$. We can define effective measurement operators according to the global results as
$$\begin{aligned} M_{+}&=\sqrt{F_{1}}\left(\Pi_{HH}+\Pi_{VV}\right)+\sqrt{F_{2}}\left(\Pi_{HV}+\Pi_{VH}\right),\\ M_{-}&=\sqrt{F_{2}}\left(\Pi_{HH}+\Pi_{VV}\right)+\sqrt{F_{1}}\left(\Pi_{HV}+\Pi_{VH}\right). \end{aligned}$$

These measurement operators can be rewritten in a more compact form by applying superoperators, $I_{\pm }(\rho _p)=M_{\pm }\rho _pM^{\dagger }_{\pm }$, on the system. The positive operator-valued measure (POVM) effects for the whole process, corresponding to the projective measurement on meter qubits, can be described as

$$E_{{\pm}}=\frac{1}{2}[\sigma_I\pm(F_{1}-F_{2})\sigma_{\rm zA}\sigma_{\rm zB}],$$
where $\sigma _I=(|H\rangle _A\langle H|+|V\rangle _A\langle V|)\otimes (|H\rangle _B\langle H|+|V\rangle _B\langle V|)$ is the identity operator of the polarization of the photon pair. The statistics associated to this POVM reveal that the measurement strength of the nonlocal generalized measurement is
$$S=F_{1}-F_{2}, \;\;\;\;\; F_{1}\geq F_{2},$$
which can be tuned actively by a proper noise loading setup [56]. For $S=1$ the meter is maximally entangled, corresponding to a strong measurement, while for $S=0$ the meter is in the maximally mixed state, corresponding to no measurement.

The corresponding state of the system conditioned on a global measurement result $i\in \{+,-\}$ can be described as

$$\rho_p^f=\dfrac{M_{i}\rho_3M^{{\dagger}}_{i}}{{\rm Tr}(M^{{\dagger}}_{i}M_{i}\rho_3)} ,$$
where Tr represents a trace operation over the spatial-mode subspace. Therefore, the nonlocal generalized measurement of the Pauli observable $\sigma ^p_{\rm zA}\otimes \sigma ^p_{\rm zB}$ will not change the entanglement of the system if it is initialized to a subspace consisting of an even (odd) polarization parity, which features the distinct property of nonlocal measurement [43,47,50].

4. Measurement strength and concurrence

The measurement strength is closely related to the entanglement of the meter and thus the probability of state $|\Phi ^+\rangle _s$ in the mixed meter state. The entanglement of a mixed state of a bipartite system can be described by the entanglement of formation [75], an explicit formula of which can be described as [76,77]

$$\begin{aligned}E\left(\rho\right)&={\mathcal{E}}\left[C \left(\rho\right)\right],\\ {\mathcal{E}}(C)&=h\left(\dfrac{1}{1-C^2}\right), \end{aligned}$$
where $h(x)=-x{\rm log}_2x-(1-x){\rm log}_2(1-x)$, $\rho$ is the density matrix of a bipartite system, $C\left (\rho \right )$ is referred to as the concurrence.

In practice, ${\mathcal {E}}(C)$ monotonically increases with the concurrence $C$, and we can take the concurrence as a measure of entanglement with

$$C\left(\rho\right)=\max\left\{ 0,\lambda_{1}-\lambda_{2}-\lambda_{3}-\lambda_{4}\right\}.$$

Here $\lambda _{i}$ with ${i}\in \left \{1,2,3,4\right \}$ are eigenvalues of the Hermitian matrix $R=\sqrt {\sqrt {\rho }\widetilde {\rho }\sqrt {\rho }}$ and thus can also be referred to as the square roots of the eigenvalues of the matrix $\rho \widetilde {\rho }$. The $\lambda _{i}$ is in a descending order with $\lambda _{1}\geq \lambda _{2}\geq \lambda _{3}\geq \lambda _{4}$. The spin-flip state $\widetilde {\rho }$ of the bipartite system is [77]

