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

Complete and faithful hyperentangled-Bell-state analysis of photon systems using a failure-heralded and fidelity-robust quantum gate

Open Access Open Access

Abstract

Hyperentangled-Bell-state analysis (HBSA) represents a key step in many quantum information processing schemes that utilize hyperentangled states. In this paper, we present a complete and faithful HBSA scheme for two-photon quantum systems hyperentangled in both the polarization and spatial-mode degrees of freedom, using a failure-heralded and fidelity-robust quantum swap gate for the polarization states of two photons (P-SWAP gate), constructed with a singly charged semiconductor quantum dot (QD) in a double-sided optical microcavity (double-sided QD-cavity system) and some linear-optical elements. Compared with the previously proposed complete HBSA schemes using different auxiliary tools such as parity-check quantum nondemonlition detectors or additional entangled states, our scheme significantly simplifies the analysis process and saves the quantum resource. Unlike the previous schemes based on the ideal optical giant circular birefringence induced by a single-electron spin in a double-sided QD-cavity system, our scheme guarantees the robust fidelity and relaxes the requirement on the QD-cavity parameters. These features indicate that our scheme may be more feasible and useful in practical applications based on the photonic hyperentanglement.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Quantum entanglement plays an important role in quantum information science and technology. It is a unique quantum resource for many branches of quantum information processing (QIP), such as quantum key distribution [1,2], quantum dense coding [3,4], quantum teleportation [5], quantum secret sharing [6], quantum secure direct communication [7,8], and quantum computation [9,10]. Of the many different kinds of entangled states, Bell states of two qubit systems, which are the maximally entangled and mutually orthogonal basis states for the Hilbert space, are the most popular and crucial ones. In many quantum cryptography and quantum communication schemes, a key step is the implementation of the joint Bell-state measurement [11]. The number of the bases that can be unambiguously discriminated usually determines the capacity of the quantum communication process. Therefore, much effort has been made to perform the Bell-state analysis (BSA) on various quantum systems, and the ultimate goal is the complete and faithful discrimination of the Bell states.

Photon systems have been identified as ideal candidates for QIP. However, the complete BSA of photon pairs in one degree of freedom (DOF) cannot be achieved using only linear-optical elements. It has been proven that the four two-photon Bell states can only be classified into three groups, and the optimal success probability of state analysis is 50% with only linear optics [1214]. By the use of a quantum nondemonlition detector (QND), the success probability can be improved to unity [15]. On the other hand, photons possess several DOFs, such as polarization, spatial mode, time bin, frequency, and orbital angular momentum (OAM), and each DOF can be manipulated independently. In particular, photonic hyperentanglement [16,17], which involves photons simultaneously entangled in several DOFs, can be used to realize the complete BSA in one DOF [1824]. In 1998, Kwiat et al. [18] first investigated the method to completely distinguish the four polarization Bell states based on the hyperentanglement in both the polarization and momentum DOFs. In 2003, Walborn et al. [19] proposed a hyperentanglement-based scheme for the complete BSA in the polarization DOF or the momentum DOF with linear optics. The complete BSA experiments have been demonstrated with polarization-time-bin [21], polarization-momentum [22], and polarization-OAM [23] hyperentanglements. In 2017, Williams et al. [24] reported the demonstration of superdense coding over optical fiber links with a complete BSA enabled by polariztion-time-bin hyperentanglement. In addition to the complete BSA, hyperentanglement has many important applications, such as deterministic entanglement purification [2529], high-capacity quantum communication [3033], and efficient quantum computation [3436]. Hence, the generation [3739] and manipulation [4042] of hyperentanglement have been widely studied in the past years.

Hyperentangled-Bell-state analysis (HBSA) of photons in more than one DOF, which is very useful for QIP based on the photonic hyperentanglement, has attracted much attention in recent years [4346]. Considering two photons simultaneously entangled in two qubitlike DOFs, there are 16 orthogonal hyperentangled Bell states in a large Hibert space. It is shown that the 16 two-photon hyperentangled Bell states can only be classified into seven groups with only linear optics. When QNDs are utilized, the 16 states can be discriminated completely. In 2010, Sheng et al. [47] proposed the first two-step HBSA scheme for the complete discrimination of the 16 hyperentangled Bell states in the polarization and spatial-mode DOFs, using QNDs constructed with cross-Kerr nonlinearities. In this scheme, the spatial-mode entanglement is analyzed in the first step and the two photons are preserved using two different QNDs. The polarization entanglement is analyzed in the second step with the information obtained in the first step and another QND. The principle can be generalized to perform the hyperentangled Greenberger-Horne-Zeilinger-state analysis [48]. In 2016, Li et al. [49] proposed an efficient self-assisted complete polarization-spatial HBSA scheme, using less QNDs constructed with cross-Kerr nonlinearities. This scheme preserves not only the photons but also the spatial-mode entanglement in the first step, and uses the preserved spatial-mode entanglement as an ancillary to analyze the polarization entanglement in the second step. The self-assisted principle can also be used to analysis the polarization-time-bin hyperentanglement [50]. Recently, an interesting scheme was presented for the complete analysis of 16 hyperentangled Bell states in the polarization and first longitudinal momentum DOFs, assisted by the auxiliary hyperentanglement in the frequency and second longitudinal momentum DOFs [51].

The singly charged semiconductor quantum dot (QD) is a promising solid-state physical system for QIP. The QD-confined electron spin can be used to store and process quantum information due to the long electron-spin coherence time (typically around several microseconds), which is limited by the spin-relaxation time. And the spin initialization, manipulation, and measurement have been well investigated in the past decades. Current micro- and nano-fabrication techniques provide very efficient ways to incorporate the QDs into optical resonant cavities, which is of great importance for the realization of coherent manipulations of quantum information and the development of practical quantum devices [5255]. For example, based on the optical giant circular birefringence (GCB) induced by a single-electron spin in a QD inside a single-sided or double-sided optical microcavity due to the effect of cavity quantum electrodynamics (QED), one can realize different types of photon-spin quantum entangling gate (quantum interface) [5658], which can be widely used in designing scalable and all-optical QIP schemes [5963], including photonic complete BSA [6466] and HBSA [6771] schemes. In 2012, Ren et al. [67] proposed a complete HBSA scheme using QNDs constructed with single-sided QD-cavity systems, based on the ideal GCB in the reflection geometry. In the same year, Wang et al. [68] presented a different complete HBSA scheme with double-sided QD-cavity systems based on the ideal GCB in the reflection/transmission geometry. In 2016, Wang et al. [69] proposed an interesting complete HBSA scheme and a complete nondestructive HBSA scheme using error-detected blocks constructed with double-sided QD-cavity systems, which can work effectively even when the ideal GCB cannot be realized. In 2018, Zeng presented a self-assisted complete HBSA scheme using single-sided QD-cavity systems to construct QNDs [70].

In this paper, we present a complete and faithful HBSA scheme for two-photon quantum systems hyperentangled in both the polarization and spatial-mode DOFs, using a failure-heralded and fidelity-robust quantum swap gate for the polarization states of two hyper-encoded photons (P-SWAP gate). Compared with the previously proposed complete HBSA schemes, our scheme can also distinguish the 16 two-photon polarization-spatial hyperentangled Bell states completely, but has several distinct features: (i) Our scheme uses the P-SWAP gate rather than the parity-check QNDs, which significantly simplifies the state analysis process. There is no need to pause the state analysis process midway to read out the parity information and confirm the spatial modes of the two photons. After all the operations and measurements have been successfully performed on the photons, we can identify the initial hyperentangled Bell state according the final single-photon measurement outcomes. So the scheme can be implemented in one shot. (ii) The P-SWAP gate is constructed with only one double-sided QD-cavity system and some linear-optical elements, which greatly reduces the number of the required QD-cavity system and saves the quantum resource. (iii) The failure of the P-SWAP gate operation is heralded by the clicks of single-photon detectors in the quantum circuit and the gate fidelity is robust when the operation succeeds, since the reflection and transmission coefficients of the double-sided QD-cavity system only appear in the global coefficient of the final output state of the gate. This makes our scheme works faithfully (without the error introduced by the infidelity of the P-SWAP operation), even when the QD-cavity parameters cannot satisfy the conditions for the ideal GCB. Therefore, our scheme may be more feasible and useful, which can be faithfully used in various applications, such as quantum hyperdense coding, teleportation via hyperentangled channels, hyerentanglement swapping, and others.