$$\widetilde{\rho}=\left(\sigma_{\rm y}\otimes\sigma_{\rm y}\right)\rho^{*}\left(\sigma_{\rm y}\otimes\sigma_{\rm y}\right),$$
where $\sigma _{\rm y}= \left ( \begin {smallmatrix} 0 & -i \\ i & 0 \\ \end {smallmatrix} \right )$ is the Pauli operator in the standard basis and $\rho ^{*}$ is the conjugation of density matrix $\rho$. Therefore, we have $0\leq C\left (\rho \right )\leq 1$. For $C=0$, the bipartite system is separable without any entanglement, while for $C=1$, the bipartite system is maximally entangled.

In our nonlocal generalized measurement with linear optics, the meter qubits are encoded by the spatial modes of a photon pair, and their state just before the measurement is $\rho _{s2}=F_{1}|{\Phi ^{+}}\rangle _{s}\langle {\Phi ^{+}}|+F_{2}|{\Psi ^{+}} \rangle _{s}\langle {\Psi ^{+}}|$, which is a conjugate matrix with $\rho _{s2}=\rho _{s2}^{*}$. The spin-flip state of the meter, according to Eq. (20), can be described as

$$\widetilde{\rho}_{s2}=\frac{1}{2}\left(\begin{array}{cccc} F_{1} & 0 & 0 & F_{1}\\ 0 & F_{2} & F_{2} & 0\\ 0 & F_{2} & F_{2} & 0\\ F_{1} & 0 & 0 & F_{1} \end{array}\right).$$

The corresponding matrix $N$ for calculating the concurrence of the meter state can be described as follows:

$$N=\rho_{s2}\widetilde{\rho}_{s2}=\frac{1}{2}\left(\begin{array}{cccc} F_{1}^{2} & 0 & 0 & F_{1}^{2}\\ 0 & F_{2}^{2} & F_{2}^{2} & 0\\ 0 & F_{2}^{2} & F_{2}^{2} & 0\\ F_{1}^{2} & 0 & 0 & F_{1}^{2} \end{array}\right).$$

The four eigenvalues of matrix $N$ in the descending order for $F_1\geq F_2$ are $\lambda _{1}=F_{1}^{2}$, $\lambda _{2}= F_{2}^{2}$, $\lambda _{3}=0$, and $\lambda _{4}= 0$, respectively. Therefore, the concurrence of the meter state is

$$C_{1}=F_{1}-F_{2}, \;\;\;\;\; F_{1} \geq F_{2},$$
which exactly equals the measurement strength $S$ shown in Eq. (16).

For different measurement strengths, the mixed entanglement of the spatial mode required for measuring the polarization of two spacelike separated photons should be changed accordingly. For $C_1$ approaches unity, a strong nonlocal measurement can be achieved, while for $C_1$ approaches zero, only a weak measurement can be achieved. Furthermore, for a general $C_1$, a nonlocal measurement at any intermediate strength $S=C_1$ can be achieved with the protocol presented in Fig. 1. Generally, the concurrence $C_1$ of the meter is assumed to be no less than the measurement strength, otherwise entanglement purification will be required to increase the concurrence of the meter to the desired level [78] . For the specific measurement protocol implemented with linear optics, we use the polarization and the spatial modes of a photon pair as the system and the meter, respectively. The distribution of spatial mode entanglement can be faithful, and efficient distribution of high-dimensional spatial mode entanglement through a $11$ km multicore fiber has been experimentally demonstrated with an average fidelity of $F=0.921$ [73]. Therefore, the linear protocol can be implemented directly for nonlocal measurements with a weak and/or mediate measurement strength. Note that the aforementioned results can also be obtained for other systems, when the meter state is initialized to a similar mixed state and the corresponding two-qubit CNOT gate is available [7981].