2. Interaction between an incident photon and a singly charged QD in a double-sided optical microcavity

As shown in Fig. 1(a), we consider a singly charged semiconductor QD, for instance, a self-assembled In(Ga)As QD or a GaAs interface QD, placed at the antinode of a double-sided optical micropillar cavity, where the two GaAs/Al(Ga)As distributed Bragg reflectors (DBRs) and the transverse index guiding provide the three-dimensional confinement of light. The two DBRs of the microcavity are made partially reflective (double sided) and symmetric, and the cross section is made circular.

 figure: Fig. 1.

Fig. 1. (a) A singly charged QD in a double-sided optical micropillar cavity. The two distributed Bragg reflectors (DBRs) and the transverse index guiding provide the three-dimensional confinement of light. The two DBRs of the microcavity are made partially reflective (double sided) and symmetric, and the cross section is made circular. (b) The relevant energy levels and the optical spin selection rule for $X^-$ transitions due to the Pauli exclusion principle and the conservation of total spin angular momentum (see text).

Download Full Size | PDF

The optical properties of singly charged QDs are dominated by the spin-dependent optical transitions of the negatively charged exciton $X^-$ which consists of two electrons bound to one hole [72]. The relevant energy levels and the optical spin selection rule for $X^-$ transitions are shown in Fig. 1(b). Here, $\left | \uparrow \right \rangle$ and $\left | \downarrow \right \rangle$ represent the single-electron spin states with ${J_z} = \pm 1/2$, $\left | \Uparrow \right \rangle$ and $\left | \Downarrow \right \rangle$ represent the heavy-hole spin states with ${J_z} = \pm 3/2$, and the two electron spins are antiparallel in the $X^-$ states $\left | { \uparrow \downarrow \Uparrow } \right \rangle$ and $\left | { \downarrow \uparrow \Downarrow } \right \rangle$. The spin-quantization axis $z$ is along the QD growth direction, i.e., the normal direction of the cavity and the propagation direction of the input (or output) light. When we consider an incident single photon interacts with the charged QD, the photon spin ${s_z}$ can also be defined with respect to the $z$ axis. Due to the Pauli exclusion principle and the conservation of total spin angular momentum, there are only two optically allowed dipole transitions between the electron states and the $X^-$ states: the photon in the state $\left | {{L^ \downarrow }} \right \rangle$ or $\left | {{R^ \uparrow }} \right \rangle$ (${s_z} = + 1$) can only couple to the transition $\left | \uparrow \right \rangle \leftrightarrow \left | { \uparrow \downarrow \Uparrow } \right \rangle$, and the photon in the state $\left | {{R^ \downarrow }} \right \rangle$ or $\left | {{L^ \uparrow }} \right \rangle$ (${s_z} = - 1$) can only couple to the transition $\left | \downarrow \right \rangle \leftrightarrow \left | { \downarrow \uparrow \Downarrow } \right \rangle$ [see Fig. 1(b)]. Here, $\left | R \right \rangle$ and $\left | L \right \rangle$ represent the right- and left-circularly polarized states of the photon, respectively. The superscript arrows indicate the photon propagation directions along the $z$ axis.

The reflection and transmission coefficients of the double-sided QD-cavity system can be obtained by solving the Heisenberg equations of motion for the cavity-field operator $\hat a$ and the $X^-$ dipole operator ${\hat \sigma _ - }$ in the interaction picture, and the input-output relations [58]

$$\begin{aligned}&\frac{{d\hat a}}{{dt}} = - \left[ {i\left( {{\omega _c} - \omega } \right) + \kappa + \frac{{{\kappa _s}}}{2}} \right]\hat a - g{\hat \sigma _ - } - \sqrt \kappa {\hat a_{in}} - \sqrt \kappa {\hat a_{in}}^\prime + \hat H, \\ &\frac{{d{{\hat \sigma }_ - }}}{{dt}} = - \left[ {i\left( {{\omega _{{X^ - }}} - \omega } \right) + \frac{\gamma }{2}} \right]{{\hat \sigma }_ - } - g{{\hat \sigma }_z}\hat a + \hat G,\\ &{{\hat a}_r} = {{\hat a}_{in}} + \sqrt \kappa \hat a,\\ &{{\hat a}_t} = {{\hat a}_{in}}^\prime + \sqrt \kappa \hat a, \end{aligned}$$
where $\omega$, $\omega _c$, and $\omega _{X^ - }$ are the frequencies of the incident photon, cavity mode, and $X^-$ transition, respectively. $\kappa$ and $\kappa _s$ are the cavity-field decay rates into the input/output modes and the leaky modes (side leakage), respectively. $\gamma$ is the $X^-$ dipole decay rate. $g$ is the $X^-$-cavity coupling strength. $\hat H$ and $\hat G$ are the noise operators. $\hat a_{in}$, $\hat a_{in}'$ and $\hat a_r$, $\hat a_t$ are the input- and output-field operators. In the weak-excitation approximation, we can obtain the steady-state reflection and transmission coefficients $r(\omega )$ and $t(\omega )$ for a coupled QD-cavity system as [58]
$$\begin{aligned}&r(\omega ) = 1 + t(\omega ), \\ &t(\omega ) = \frac{{ - \kappa \left[ {i({\omega _{{X^ - }}} - \omega ) + \frac{\gamma }{2}} \right]}}{{\left[ {i({\omega _{{X^ - }}} - \omega ) + \frac{\gamma }{2}} \right]\left[ {i({\omega _c} - \omega ) + \kappa + \frac{{{\kappa _s}}}{2}} \right] + {g^2}}}. \end{aligned}$$
By setting $g=0$, the reflection and transmission coefficients $r_0(\omega )$ and $t_0(\omega )$ for a bare cavity with QD uncoupled to the cavity can be obtained as
$$\begin{aligned}&{r_0}(\omega ) = \frac{{i({\omega _c} - \omega ) + \frac{{{\kappa _s}}}{2}}}{{i({\omega _c} - \omega ) + \kappa + \frac{{{\kappa _s}}}{2}}}, \\ &{t_0}(\omega ) = \frac{{ - \kappa }}{{i({\omega _c} - \omega ) + \kappa + \frac{{{\kappa _s}}}{2}}}. \end{aligned}$$
Due to the spin selection rule and the cavity-QED effect, if the electron spin in the QD is in the state $\left | \uparrow \right \rangle$, the photon in the state $\left | {{L^ \downarrow }} \right \rangle$ or $\left | {{R^ \uparrow }} \right \rangle$ feels a coupled QD-cavity system with the reflection coefficient $r$ and transmission coefficient $t$, whereas the photon in the state $\left | {{R^ \downarrow }} \right \rangle$ or $\left | {{L^ \uparrow }} \right \rangle$ feels a bare cavity with the reflection coefficient $r_0$ and transmission coefficient $t_0$. As photon polarization is commonly defined with respect to the direction of propagation, both the photon polarization and propagation direction will be flipped upon reflection. Conversely, if the electron spin in the QD is in the state $\left | \downarrow \right \rangle$, the photon in the state $\left | {{R^ \downarrow }} \right \rangle$ or $\left | {{L^ \uparrow }} \right \rangle$ feels a coupled QD-cavity system, whereas the photon in the state $\left | {{L^ \downarrow }} \right \rangle$ or $\left | {{R^ \uparrow }} \right \rangle$ feels a bare cavity. The interaction between the incident photon and the double-sided QD-cavity system can therefore be described as
$$\begin{aligned} &\left| {{L^ \downarrow }, \uparrow } \right\rangle \to r\left| {{R^ \uparrow }, \uparrow } \right\rangle + t\left| {{L^ \downarrow }, \uparrow } \right\rangle, \left| {{R^ \downarrow }, \uparrow } \right\rangle \to {r_0}\left| {{L^ \uparrow }, \uparrow } \right\rangle + {t_0}\left| {{R^ \downarrow }, \uparrow } \right\rangle, \\ &\left| {{R^ \uparrow }, \uparrow } \right\rangle \to r\left| {{L^ \downarrow }, \uparrow } \right\rangle + t\left| {{R^ \uparrow }, \uparrow } \right\rangle, \left| {{L^ \uparrow }, \uparrow } \right\rangle \to {r_0}\left| {{R^ \downarrow }, \uparrow } \right\rangle + {t_0}\left| {{L^ \uparrow }, \uparrow } \right\rangle,\\ &\left| {{R^ \downarrow }, \downarrow } \right\rangle \to r\left| {{L^ \uparrow }, \downarrow } \right\rangle + t\left| {{R^ \downarrow }, \downarrow } \right\rangle, \left| {{L^ \downarrow }, \downarrow } \right\rangle \to {r_0}\left| {{R^ \uparrow }, \downarrow } \right\rangle + {t_0}\left| {{L^ \downarrow }, \downarrow } \right\rangle,\\ &\left| {{L^ \uparrow }, \downarrow } \right\rangle \to r\left| {{R^ \downarrow }, \downarrow } \right\rangle + t\left| {{L^ \uparrow }, \downarrow } \right\rangle, \left| {{R^ \uparrow }, \downarrow } \right\rangle \to {r_0}\left| {{L^ \downarrow }, \downarrow } \right\rangle + {t_0}\left| {{R^ \uparrow }, \downarrow } \right\rangle. \end{aligned}$$
It has been shown that the double-sided QD-cavity system can show large reflectance and transmittance difference between the coupled and uncoupled cases, which is the so-called GCB. For example, when ${\omega _c} = {\omega _{{X^ - }}}$ and ${\kappa _s} \to 0$, the ideal GCB ($\left | {{r_0}} \right | \to 0$, $\left | {{t_0}} \right | \to 1$, $\left | r \right | \to 1$, and $\left | t \right | \to 0$) can be achieved in the strong-coupling regime with $g\;>\;\left ( {\kappa ,\gamma } \right )$ and in the central frequency regime with $\left | {\omega - {\omega _c}} \right |\;<\;\kappa$ [58,64]. Based on the ideal GCB in the reflection/transmission geometry, a kind of near-deterministic photon-spin quantum entangling gate (quantum interface) can be built, which has been widely used in QIP schemes. However, although significant progress has been made, these conditions remain represent a big challenge for practical QD-cavity systems. If the QD-cavity parameters cannot strictly satisfy the conditions for the ideal GCB, the practical fidelities of many QIP schemes would inevitably be reduced.