5. Discussion and summary

We show that a generalized quantum measurement of a product observable can be implemented with mixed entanglement, which in practice can be more resource-efficient compared with those based on quantum erasure [43] or partially entangled pure states [50]. We can measure a product observable $\sigma _{\rm z}\otimes \sigma _{\rm z}$ on an arbitrary state of a spacelike separated two-qubit system, when the meter is prepared in a mixed entangled state composed of $|\Phi ^+\rangle _{M}$ and $|\Psi ^+\rangle _{M}$. This protocol is thus particularly useful for the case with a bit-flip quantum channel, a phase-flip quantum channel or a bit-phase flip channel [82], which evolves a maximally entangled state into a mixed one consisting of two components $|\Phi ^+\rangle _{M}$ and $|\Psi ^+\rangle _{M}$, up to single-qubit rotations of meter qubits. For a bit-phase flip channel, the single-qubit rotations can be implemented by $\sigma ^M_{\rm A}\otimes \sigma ^M_{\rm B}$ with $\sigma ^M_{\rm A}=|0\rangle \langle 0|-i|1\rangle \langle 1|$ and $\sigma ^M_{\rm B}=|0\rangle \langle 0|+i|1\rangle \langle 1|$ rotations of meter qubits A and B, respectively.

The mixed entangled meter in principle can be referred to as a binary with even and odd parities, and thus leads to distinct results after its interaction with the system, which is the essential for our nonlocal generalized measurement using mixed entanglement. We can perform the nonlocal generalized measurement directly when the required measurement strength equals the concurrence of the meter, rather than performing entanglement purification [5763] to prepare maximally entangled states for the measurement with quantum erasure. Furthermore, this protocol can be generalized to measure other product observables [43,50] by introducing additional single-qubit rotations. For instance, to measure the nonlocal product observable $\sigma _{\rm x}\otimes \sigma _{\rm x}$, a Hadamard operation should be applied on each system qubit before and after the nonlocal generalized quantum measurement described in Fig. 1.

In summary, we have shown that it is possible to perform a nonlocal generalized quantum measurement of a product observable with mixed entanglement. The measurement strength equals the concurrence of the meter, and thus an arbitrary-strength nonlocal measurement can be achieved by properly modifying the entanglement of the meter. Furthermore, we present a feasible implementation protocol with linear optics. The post-measurement states and statistics after the nonlocal measurement are the same as those based on quantum erasure with maximal entanglement. All these distinguishing features make our protocol efficient for various applications involving nonlocal product observables in combination with tests of quantum foundations in nonlocal scenarios.

Funding

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

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. J. von Neumann, “Mathematische grundlagen der quantenmechanik,” Am. Math. Mon. 42(4), 237–239 (1935). [CrossRef]  

2. Y. Aharonov, D. Z. Albert, and L. Vaidman, “How the result of a measurement of a component of the spin of a spin-1/2 particle can turn out to be 100,” Phys. Rev. Lett. 60(14), 1351–1354 (1988). [CrossRef]  

3. A. G. Kofman, S. Ashhab, and F. Nori, “Nonperturbative theory of weak pre-and post-selected measurements,” Phys. Rep. 520(2), 43–133 (2012). [CrossRef]  

4. J. Dressel, M. Malik, F. M. Miatto, A. N. Jordan, and R. W. Boyd, “Colloquium: understanding quantum weak values: basics and applications,” Rev. Mod. Phys. 86(1), 307–316 (2014). [CrossRef]  

5. J. Dressel, K. Y. Bliokh, and F. Nori, “Classical field approach to quantum weak measurements,” Phys. Rev. Lett. 112(11), 110407 (2014). [CrossRef]  

6. J. T. Monroe, N. Yunger Halpern, T. Lee, and K. W. Murch, “Weak measurement of a superconducting qubit reconciles incompatible operators,” Phys. Rev. Lett. 126(10), 100403 (2021). [CrossRef]  

7. F. De Zela, “Role of weak values in strong measurements,” Phys. Rev. A 105(4), 042202 (2022). [CrossRef]  

8. O. Hosten and P. Kwiat, “Observation of the spin hall effect of light via weak measurements,” Science 319(5864), 787–790 (2008). [CrossRef]  

9. X. Zhou, Z. Xiao, H. Luo, and S. Wen, “Experimental observation of the spin Hall effect of light on a nanometal film via weak measurements,” Phys. Rev. A 85(4), 043809 (2012). [CrossRef]  

10. L. Zhou, Y. Turek, C. P. Sun, and F. Nori, “Weak-value amplification of light deflection by a dark atomic ensemble,” Phys. Rev. A 88(5), 053815 (2013). [CrossRef]  

11. S. Lin, J. Hong, Z. Chen, Y. Chen, and X. Zhou, “Optimal weak measurement in the photonic spin Hall effect for arbitrary linear polarization incidence,” Opt. Express 30(3), 4096–4105 (2022). [CrossRef]  

12. Y. Wang, W. Zhang, S. Chen, S. Wen, and H. Luo, “Multiple-weak-value quantum measurement for precision estimation of time delay,” Phys. Rev. A 105(3), 033521 (2022). [CrossRef]  

13. J. S. Lundeen and C. Bamber, “Procedure for direct measurement of general quantum states using weak measurement,” Phys. Rev. Lett. 108(7), 070402 (2012). [CrossRef]  

14. E. Sjöqvist, “Geometric phase in weak measurements,” Phys. Lett. A 359(3), 187–189 (2006). [CrossRef]  

15. X.-Y. Xu, Y. Kedem, K. Sun, L. Vaidman, C.-F. Li, and G.-C. Guo, “Phase estimation with weak measurement using a white light source,” Phys. Rev. Lett. 111(3), 033604 (2013). [CrossRef]  

16. M. Cormann, M. Remy, B. Kolaric, and Y. Caudano, “Revealing geometric phases in modular and weak values with a quantum eraser,” Phys. Rev. A 93(4), 042124 (2016). [CrossRef]  

17. Q. Hu, T. Yusufu, and Y. Turek, “Quantum state engineering using weak measurements,” Phys. Rev. A 105(2), 022608 (2022). [CrossRef]  

18. Y. Aharonov, A. Botero, S. Popescu, B. Reznik, and J. Tollaksen, “Revisiting Hardy’s paradox: counterfactual statements, real measurements, entanglement and weak values,” Phys. Lett. A 301(3-4), 130–138 (2002). [CrossRef]  

19. N. S. Williams and A. N. Jordan, “Weak values and the Leggett-Garg inequality in solid-state qubits,” Phys. Rev. Lett. 100(2), 026804 (2008). [CrossRef]  

20. C. Emary, N. Lambert, and F. Nori, “Leggett-Garg inequalities,” Rep. Prog. Phys. 77(1), 016001 (2014). [CrossRef]  

21. M. Huang, R.-K. Lee, L. Zhang, S.-M. Fei, and J. Wu, “Simulating broken $\mathcal {PT}$-symmetric Hamilton systems by weak measurement,” Phys. Rev. Lett. 123(8), 080404 (2019). [CrossRef]  

22. 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]  

23. J. S. Bell, “On the Einstein Podolsky Rosen paradox,” Physics Physique Fizika 1(3), 195–200 (1964). [CrossRef]  

24. F. Laloë, Do we really understand quantum mechanics? (Cambridge University, 2019).

25. L. Landau and R. Peierls, “Erweiterung des unbestimmtheitsprinzips für die relativistische quantentheorie,” Z. Physik 69(1-2), 56–69 (1931). [CrossRef]  

26. Y. Aharonov and D. Z. Albert, “Can we make sense out of the measurement process in relativistic quantum mechanics?” Phys. Rev. D 24(2), 359–370 (1981). [CrossRef]  

27. Y. Aharonov, D. Z. Albert, and L. Vaidman, “Measurement process in relativistic quantum theory,” Phys. Rev. D 34(6), 1805–1813 (1986). [CrossRef]  

28. S. Popescu and L. Vaidman, “Causality constraints on nonlocal quantum measurements,” Phys. Rev. A 49(6), 4331–4338 (1994). [CrossRef]  