3. An error-detected circuit unit constructed with a double-sided QD-cavity system and some linear-optical elements

We can construct an error-detected circuit unit with a double-sided QD-cavity system and some linear-optical elements for fidelity-robust QIP. The schematic diagram of the error-detected circuit unit is shown in Fig. 2. Here, C is an optical circulator and D is a single-photon detector. BS represents a 50:50 beam splitter, which performs the Hadamard operation [$\left | {{i_1}} \right \rangle \leftrightarrow \left ( {\left | {{j_1}} \right \rangle + \left | {{j_2}} \right \rangle } \right )/\sqrt 2$ and $\left | {{i_2}} \right \rangle \leftrightarrow \left ( {\left | {{j_1}} \right \rangle - \left | {{j_2}} \right \rangle } \right )/\sqrt 2$] on the spatial-mode DOF of single photons. H (H$_{j1}$ or H$_{j2}$) represents a half-wave plate oriented at ${22.5^ \circ }$, which performs the Hadamard operation [$\left | R \right \rangle \leftrightarrow \left ( {\left | R \right \rangle + \left | L \right \rangle } \right )/\sqrt 2$ and $\left | L \right \rangle \leftrightarrow \left ( {\left | R \right \rangle - \left | L \right \rangle } \right )/\sqrt 2$] on the polarization DOF of single photons. M (M$_{j1}$ or M$_{j2}$) is a mirror. Suppose we always inject the right-circularly polarized photon $\left | R \right \rangle$ into the circuit unit from the path $i_1$ (see Fig. 2). If the electron spin in the QD is initially in a superposition state $\left | + \right \rangle = \left ( {\left | \uparrow \right \rangle + \left | \downarrow \right \rangle } \right )/\sqrt 2$, after the photon passes through the block composed of BS, H$_{j1}$, H$_{j2}$, M$_{j1}$, M$_{j2}$, and QD-cavity system, the total state for the photon and the electron spin is changed from $\left | {{\Phi _0}} \right \rangle = \left | R \right \rangle \otimes \left | + \right \rangle$ to

$$\left| {{\Phi _1}} \right\rangle = D\left| R \right\rangle \left| {{i_1}} \right\rangle \left| + \right\rangle + T\left| L \right\rangle \left| {{i_2}} \right\rangle \left| - \right\rangle,$$
where $D = (r + t + {r_0} + {t_0})/2$ and $T = (r + t - {r_0} - {t_0})/2$ represent the generalized reflection and transmission coefficients of the block, respectively [69]. $\left | - \right \rangle = \left ( {\left | \uparrow \right \rangle - \left | \downarrow \right \rangle } \right )/\sqrt 2$. Similarly, if the electron spin in the QD is initially in the superposition state $\left | - \right \rangle$, after the photon passes through the block, the total state for the photon and the electron spin is changed from $\left | {{\Psi _0}} \right \rangle = \left | R \right \rangle \otimes \left | - \right \rangle$ to
$$\left| {{\Psi _1}} \right\rangle = D\left| R \right\rangle \left| {{i_1}} \right\rangle \left| - \right\rangle + T\left| L \right\rangle \left| {{i_2}} \right\rangle \left| + \right\rangle.$$

 figure: Fig. 2.

Fig. 2. Schematic diagram of the error-detected circuit unit constructed with a double-sided QD-cavity system and some linear-optical elements. Here, C is an optical circulator and D is a single-photon detector. BS represents a 50:50 beam splitter, which performs the spatial-mode Hadamard operation [$\left | {{i_1}} \right \rangle \leftrightarrow \left ( {\left | {{j_1}} \right \rangle + \left | {{j_2}} \right \rangle } \right )/\sqrt 2$ and $\left | {{i_2}} \right \rangle \leftrightarrow \left ( {\left | {{j_1}} \right \rangle - \left | {{j_2}} \right \rangle } \right )/\sqrt 2$]. H (H$_{j1}$ or H$_{j2}$) represents a half-wave plate oriented at ${22.5^ \circ }$, which performs the polarization Hadamard operation [$\left | R \right \rangle \leftrightarrow \left ( {\left | R \right \rangle + \left | L \right \rangle } \right )/\sqrt 2$ and $\left | L \right \rangle \leftrightarrow \left ( {\left | R \right \rangle - \left | L \right \rangle } \right )/\sqrt 2$]. M (M$_{j1}$ or M$_{j2}$) is a mirror.

Download Full Size | PDF

From Eqs. (5) and (6), one can observe that there are two terms in each final state, in which the first term is usually defined as the error term and the second term is our desired state. If the photon is transmitted through the block (which corresponds to the second term with the probability amplitude $T$), the polarization of the photon would be flipped and the spin superposition state of the electron would be changed from $\left | + \right \rangle$ to $\left | - \right \rangle$ or from $\left | - \right \rangle$ to $\left | + \right \rangle$. Then we can utilize the transmitted photon and the QD-cavity system to accomplish the task of fidelity-robust QIP. Conversely, if the photon is reflected by the block (which corresponds to the first term with the probability amplitude $D$), the electron spin state would not change and the reflected photon would be detected by the single-photon detector D. Thus the click of the detector indicates the case in which the error is detected and the corresponding task of the circuit unit runs fail. Such an error-detected circuit unit has been used to implement quantum logic gates [7377], entanglement generation and analysis [69,74], and entanglement purification and concentration [7880] very recently.

4. Implementation of a failure-heralded and fidelity-robust quantum swap gate for the polarization states of two hyper-encoded photons

Suppose two photons $A$ and $B$ are hyper encoded in the polarization and spatial-mode DOFs as

$$\begin{aligned}&{\left| \varphi \right\rangle _A} = {({\alpha _1}\left| R \right\rangle + {\beta _1}\left| L \right\rangle )_A} \otimes {({\chi _1}\left| {{a_1}} \right\rangle + {\delta _1}\left| {{a_2}} \right\rangle )_A},\\ &{\left| \varphi \right\rangle _B} = {({\alpha _2}\left| R \right\rangle + {\beta _2}\left| L \right\rangle )_B} \otimes {({\chi _2}\left| {{b_1}} \right\rangle + {\delta _2}\left| {{b_2}} \right\rangle )_B}, \end{aligned}$$
where the subscripts $A$ and $B$ denote the two photons. $\left | {{a_1}} \right \rangle$ and $\left | {{a_2}} \right \rangle$ ($\left | {{b_1}} \right \rangle$ and $\left | {{b_2}} \right \rangle$) are the different spatial-mode states of the photon $A$ ($B$), respectively. ${\left | {{\alpha _k}} \right |^2} + {\left | {{\beta _k}} \right |^2} = {\left | {{\chi _k}} \right |^2} + {\left | {{\delta _k}} \right |^2} = 1$ ($k = 1,2$). A quantum swap gate for the polarization states of two hyper-encoded photons (P-SWAP gate) can exchange the polarization states of the two photons $A$ and $B$ without affecting their spatial-mode states.