29. L. Vaidman, “Instantaneous measurement of nonlocal variables,” Phys. Rev. Lett. 90(1), 010402 (2003). [CrossRef]  

30. B. Groisman and B. Reznik, “Measurements of semilocal and nonmaximally entangled states,” Phys. Rev. A 66(2), 022110 (2002). [CrossRef]  

31. S. R. Clark, A. J. Connor, D. Jaksch, and S. Popescu, “Entanglement consumption of instantaneous nonlocal quantum measurements,” New J.Phys. 12(8), 083034 (2010). [CrossRef]  

32. C. H. Bennett, “Quantum information,” Phys. Scr. T76(1), 210 (1998). [CrossRef]  

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

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

35. W. Qin, A. Miranowicz, P.-B. Li, X.-Y. Lü, J. Q. You, and F. Nori, “Exponentially enhanced light-matter interaction, cooperativities, and steady-state entanglement using parametric amplification,” Phys. Rev. Lett. 120(9), 093601 (2018). [CrossRef]  

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

37. J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt, “Proposed experiment to test local hidden-variable theories,” Phys. Rev. Lett. 23(15), 880–884 (1969). [CrossRef]  

38. D. Beckman, D. Gottesman, M. A. Nielsen, and J. Preskill, “Causal and localizable quantum operations,” Phys. Rev. A 64(5), 052309 (2001). [CrossRef]  

39. D. Gottesman, Stabilizer codes and quantum error correction (California Institute of Technology, 1997).

40. K. Joulain, J. Drevillon, Y. Ezzahri, and J. Ordonez-Miranda, “Quantum thermal transistor,” Phys. Rev. Lett. 116(20), 200601 (2016). [CrossRef]  

41. B.-Q. Guo, T. Liu, and C.-S. Yu, “Quantum thermal transistor based on qubit-qutrit coupling,” Phys. Rev. E 98(2), 022118 (2018). [CrossRef]  

42. Y. Zhang, Z. Yang, X. Zhang, B. Lin, G. Lin, and J. Chen, “Coulomb-coupled quantum-dot thermal transistors,” EPL 122(1), 17002 (2018). [CrossRef]  

43. A. Brodutch and E. Cohen, “Nonlocal measurements via quantum erasure,” Phys. Rev. Lett. 116(7), 070404 (2016). [CrossRef]  

44. M. O. Scully and K. Drühl, “Quantum eraser: a proposed photon correlation experiment concerning observation and "delayed choice" in quantum mechanics,” Phys. Rev. A 25(4), 2208–2213 (1982). [CrossRef]  

45. T. J. Herzog, P. G. Kwiat, H. Weinfurter, and A. Zeilinger, “Complementarity and the quantum eraser,” Phys. Rev. Lett. 75(17), 3034–3037 (1995). [CrossRef]  

46. A. Peruzzo, P. Shadbolt, N. Brunner, S. Popescu, and J. L. O’Brien, “A quantum delayed-choice experiment,” Science 338(6107), 634–637 (2012). [CrossRef]  

47. Y. Li, H.-S. Zhong, Y.-H. Luo, L.-C. Peng, C.-Y. Lu, N.-L. Liu, J. Zhang, L. Li, and J.-W. Pan, “Experimental nonlocal measurement of a product observable,” Optica 6(9), 1199–1202 (2019). [CrossRef]  

48. X.-Y. Xu, W.-W. Pan, Q.-Q. Wang, J. Dziewior, L. Knips, Y. Kedem, K. Sun, J.-S. Xu, Y.-J. Han, C.-F. Li, G.-C. Guo, and L. Vaidman, “Measurements of nonlocal variables and demonstration of the failure of the product rule for a pre- and postselected pair of photons,” Phys. Rev. Lett. 122(10), 100405 (2019). [CrossRef]  