The quantum circuit for the implementation of a failure-heralded and fidelity-robust P-SWAP gate with the error-detected circuit unit is shown in Fig. 3. Here, CPBS represents a circularly polarizing beam splitter in the basis $\left \{ {\left | R \right \rangle ,\left | L \right \rangle } \right \}$, which transmits the right-circularly polarized photon component and reflects the left-circularly polarized photon component. X represents a half wave plate oriented at ${45^ \circ }$, which performs the polarization bit-flip operation ${\sigma _x} = \left | R \right \rangle \left \langle L \right | + \left | L \right \rangle \left \langle R \right |$. UBS represents an unbalanced beam splitter with the transmission coefficient $T$. S is an optical switch. The electron spin in the QD is initially prepared in the state $\left | + \right \rangle$ and the function of the error-detected circuit unit is described by Eqs. (5) and (6). To implement the P-SWAP gate operation for the two photons $A$ and $B$, we input the photons $A$ and $B$ into the quantum circuit through the spatial modes $l_1$ and $l_2$ ($l=a,b$) in sequence (see Fig. 3). In both the spatial modes $l_1$ and $l_2$, after each photon passes through the CPBS, the left-circularly polarized photon component will pass through the X, and then be injected into the error-detected circuit unit with the reflection coefficient $D$ and transmission coefficient $T$. The right-circularly polarized photon component will pass through the UBS. If there is no click of the single-photon detectors, the two components of each photon are combined again at the other CPBS. After that, a polarization Hadamard operation is performed on each photon through the H and a spin Hadamard operation [$\left | \uparrow \right \rangle \leftrightarrow \left ( {\left | \uparrow \right \rangle + \left | \downarrow \right \rangle } \right )/\sqrt 2$ and $\left | \downarrow \right \rangle \leftrightarrow \left ( {\left | \uparrow \right \rangle - \left | \downarrow \right \rangle } \right )/\sqrt 2$] is performed on the electron spin in the QD. After the above operations, the total state for the two photons and the electron is changed from $\left | {{\Omega _0}} \right \rangle = {\left | \varphi \right \rangle _A} \otimes {\left | \varphi \right \rangle _B} \otimes \left | + \right \rangle$ to (without normalization)

$$\begin{aligned}\left| {{\Omega _1}} \right\rangle =& \frac{{{T^2}}}{2}[{\alpha _1}{\alpha _2}{(\left| {RR} \right\rangle + \left| {RL} \right\rangle + \left| {LR} \right\rangle + \left| {LL} \right\rangle )_{AB}}\left| \uparrow \right\rangle \\ &+ {\alpha _1}{\beta _2}{(\left| {RR} \right\rangle - \left| {RL} \right\rangle + \left| {LR} \right\rangle - \left| {LL} \right\rangle )_{AB}}\left| \downarrow \right\rangle\\ &+ {\beta _1}{\alpha _2}{(\left| {RR} \right\rangle + \left| {RL} \right\rangle - \left| {LR} \right\rangle - \left| {LL} \right\rangle )_{AB}}\left| \downarrow \right\rangle\\ &+ {\beta _1}{\beta _2}{(\left| {RR} \right\rangle - \left| {RL} \right\rangle - \left| {LR} \right\rangle + \left| {LL} \right\rangle )_{AB}}\left| \uparrow \right\rangle]\\ &\otimes {({\chi _1}\left| {{a_1}} \right\rangle + {\delta _1}\left| {{a_2}} \right\rangle )_A} \otimes {({\chi _2}\left| {{b_1}} \right\rangle + {\delta _2}\left| {{b_2}} \right\rangle )_B}. \end{aligned}$$

 figure: Fig. 3.

Fig. 3. Quantum circuit for the implementation of a failure-heralded and fidelity-robust P-SWAP gate with the error-detected circuit unit. Here, CPBS represents a circularly polarizing beam splitter in the basis $\left \{ {\left | R \right \rangle ,\left | L \right \rangle } \right \}$, which transmits the right-circularly polarized photon component and reflects the left-circularly polarized photon component. X represents a half wave plate oriented at ${45^ \circ }$, which performs the polarization bit-flip operation ${\sigma _x} = \left | R \right \rangle \left \langle L \right | + \left | L \right \rangle \left \langle R \right |$. UBS represents an unbalanced beam splitter with the transmission coefficient $T$. S is an optical switch. Other devices are the same as that in Fig. 2.

Download Full Size | PDF

Subsequently, by using the optical switches, we let the photons pass through the quantum circuit for the second time. After the same operations as that in the first time, if there is no click of the single-photon detectors, the total state becomes

$$\begin{aligned}\left| {{\Omega _2}} \right\rangle =& \frac{{{T^4}}}{{\sqrt 2 }}[{({\alpha _2}\left| R \right\rangle + {\beta _2}\left| L \right\rangle )_A}{({\alpha _1}\left| R \right\rangle + {\beta _1}\left| L \right\rangle )_B}\left| \uparrow \right\rangle \\ &+ {({\alpha _2}\left| R \right\rangle - {\beta _2}\left| L \right\rangle )_A}{({\alpha _1}\left| R \right\rangle - {\beta _1}\left| L \right\rangle )_B}\left| \downarrow \right\rangle ]\\ &\otimes {({\chi _1}\left| {{a_1}} \right\rangle + {\delta _1}\left| {{a_2}} \right\rangle )_A} \otimes {({\chi _2}\left| {{b_1}} \right\rangle + {\delta _2}\left| {{b_2}} \right\rangle )_B}. \end{aligned}$$
Finally, we can measure the electron spin in the basis $\{ \left | \uparrow \right \rangle ,\left | \downarrow \right \rangle \}$ to complete the P-SWAP operation. In detail, if we measure the electron spin in the state $\left | \uparrow \right \rangle$, the two photons collapse to the state
$$\begin{aligned}\left| {{\Omega _3}} \right\rangle =& {T^4}{({\alpha _2}\left| R \right\rangle + {\beta _2}\left| L \right\rangle )_A} \otimes {({\chi _1}\left| {{a_1}} \right\rangle + {\delta _1}\left| {{a_2}} \right\rangle )_A} \\ &\otimes {({\alpha _1}\left| R \right\rangle + {\beta _1}\left| L \right\rangle )_B} \otimes {({\chi _2}\left| {{b_1}} \right\rangle + {\delta _2}\left| {{b_2}} \right\rangle )_B}, \end{aligned}$$
which is the ideal output state of the P-SWAP gate; if we measure the electron spin in the state $\left | \downarrow \right \rangle$, we can perform a polarization phase-flip operation ${\sigma _z} = \left | R \right \rangle \left \langle R \right | - \left | L \right \rangle \left \langle L \right |$ on each photon to transform the two photons into the desired output state.

In implementing the P-SWAP gate, the clicks of the single-photon detectors in the quantum circuit herald the failure of the gate operation. However, once the operation succeeds, the gate fidelity is, in principle, in unity, since the reflection and transmission coefficients of the double-sided QD-cavity system only appear in the global coefficient of the final state. That is, the gate fidelity is robust to the QD-cavity parameters. Without considering the photon loss due to the absorption and scattering of the linear-optical elements, and the inefficiency of the single-photon detectors, the success probability of the P-SWAP gate can be directly calculated to be $\eta = {\left | T \right |^8}$. We will discuss the success probability by considering currently achievable QD-cavity parameters in Sec. 6.

5. Complete two-photon polarization-spatial hyperentangled-Bell-state analysis using the P-SWAP gate

A two-photon hyperentangled Bell state in both the polarization and spatial-mode DOFs can be written as

$$\left| \varphi \right\rangle _{AB}^{p,s} = \left| \varepsilon \right\rangle _{AB}^p \otimes \left| \xi \right\rangle _{AB}^s,$$
where the subscripts $A$ and $B$ represent the two hyperentangled photons. The superscript $p$ represents the polarization DOF, and $\left | \varepsilon \right \rangle _{AB}^p$ is one of the four polarization Bell states
$$\begin{aligned} &\left| {{\phi ^ \pm }} \right\rangle _{AB}^p = \frac{1}{{\sqrt 2 }}{(\left| {RR} \right\rangle \pm \left| {LL} \right\rangle )_{AB}}, \\ &\left| {{\psi ^ \pm }} \right\rangle _{AB}^p = \frac{1}{{\sqrt 2 }}{(\left| {RL} \right\rangle \pm \left| {LR} \right\rangle )_{AB}}. \end{aligned}$$
The superscript $s$ represents the spatial-mode DOF, and $\left | \xi \right \rangle _{AB}^s$ is one of the four spatial-mode Bell states
$$\begin{aligned}&\left| {{\phi ^ \pm }} \right\rangle _{AB}^s = \frac{1}{{\sqrt 2 }}{(\left| {{a_1}{b_1}} \right\rangle \pm \left| {{a_2}{b_2}} \right\rangle )_{AB}}, \\ &\left| {{\psi ^ \pm }} \right\rangle _{AB}^s = \frac{1}{{\sqrt 2 }}{(\left| {{a_1}{b_2}} \right\rangle \pm \left| {{a_2}{b_1}} \right\rangle )_{AB}}. \end{aligned}$$
Our task is to completely distinguish the 16 hyperentangled Bell states.

The schematic diagram of our complete and faithful HBSA scheme for polarization-spatial hyperentangled two-photon systems is shown in Fig. 4. Here, we use the failure-heralded and fidelity-robust P-SWAP gate with only one QD-cavity system and some linear-optical elements to achieve the complete HBSA. Suppose a two-photon system $AB$ is in one of the 16 hyperentangled Bell states as described by Eq. (11), the process for the complete HBSA can be described as follows.

 figure: Fig. 4.

Fig. 4. Schematic diagram of our complete and faithful HBSA scheme for hyperentangled two-photon systems in the polarization and spatial-mode DOFs.

Download Full Size | PDF

First, we let the photon $B$ pass through the block composed of X$_1$, CPBS$_1$, and X$_2$, which can exchange the polarization and spatial-mode states of the photon $B$. Then the initial two-photon hyperentangled Bell state is transformed to a state with the polarization DOF of the photon $A$ ($B$) entangled to the spatial-mode DOF of the photon $B$ ($A$). Second, we let the two photons $A$ and $B$ pass through the failure-heralded and fidelity-robust P-SWAP gate in sequence, to exchange the polarization states of the two photons. After the successful P-SWAP operation, the two-photon state is converted into a product of two single-photon hybrid entangled states [81,82]. For example, if the initial state is ${\left | {{\phi ^ + }} \right \rangle ^p}\left | {{\phi ^ + }} \right \rangle _{AB}^s$, after the above operations, the state transformation is

$${\left| {{\phi ^ + }} \right\rangle ^p}\left| {{\phi ^ + }} \right\rangle _{AB}^s \to \frac{1}{{\sqrt 2 }}{\left( {\left| {R{a_1}} \right\rangle + \left| {L{a_2}} \right\rangle } \right)_A} \otimes \frac{1}{{\sqrt 2 }}{\left( {\left| {R{b_1}} \right\rangle + \left| {L{b_2}} \right\rangle } \right)_B}.$$
If the initial state is one of the other 15 states, the transformation can also be written in the similar form. At last, the photon $A$ ($B$) passes through the block composed of CPBS$_2$, H$_1$, and H$_2$ (CPBS$_3$, H$_3$, and H$_4$). After the two photons pass through the quantum circuit and before they are measured, the two-photon state evolves as
$$\begin{aligned}&{\left| {{\phi ^ + }} \right\rangle ^p}\left| {{\phi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_2},R} \right\rangle _A}{\left| {{b_2},R} \right\rangle _B},\\ &{\left| {{\phi ^ - }} \right\rangle ^p}\left| {{\phi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_2},R} \right\rangle _A}{\left| {{b_2},L} \right\rangle _B},\\ &{\left| {{\psi ^ + }} \right\rangle ^p}\left| {{\phi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_2},R} \right\rangle _A}{\left| {{b_1},R} \right\rangle _B},\\ &{\left| {{\psi ^ - }} \right\rangle ^p}\left| {{\phi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_2},R} \right\rangle _A}{\left| {{b_1},L} \right\rangle _B},\\ &{\left| {{\phi ^ + }} \right\rangle ^p}\left| {{\phi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_2},L} \right\rangle _A}{\left| {{b_2},R} \right\rangle _B},\\ &{\left| {{\phi ^ - }} \right\rangle ^p}\left| {{\phi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_2},L} \right\rangle _A}{\left| {{b_2},L} \right\rangle _B},\\ &{\left| {{\psi ^ + }} \right\rangle ^p}\left| {{\phi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_2},L} \right\rangle _A}{\left| {{b_1},R} \right\rangle _B},\\ &{\left| {{\psi ^ - }} \right\rangle ^p}\left| {{\phi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_2},L} \right\rangle _A}{\left| {{b_1},L} \right\rangle _B},\\ &{\left| {{\phi ^ + }} \right\rangle ^p}\left| {{\psi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_1},R} \right\rangle _A}{\left| {{b_2},R} \right\rangle _B},\\ &{\left| {{\phi ^ - }} \right\rangle ^p}\left| {{\psi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_1},R} \right\rangle _A}{\left| {{b_2},L} \right\rangle _B},\\ &{\left| {{\psi ^ + }} \right\rangle ^p}\left| {{\psi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_1},R} \right\rangle _A}{\left| {{b_1},R} \right\rangle _B},\\ &{\left| {{\psi ^ - }} \right\rangle ^p}\left| {{\psi ^ + }} \right\rangle _{AB}^s \to {\left| {{a_1},R} \right\rangle _A}{\left| {{b_1},L} \right\rangle _B},\\ &{\left| {{\phi ^ + }} \right\rangle ^p}\left| {{\psi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_1},L} \right\rangle _A}{\left| {{b_2},R} \right\rangle _B},\\ &{\left| {{\phi ^ - }} \right\rangle ^p}\left| {{\psi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_1},L} \right\rangle _A}{\left| {{b_2},L} \right\rangle _B},\\ &{\left| {{\psi ^ + }} \right\rangle ^p}\left| {{\psi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_1},L} \right\rangle _A}{\left| {{b_1},R} \right\rangle _B},\\ &{\left| {{\psi ^ - }} \right\rangle ^p}\left| {{\psi ^ - }} \right\rangle _{AB}^s \to {\left| {{a_1},L} \right\rangle _A}{\left| {{b_1},L} \right\rangle _B}. \end{aligned}$$
Therefore, the photons $A$ and $B$ can be measured independently in both the spatial-mode and polarization DOFs (see Fig. 4). The relationship between the measurement outcomes and the initial hyperentangled Bell states of the two-photon system $AB$ is shown in Table 1. According to Table 1, one can obtain the complete analysis on the hyperentangled state of the two-photon system $AB$. That is, the scheme shown in Fig. 4 can be used for the complete two-photon polarization-spatial HBSA.

Tables Icon

Table 1. The relationship between the final measurement outcomes of the two photons in the spatial-mode and polarization DOFs and the initial hyperentangled Bell states of the two-photon system. The products of two single-photon hybrid entangled states after the successful P-SWAP operations are also contained in the table, where $\left | {{\phi ^ \pm }} \right \rangle _{A(B)}^{p,s} = {(\left | {R{a_1}({b_1})} \right \rangle \pm \left | {L{a_2}({b_2})} \right \rangle )_{A(B)}}/\sqrt 2$ and $\left | {{\psi ^ \pm }} \right \rangle _{A(B)}^{p,s} = {(\left | {R{a_2}({b_2})} \right \rangle \pm \left | {L{a_1}({b_1})} \right \rangle )_{A(B)}}/\sqrt 2$ represent the single-photon hybrid entangled states.

Our scheme uses the P-SWAP gate rather than the parity-check QNDs. There is no need to pause the state analysis process midway to read out the parity information and confirm the spatial modes of the two photons, which is apparently different from some of the previous complete HBSA schemes. After all the operations and measurements have been successfully performed on the photons, we can confirm the initial hyperentangled Bell state according the final measurement outcomes. So the scheme can be implemented in one shot.