49. W.-W. Pan, X.-Y. Xu, Y. Kedem, Q.-Q. Wang, Z. Chen, M. Jan, K. Sun, J.-S. Xu, Y.-J. Han, C.-F. Li, and G.-C. Guo, “Direct measurement of a nonlocal entangled quantum state,” Phys. Rev. Lett. 123(15), 150402 (2019). [CrossRef]  

50. P. Vidil and K. Edamatsu, “Nonlocal generalized quantum measurements of bipartite spin products without maximal entanglement,” New J. Phys. 23(4), 043004 (2021). [CrossRef]  

51. 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]  

52. Y.-B. Sheng, F.-G. Deng, and G. L. Long, “Complete hyperentangled-Bell-state analysis for quantum communication,” Phys. Rev. A 82(3), 032318 (2010). [CrossRef]  

53. 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]  

54. J.-Z. Liu, N.-Y. Chen, W.-Q. Liu, H.-R. Wei, and M. Hua, “Hyperparallel transistor, router and dynamic random access memory with unity fidelities,” Opt. Express 27(15), 21380–21394 (2019). [CrossRef]  

55. 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]  

56. X.-M. Hu, C.-X. Huang, Y.-B. Sheng, L. Zhou, B.-H. Liu, Y. Guo, C. Zhang, W.-B. Xing, Y.-F. Huang, C.-F. Li, and G.-C. Guo, “Long-distance entanglement purification for quantum communication,” Phys. Rev. Lett. 126(1), 010503 (2021). [CrossRef]  

57. J.-W. Pan, C. Simon, Č. Brukner, and A. Zeilinger, “Entanglement purification for quantum communication,” Nature 410(6832), 1067–1070 (2001). [CrossRef]  

58. T. Yamamoto, M. Koashi, Ş. K. Özdemir, and N. Imoto, “Experimental extraction of an entangled photon pair from two identically decohered pairs,” Nature 421(6921), 343–346 (2003). [CrossRef]  

59. Y.-B. Sheng and F.-G. Deng, “One-step deterministic polarization-entanglement purification using spatial entanglement,” Phys. Rev. A 82(4), 044305 (2010). [CrossRef]  

60. B.-C. Ren, F.-F. Du, and F.-G. Deng, “Hyperentanglement concentration for two-photon four-qubit systems with linear optics,” Phys. Rev. A 88(1), 012302 (2013). [CrossRef]  

61. T.-J. Wang, L.-L. Liu, R. Zhang, C. Cao, and C. Wang, “One-step hyperentanglement purification and hyperdistillation with linear optics,” Opt. Express 23(7), 9284–9294 (2015). [CrossRef]  

62. T.-J. Wang, S.-C. Mi, and C. Wang, “Hyperentanglement purification using imperfect spatial entanglement,” Opt. Express 25(3), 2969–2982 (2017). [CrossRef]  

63. P.-S. Yan, L. Zhou, W. Zhong, and Y.-B. Sheng, “Feasible measurement-based entanglement purification in linear optics,” Opt. Express 29(6), 9363–9384 (2021). [CrossRef]  

64. C.-C. Luo, L. Zhou, W. Zhong, and Y.-B. Sheng, “Purification for hybrid logical qubit entanglement,” Quantum Inf. Process. 21(8), 300 (2022). [CrossRef]  

65. J.-W. Pan, Z.-B. Chen, C.-Y. Lu, H. Weinfurter, A. Zeilinger, and M. Żukowski, “Multiphoton entanglement and interferometry,” Rev. Mod. Phys. 84(2), 777–838 (2012). [CrossRef]  

66. X.-L. Wang, Y.-H. Luo, H.-L. Huang, M.-C. Chen, Z.-E. Su, C. Liu, C. Chen, W. Li, Y.-Q. Fang, X. Jiang, J. Zhang, L. Li, N.-L. Liu, C.-Y. Lu, and J.-W. Pan, “18-qubit entanglement with six photons’ three degrees of freedom,” Phys. Rev. Lett. 120(26), 260502 (2018). [CrossRef]  