Our scheme can also be directly used for analyzing the polarization-entangled Bell states, as the polarization-entangled Bell states can be seen as the specific hyperentangled Bell states with certain spatial modes. For example, suppose a two-photon system $AB$ is in one of the four polarization-entangled Bell states as described by Eq. (12), we can input the photons $A$ and $B$ into the quantum circuit shown in Fig. 4 through the input ports $a_1$ and $b_1$, which means the initial two-photon state has the decided spatial modes. After the same process as that for the complete HBSA and before the photons are measured, the two-photon state evolves as

$$\begin{aligned}&\left| {{\phi ^ + }} \right\rangle _{AB}^p\left| {{a_1}{b_1}} \right\rangle \to \frac{1}{{\sqrt 2 }}{(\left| {{a_2},R} \right\rangle + \left| {{a_2},L} \right\rangle )_A}{\left| {{b_2},R} \right\rangle _B},\\ &\left| {{\phi ^ - }} \right\rangle _{AB}^p\left| {{a_1}{b_1}} \right\rangle \to \frac{1}{{\sqrt 2 }}{(\left| {{a_2},R} \right\rangle + \left| {{a_2},L} \right\rangle )_A}{\left| {{b_2},L} \right\rangle _B},\\ &\left| {{\psi ^ + }} \right\rangle _{AB}^p\left| {{a_1}{b_1}} \right\rangle \to \frac{1}{{\sqrt 2 }}{(\left| {{a_2},R} \right\rangle + \left| {{a_2},L} \right\rangle )_A}{\left| {{b_1},R} \right\rangle _B},\\ &\left| {{\psi ^ - }} \right\rangle _{AB}^p\left| {{a_1}{b_1}} \right\rangle \to \frac{1}{{\sqrt 2 }}{(\left| {{a_2},R} \right\rangle + \left| {{a_2},L} \right\rangle )_A}{\left| {{b_1},L} \right\rangle _B}. \end{aligned}$$
Therefore, one can measure the photon $B$ in both the spatial-mode and polarization DOFs, and obtain the complete analysis on the initial polarization-entangled Bell state of the two-photon system $AB$.

6. Discussion and summary

Up to now, we have fully presented our complete and faithful two-photon polarization-spatial HBSA scheme. The key element for our scheme is the failure-heralded and fidelity-robust P-SWAP gate. As mentioned above, the gate fidelity is robust to the QD-cavity parameters when the operation succeeds, as the reflection and transmission coefficients only appear in the global coefficient of the output state. This makes our scheme works faithfully (without the error introduced by the infidelity of the P-SWAP operation), which can be used in various applications, especially the safe and high-capacity quantum communication processes [78]. The efficiency of our scheme is mainly limited by the success probability of the P-SWAP gate, which has been calculated to be

$$\eta = {\left| T \right|^8}.$$
This indicates that the efficiency is constrained by the QD-cavity parameters. Figure 5 shows the calculated success probability $\eta$ as a function of $g/\kappa$ and $\kappa _s/\kappa$. Here, $\omega = {\omega _c} = {\omega _{{X^ - }}}$ is assumed and $\gamma = 0.1\kappa$ is taken by considering the practical QD-cavity parameters. From Fig. 5, one can see that the gate can be made very efficient by optimizing the QD-cavity system (e.g., by increasing the QD-cavity coupling strength and suppressing the cavity side leakage). As considered in [5658], if we set $g/\kappa = 2.4$ (which is achievable for the In(Ga)As QD-micropillar cavity system [83,84]), the success probability can achieve $76.46\%$ with $\kappa _s=0.05\kappa$.

 figure: Fig. 5.

Fig. 5. Calculated success probability of the P-SWAP gate as a function of $g/\kappa$ and $\kappa _s/\kappa$. Here, $\omega = {\omega _c} = {\omega _{{X^ - }}}$ is assumed and $\gamma = 0.1\kappa$ is taken by considering the practical QD-cavity parameters.

Download Full Size | PDF

Although the fidelity of the P-SWAP gate is robust to the QD-cavity parameters, there may exist some other imperfections that should be considered. As noted by Hu et al. [56,57], the perfect spin-selection rule only holds for an ideal QD that is symmetric in both the QD shape and the strain field distribution such that there is no spin-level mixing or splitting at zero magnetic field. A realistic QD which is generally asymmetric could be made symmetric via thermal annealing [85], tuning the QD size [86], or applying an electric/magnetic field [87]. On the other hand, the heavy-light hole mixing in a realistic QD can also affect the spin-selection rule [57,58]. For self-assembled In(Ga)As QDs, the hold mixing in the valence band is in the order of a few percent [88,89], which would reduce the gate fidelity by a few percent. The hold mixing could be reduced by engineering the shape and size of QDs or by utilizing different types of QDs.

The electron-spin coherence time $T_2$ in the QD is an important parameter in implementing our HBSA scheme. Considering the spin decoherence, the high-fidelity P-SWAP gate operation can only be achieved when the operational time is much shorter than the spin-coherence time. It has been shown that In(Ga)As or GaAs charged QDs have the long electron-spin coherence time (${T_2} \sim \mu$s) [90,91], which is limited by the spin-relaxation time (${T_1} \sim$ ms) [92,93]. Recent experiments also show that spin echo or dynamical decoupling techniques can be used to preserve the spin coherence. Based on the ultrafast optical spin echo technique, $T_2 = 1$ $\mu$s has been demonstrated in a In(Ga)As QD [94]. On the other hand, the exciton dephasing can also reduce the gate fidelity by amount of $\left [ {1 - \exp \left ( { - \tau /\Gamma } \right )} \right ]$, where $\tau$ is the cavity photon lifetime and $\Gamma$ is the exciton coherence time [58]. In self-assembled In(Ga)As QDs, as the optical coherence time of excitons can be ten times longer than the cavity lifetime [95], the optical dephasing of the $X^-$ can only slightly reduce the fidelity by less then $10\%$. The spin dephasing of the $X^-$ can be safely neglected in our considerations, as the spin dephasing is mainly related to the hole spin, whose coherence time can be three orders of magnitude longer than the cavity photon lifetime [96].

The electron-spin initialization, manipulation, and measurement required in our scheme can be efficiently achieved. For example, the spin superposition state can be prepared by optical pumping and/or optical cooling followed by spin-flip Raman transitions [97], or single-spin rotations using nanosecond electron spin resonance (ESR) microwave pulses [91] or picosecond/femtosecond optical pulses [98]. The photon-spin entangling gate based on the ideal GCB can also be used to initialize and measure the spin. The spin manipulation and measurement time scales could be much shorter than the spin coherence time.

In summary, we have presented a complete and faithful two-photon polarization-spatial HBSA scheme, using a failure-heralded and fidelity-robust P-SWAP gate constructed with a double-sided QD-cavity system. The HBSA is implemented with three steps: first, the polarization and spatial-mode states of the photon $B$ are exchanged, by which the hyperentangled Bell state is transformed to the state with the polarization DOF of the photon $A$ ($B$) entangled to the spatial-mode DOF of the photon $B$ ($A$); second, the polarization state of the photon $A$ is exchanged with the polarization state of the photon $B$, which transforms the two-photon state into the product of two single-photon hybrid entangled states; at last, the single-photon hybrid entangled states are measured. Compared with previous schemes, our scheme not only simplifies the state analysis process and saves the quantum resource, but also guarantees the robust fidelity and relaxes the experimental requirement. Therefore, our scheme may be more feasible and useful for practical applications based on the photonic hyperentanglement.

Funding

National Natural Science Foundation of China (61671085, 61701035).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. A. K. Ekert, “Quantum cryptography based on Bell’s theorem,” Phys. Rev. Lett. 67(6), 661–663 (1991). [CrossRef]  

2. G.-L. Long and X.-S. Liu, “Theoretically efficient high-capacity quantum-key-distribution scheme,” Phys. Rev. A 65(3), 032302 (2002). [CrossRef]  

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

4. X.-S. Liu, G.-L. Long, D.-M. Tong, and F. Li, “General scheme for superdense coding between multiparties,” Phys. Rev. A 65(2), 022304 (2002). [CrossRef]  

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

6. M. Hillery, V. Bužek, and A. Berthiaume, “Quantum secret sharing,” Phys. Rev. A 59(3), 1829–1834 (1999). [CrossRef]  

7. 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]  

8. W. Zhang, D.-S. Ding, Y.-B. Sheng, L. Zhou, B.-S. Shi, and G.-C. Guo, “Quantum secure direct communication with quantum memory,” Phys. Rev. Lett. 118(22), 220501 (2017). [CrossRef]  

9. R. Raussendorf and H. J. Briegel, “A one-way quantum computer,” Phys. Rev. Lett. 86(22), 5188–5191 (2001). [CrossRef]  

10. R. Raussendorf, D. E. Browne, and H. J. Briegel, “Measurement-based quantum computation on cluster states,” Phys. Rev. A 68(2), 022312 (2003). [CrossRef]  

11. N. Lütkenhaus, J. Calsamiglia, and K.-A. Suominen, “Bell measurements for teleportation,” Phys. Rev. A 59(5), 3295–3300 (1999). [CrossRef]  

12. S. Ghosh, G. Kar, A. Roy, A. Sen, and U. Sen, “Distinguishability of Bell states,” Phys. Rev. Lett. 87(27), 277902 (2001). [CrossRef]  

13. J. Calsamiglia and N. Lütkenhaus, “Maximum efficiency of a linear-optical Bell-state analyzer,” Appl. Phys. B 72(1), 67–71 (2001). [CrossRef]  

14. J. Calsamiglia, “Generalized measurements by linear elements,” Phys. Rev. A 65(3), 030301 (2002). [CrossRef]  

15. S. D. Barrett, P. Kok, K. Nemoto, R. G. Beausoleil, W. J. Munro, and T. P. Spiller, “Symmetry analyzer for nondestructive Bell-state detection using weak nonlinearities,” Phys. Rev. A 71(6), 060302 (2005). [CrossRef]  

16. P. G. Kwiat, “Hyper-entangled states-12,” J. Mod. Opt. 44(11-12), 2173–2184 (1997). [CrossRef]  

17. 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]  

18. P. G. Kwiat and H. Weinfurter, “Embedded Bell-state analysis,” Phys. Rev. A 58(4), R2623–R2626 (1998). [CrossRef]  

19. S. Walborn, S. Pádua, and C. Monken, “Hyperentanglement-assisted Bell-state analysis,” Phys. Rev. A 68(4), 042313 (2003). [CrossRef]  

20. X.-F. Ren, G.-P. Guo, and G.-C. Guo, “Complete Bell-states analysis using hyper-entanglement,” Phys. Lett. A 343(1-3), 8–11 (2005). [CrossRef]  

21. C. Schuck, G. Huber, C. Kurtsiefer, and H. Weinfurter, “Complete deterministic linear optics Bell state analysis,” Phys. Rev. Lett. 96(19), 190501 (2006). [CrossRef]  

22. M. Barbieri, G. Vallone, P. Mataloni, and F. De Martini, “Complete and deterministic discrimination of polarization Bell states assisted by momentum entanglement,” Phys. Rev. A 75(4), 042317 (2007). [CrossRef]  

23. J. T. Barreiro, T.-C. Wei, and P. G. Kwiat, “Beating the channel capacity limit for linear photonic superdense coding,” Nat. Phys. 4(4), 282–286 (2008). [CrossRef]  

24. B. P. Williams, R. J. Sadlier, and T. S. Humble, “Superdense coding over optical fiber links with complete Bell-state measurements,” Phys. Rev. Lett. 118(5), 050501 (2017). [CrossRef]  

25. 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]  

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

27. X.-H. Li, “Deterministic polarization-entanglement purification using spatial entanglement,” Phys. Rev. A 82(4), 044304 (2010). [CrossRef]  

28. F.-G. Deng, “One-step error correction for multipartite polarization entanglement,” Phys. Rev. A 83(6), 062316 (2011). [CrossRef]  

29. Y.-B. Sheng and L. Zhou, “Deterministic polarization entanglement purification using time-bin entanglement,” Laser Phys. Lett. 11(8), 085203 (2014). [CrossRef]  

30. T.-J. Wang, T. Li, F.-F. Du, and F.-G. Deng, “High-capacity quantum secure direct communication based on quantum hyperdense coding with hyperentanglement,” Chin. Phys. Lett. 28(4), 040305 (2011). [CrossRef]  

31. X.-L. Wang, X.-D. Cai, Z.-E. Su, M.-C. Chen, D. Wu, L. Li, N.-L. Liu, C.-Y. Lu, and J.-W. Pan, “Quantum teleportation of multiple degrees of freedom of a single photon,” Nature 518(7540), 516–519 (2015). [CrossRef]  

32. F.-Z. Wu, G.-J. Yang, H.-B. Wang, J. Xiong, F. Alzahrani, A. Hobiny, and F.-G. Deng, “High-capacity quantum secure direct communication with two-photon six-qubit hyperentangled states,” Sci. China Phys. Mech. Astron. 60(12), 120313 (2017). [CrossRef]  

33. X.-M. Hu, Y. Guo, B.-H. Liu, Y.-F. Huang, C.-F. Li, and G.-C. Guo, “Beating the channel capacity limit for superdense coding with entangled ququarts,” Sci. Adv. 4(7), eaat9304 (2018). [CrossRef]  

34. K. Chen, C.-M. Li, Q. Zhang, Y.-A. Chen, A. Goebel, S. Chen, A. Mair, and J.-W. Pan, “Experimental realization of one-way quantum computing with two-photon four-qubit cluster states,” Phys. Rev. Lett. 99(12), 120503 (2007). [CrossRef]  

35. G. Vallone, E. Pomarico, F. De Martini, and P. Mataloni, “Active one-way quantum computation with two-photon four-qubit cluster states,” Phys. Rev. Lett. 100(16), 160502 (2008). [CrossRef]  

36. T.-J. Wang and C. Wang, “Parallel quantum computing teleportation for spin qubits in quantum dot and microcavity coupled system,” IEEE J. Sel. Top. Quantum Electron. 21(3), 91–97 (2015). [CrossRef]  

37. J. T. Barreiro, N. K. Langford, N. A. Peters, and P. G. Kwiat, “Generation of hyperentangled photon pairs,” Phys. Rev. Lett. 95(26), 260501 (2005). [CrossRef]  

38. D. Bhatti, J. von Zanthier, and G. S. Agarwal, “Entanglement of polarization and orbital angular momentum,” Phys. Rev. A 91(6), 062303 (2015). [CrossRef]  

39. 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]  

40. B.-C. Ren and G.-L. Long, “General hyperentanglement concentration for photon systems assisted by quantum-dot spins inside optical microcavities,” Opt. Express 22(6), 6547–6561 (2014). [CrossRef]  

41. 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]  

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

43. T.-C. Wei, J. T. Barreiro, and P. G. Kwiat, “Hyperentangled Bell-state analysis,” Phys. Rev. A 75(6), 060305 (2007). [CrossRef]  

44. N. Pisenti, C. P. E. Gaebler, and T. W. Lynn, “Distinguishability of hyperentangled Bell states by linear evolution and local projective measurement,” Phys. Rev. A 84(2), 022340 (2011). [CrossRef]  

45. X.-H. Li and S. Ghose, “Hyperentangled Bell-state analysis and hyperdense coding assisted by auxiliary entanglement,” Phys. Rev. A 96(2), 020303 (2017). [CrossRef]  

46. C.-Y. Gao, B.-C. Ren, Y.-X. Zhang, Q. Ai, and F.-G. Deng, “The linear optical unambiguous discrimination of hyperentangled Bell states assisted by time bin,” Ann. of Phys. 531(10), 1900201 (2019). [CrossRef]  

47. 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]  

48. Y. Xia, Q.-Q. Chen, J. Song, and H.-S. Song, “Efficient hyperentangled Greenberger–Horne–Zeilinger states analysis with cross-Kerr nonlinearity,” J. Opt. Soc. Am. B 29(5), 1029–1037 (2012). [CrossRef]  

49. X.-H. Li and S. Ghose, “Self-assisted complete maximally hyperentangled state analysis via the cross-Kerr nonlinearity,” Phys. Rev. A 93(2), 022302 (2016). [CrossRef]  

50. X.-H. Li and S. Ghose, “Complete hyperentangled Bell state analysis for polarization and time-bin hyperentanglement,” Opt. Express 24(16), 18388–18398 (2016). [CrossRef]  

51. G.-Y. Wang, B.-C. Ren, F.-G. Deng, and G.-L. Long, “Complete analysis of hyperentangled Bell states assisted with auxiliary hyperentanglement,” Opt. Express 27(6), 8994–9003 (2019). [CrossRef]  

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

53. Y.-F. Xiao and Q.-H. Gong, “Optical microcavity: from fundamental physics to functional photonics devices,” Sci. Bull. 61(3), 185–186 (2016). [CrossRef]  

54. C. Cao, Y.-W. Duan, X. Chen, R. Zhang, T.-J. Wang, and C. Wang, “Implementation of single-photon quantum routing and decoupling using a nitrogen-vacancy center and a whispering-gallery-mode resonator-waveguide system,” Opt. Express 25(15), 16931–16946 (2017). [CrossRef]  

55. Q.-H. Song, “Emerging opportunities for ultra-high Q whispering gallery mode microcavities,” Sci. China Phys. Mech. Astron. 62(7), 74231 (2019). [CrossRef]  

56. C. Y. Hu, A. Young, J. L. O’Brien, W. J. Munro, and J. G. Rarity, “Giant optical Faraday rotation induced by a single-electron spin in a quantum dot: Applications to entangling remote spins via a single photon,” Phys. Rev. B 78(8), 085307 (2008). [CrossRef]  