67. Z. Gao, T. Li, and Z. Li, “Deterministic measurement-device-independent quantum secret sharing,” Sci. China Phys. Mech. Astron. 63(12), 120311 (2020). [CrossRef]  

68. I. Šupić, J.-D. Bancal, Y. Cai, and N. Brunner, “Genuine network quantum nonlocality and self-testing,” Phys. Rev. A 105(2), 022206 (2022). [CrossRef]  

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

70. J.-W. Ying, L. Zhou, and Y.-B. Sheng, “Measurement-device-independent one-step quantum secure direct communication,” Chin. Phys. B 31(12), 120303 (2022). [CrossRef]  

71. 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]  

72. X.-X. Ju, W. Zhong, Y.-B. Sheng, and L. Zhou, “Measurement-device-independent quantum secret sharing with hyper-encoding,” Chin. Phys. B 31(10), 100302 (2022). [CrossRef]  

73. X.-M. Hu, W.-B. Xing, B.-H. Liu, D.-Y. He, H. Cao, Y. Guo, C. Zhang, H. Zhang, Y.-F. Huang, C.-F. Li, and G.-C. Guo, “Efficient distribution of high-dimensional entanglement through 11 km fiber,” Optica 7(7), 738–743 (2020). [CrossRef]  

74. K. Banaszek, “Fidelity balance in quantum operations,” Phys. Rev. Lett. 86(7), 1366–1369 (2001). [CrossRef]  

75. C. H. Bennett, D. P. DiVincenzo, J. A. Smolin, and W. K. Wootters, “Mixed-state entanglement and quantum error correction,” Phys. Rev. A 54(5), 3824–3851 (1996). [CrossRef]  

76. S. Hill and W. K. Wootters, “Entanglement of a pair of quantum bits,” Phys. Rev. Lett. 78(26), 5022–5025 (1997). [CrossRef]  

77. W. K. Wootters, “Entanglement of formation of an arbitrary state of two qubits,” Phys. Rev. Lett. 80(10), 2245–2248 (1998). [CrossRef]  

78. P.-S. Yan, L. Zhou, W. Zhong, and Y.-B. Sheng, “Advances in quantum entanglement purification,” Sci. China-Phys. Mech. Astron. 66(2), 250301 (2023).

79. I. Buluta, S. Ashhab, and F. Nori, “Natural and artificial atoms for quantum computation,” Rep. Prog. Phys. 74(10), 104401 (2011). [CrossRef]  

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

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

82. M. Nielsen and I. Chuang, Quantum computation and quantum information (Cambridge University, 2000).

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 (2)

Fig. 1.
Fig. 1. Schematics for nonlocal generalized quantum measurement of the product Pauli observable $\sigma _{\rm zA}\otimes \sigma _{\rm zB}$ using mixed entanglement. Here $\rho _S$ and $\rho _M$ represent the states of the system and the meter, respectively. Alice and Bob perform a controlled-NOT gate on their qubits before measuring their meters in $\sigma _{\rm zA}$ and $\sigma _{\rm zB}$ bases.
Fig. 2.
Fig. 2. Schematics of nonlocal generalized measurement on bipartite spin products with mixed entanglement in a linear optical system. The photon pairs are generated by an SPDC source. The photon polarization is used as the system, and the spatial mode is used as the meter. The beam splitter (BS) implements a Hadamard operation on the spatial mode. Polarization beam splitter (PBS) transmits $H$-polarized photons and reflects $V$-polarized ones, and thus can be referred to as a CNOT gate on the polarization (control) and spatial mode (target) of the same photon. HWP and QWP are, respectively, half-wave plate and quarter-wave plate, which are combined to implement polarization rotations. The upper (lower) modes are denoted as $a_{1}$ and $b_{1}$ ($a_{2}$ and $b_{2}$) for photons sent to Alice and Bob, respectively.

Tables (2)

Tables Icon

Table 1. Relationship between local measurement results and global results of the nonlocal product measurement.

Tables Icon

Table 2. Relationship between local measurement results and the global results of the nonlocal generalized product measurement of σ z A p σ z B p on two polarized photons.

Equations (23)

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

ρ S = a 1 | Φ + S Φ + | + a 2 | Ψ + S Ψ + | + a 3 | Ψ S Ψ | + a 4 | Φ S Φ | ,
| Φ ± S = 1 2 ( | 00 S ± | 11 S ) , | Ψ ± S = 1 2 ( | 01 S ± | 10 S ) ,
ρ M = F 1 | Φ + M Φ + | + F 2 | Ψ + M Ψ + | ,
| Φ ± M = 1 2 ( | 00 M ± | 11 M ) , | Ψ ± M = 1 2 ( | 01 M ± | 10 M ) .
| Ψ 1 = 1 2 ( | a 1 b 1 + | a 2 b 2 ) ( | H H + | V V ) .
ρ 1 = ρ s 1 ρ p 1 = ( F 1 | Φ + s Φ + | + F 2 | Φ s Φ | ) ( a 1 | Φ + p Φ + | + a 2 | Ψ + p Ψ + | + a 3 | Ψ p Ψ | + a 4 | Φ p Φ | ) .
| Φ ± s = 1 2 ( | a 1 b 1 M ± | a 2 b 2 M ) , | Ψ ± s = 1 2 ( | a 1 b 2 M ± | a 2 b 1 M ) .
| Φ ± p = 1 2 ( | H H S ± | V V S ) , | Ψ ± p = 1 2 ( | H V S ± | V H S ) ,
ρ p 2 = c 1 | Φ + p Φ + | + c 2 | Ψ + p Ψ + | + c 3 | Ψ p Ψ | + c 4 | Φ p Φ | ,
ρ 2 = ρ s 2 ρ p 2 ,
ρ 3 = F 1 ρ s p 1 + F 2 ρ s p 2 ,
ρ s p 1 = ( c 1 | Φ + p Φ + | + c 4 | Φ p Φ | ) | Φ + s Φ + | + ( c 2 | Ψ + p Ψ + | + c 3 | Ψ p Ψ | ) | Ψ + s Ψ + | , ρ s p 2 = ( c 1 | Φ + p Φ + | + c 4 | Φ p Φ | ) | Ψ + s Ψ + | + ( c 2 | Ψ + p Ψ + | + c 3 | Ψ p Ψ | ) | Φ + s Φ + | .
M + + = F 1 2 ( Π H H + Π V V ) + F 2 2 ( Π H V + Π V H ) , M + = F 2 2 ( Π H H + Π V V ) + F 1 2 ( Π H V + Π V H ) , M + = F 2 2 ( Π H H + Π V V ) + F 1 2 ( Π H V + Π V H ) , M = F 1 2 ( Π H H + Π V V ) + F 2 2 ( Π H V + Π V H ) ,
M + = F 1 ( Π H H + Π V V ) + F 2 ( Π H V + Π V H ) , M = F 2 ( Π H H + Π V V ) + F 1 ( Π H V + Π V H ) .
E ± = 1 2 [ σ I ± ( F 1 F 2 ) σ z A σ z B ] ,
S = F 1 F 2 , F 1 F 2 ,
ρ p f = M i ρ 3 M i T r ( M i M i ρ 3 ) ,
E ( ρ ) = E [ C ( ρ ) ] , E ( C ) = h ( 1 1 C 2 ) ,
C ( ρ ) = max { 0 , λ 1 λ 2 λ 3 λ 4 } .
ρ ~ = ( σ y σ y ) ρ ( σ y σ y ) ,
ρ ~ s 2 = 1 2 ( F 1 0 0 F 1 0 F 2 F 2 0 0 F 2 F 2 0 F 1 0 0 F 1 ) .
N = ρ s 2 ρ ~ s 2 = 1 2 ( F 1 2 0 0 F 1 2 0 F 2 2 F 2 2 0 0 F 2 2 F 2 2 0 F 1 2 0 0 F 1 2 ) .
C 1 = F 1 F 2 , F 1 F 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.