57. C. Y. Hu, W. J. Munro, and J. G. Rarity, “Deterministic photon entangler using a charged quantum dot inside a microcavity,” Phys. Rev. B 78(12), 125318 (2008). [CrossRef]  

58. C. Y. Hu, W. J. Munro, J. L. O’Brien, and J. G. Rarity, “Proposed entanglement beam splitter using a quantum-dot spin in a double-sided optical microcavity,” Phys. Rev. B 80(20), 205326 (2009). [CrossRef]  

59. C. Cao, X. Chen, Y.-W. Duan, L. Fan, R. Zhang, T.-J. Wang, and C. Wang, “Concentrating partially entangled W-class states on nonlocal atoms using low-Q optical cavity and linear optical elements,” Sci. China Phys. Mech. Astron. 59(10), 100315 (2016). [CrossRef]  

60. C. Y. Hu, “Spin-based single-photon transistor, dynamic random access memory, diodes, and routers in semiconductors,” Phys. Rev. B 94(24), 245307 (2016). [CrossRef]  

61. C. Y. Hu, “Photonic transistor and router using a single quantum-dot-confined spin in a single-sided optical microcavity,” Sci. Rep. 7(1), 45582 (2017). [CrossRef]  

62. B.-Y. Xia, C. Cao, Y.-H. Han, and R. Zhang, “Universal photonic three-qubit quantum gates with two degrees of freedom assisted by charged quantum dots inside single-sided optical microcavities,” Laser Phys. 28(9), 095201 (2018). [CrossRef]  

63. C. Cao, Y.-H. Han, L. Zhang, L. Fan, Y.-W. Duan, and R. Zhang, “High-fidelity universal quantum controlled gates on electron-spin qubits in quantum dots inside single-sided optical microcavities,” Adv. Quantum Technol. 2(10), 1900081 (2019). [CrossRef]  

64. C. Bonato, F. Haupt, S. S. Oemrawsingh, J. Gudat, D. Ding, M. P. van Exter, and D. Bouwmeester, “CNOT and Bell-state analysis in the weak-coupling cavity QED regime,” Phys. Rev. Lett. 104(16), 160503 (2010). [CrossRef]  

65. C. Y. Hu and J. G. Rarity, “Loss-resistant state teleportation and entanglement swapping using a quantum-dot spin in an optical microcavity,” Phys. Rev. B 83(11), 115303 (2011). [CrossRef]  

66. B.-C. Ren, H.-R. Wei, M. Hua, T. Li, and F.-G. Deng, “Photonic spatial Bell-state analysis for robust quantum secure direct communication using quantum dot-cavity systems,” Eur. Phys. J. D 67(2), 30 (2013). [CrossRef]  

67. B.-C. Ren, H.-R. Wei, M. Hua, T. Li, and F.-G. Deng, “Complete hyperentangled-Bell-state analysis for photon systems assisted by quantum-dot spins in optical microcavities,” Opt. Express 20(22), 24664–24677 (2012). [CrossRef]  

68. T.-J. Wang, Y. Lu, and G.-L. Long, “Generation and complete analysis of the hyperentangled Bell state for photons assisted by quantum-dot spins in optical microcavities,” Phys. Rev. A 86(4), 042337 (2012). [CrossRef]  

69. G.-Y. Wang, Q. Ai, B.-C. Ren, T. Li, and F.-G. Deng, “Error-detected generation and complete analysis of hyperentangled Bell states for photons assisted by quantum-dot spins in double-sided optical microcavities,” Opt. Express 24(25), 28444–28458 (2016). [CrossRef]  

70. Z. Zeng, “Self-assisted complete hyperentangled Bell state analysis using quantum-dot spins in optical microcavities,” Laser Phys. Lett. 15(5), 055204 (2018). [CrossRef]  

71. Y.-Y. Zheng, L.-X. Liang, and M. Zhang, “Error-heralded generation and self-assisted complete analysis of two-photon hyperentangled Bell states through single-sided quantum-dot-cavity systems,” Sci. China Phys. Mech. Astron. 62(7), 970312 (2019). [CrossRef]  

72. R. Warburton, C. Dürr, K. Karrai, J. Kotthaus, G. Medeiros-Ribeiro, and P. Petroff, “Charged excitons in self-assembled semiconductor quantum dots,” Phys. Rev. Lett. 79(26), 5282–5285 (1997). [CrossRef]  

73. G.-Y. Wang, T. Li, Q. Ai, and F.-G. Deng, “Self-error-corrected hyperparallel photonic quantum computation working with both the polarization and the spatial-mode degrees of freedom,” Opt. Express 26(18), 23333–23346 (2018). [CrossRef]  

74. H.-R. Wei, N.-Y. Chen, and J.-Z. Liu, “Heralded universal quantum gate and entangler assisted by imperfect double-sided quantum-dot-microcavity systems,” Ann. of Phys. 530(8), 1800071 (2018). [CrossRef]  

75. J.-Z. Liu, H.-R. Wei, and N.-Y. Chen, “A heralded and error-rejecting three-photon hyper-parallel quantum gate through cavity-assisted interactions,” Sci. Rep. 8(1), 1885 (2018). [CrossRef]  

76. 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]  

77. A.-P. Liu, L.-Y. Cheng, Q. Guo, S.-L. Su, H.-F. Wang, and S. Zhang, “Error-detected single-photon quantum routing using a quantum dot and a double-sided microcavity system,” Chin. Phys. B 28(2), 020301 (2019). [CrossRef]  

78. G.-Y. Wang, T. Li, Q. Ai, A. Alsaedi, T. Hayat, and F.-G. Deng, “Faithful entanglement purification for high-capacity quantum communication with two-photon four-qubit systems,” Phys. Rev. Appl. 10(5), 054058 (2018). [CrossRef]  

79. F.-F. Du, Y.-T. Liu, Z.-R. Shi, Y.-X. Liang, J. Tang, and J. Liu, “Efficient hyperentanglement purification for three-photon systems with the fidelity-robust quantum gates and hyperentanglement link,” Opt. Express 27(19), 27046–27061 (2019). [CrossRef]  

80. A.-P. Liu, L.-Y. Cheng, Q. Guo, S.-L. Su, H.-F. Wang, and S. Zhang, “Heralded entanglement concentration of nonlocal photons assisted by doublesided optical microcavities,” Phys. Scr. 94(9), 095103 (2019). [CrossRef]  

81. J. T. Barreiro, T.-C. Wei, and P. G. Kwiat, “Remote preparation of single-photon "hybrid" entangled and vector-polarization states,” Phys. Rev. Lett. 105(3), 030407 (2010). [CrossRef]  

82. Y.-B. Sheng, L. Zhou, W.-W. Cheng, L.-Y. Gong, L. Wang, and S.-M. Zhao, “Complete Bell-state analysis for a single-photon hybrid entangled state,” Chin. Phys. B 22(3), 030314 (2013). [CrossRef]  

83. J. P. Reithmaier, G. Sęk, A. Löffler, C. Hofmann, S. Kuhn, S. Reitzenstein, L. Keldysh, V. Kulakovskii, T. Reinecke, and A. Forchel, “Strong coupling in a single quantum dot–semiconductor microcavity system,” Nature 432(7014), 197–200 (2004). [CrossRef]  

84. S. Reitzenstein, C. Hofmann, A. Gorbunov, M. Strauß, S. Kwon, C. Schneider, A. Löffler, S. Höfling, M. Kamp, and A. Forchel, “AlAs/GaAs micropillar cavities with quality factors exceeding 150.000,” Appl. Phys. Lett. 90(25), 251109 (2007). [CrossRef]  

85. W. Langbein, P. Borri, U. Woggon, V. Stavarache, D. Reuter, and A. Wieck, “Control of fine-structure splitting and biexciton binding in InxGa1−x as quantum dots by annealing,” Phys. Rev. B 69(16), 161301 (2004). [CrossRef]  

86. R. Seguin, A. Schliwa, S. Rodt, K. Pötschke, U. Pohl, and D. Bimberg, “Size-dependent fine-structure splitting in self-organized InAs/GaAs quantum dots,” Phys. Rev. Lett. 95(25), 257402 (2005). [CrossRef]  

87. R. Stevenson, R. Young, P. See, D. Gevaux, K. Cooper, P. Atkinson, I. Farrer, D. Ritchie, and A. Shields, “Magnetic-field-induced reduction of the exciton polarization splitting in InAs quantum dots,” Phys. Rev. B 73(3), 033306 (2006). [CrossRef]  

88. T. Calarco, A. Datta, P. Fedichev, E. Pazy, and P. Zoller, “Spin-based all-optical quantum computation with quantum dots: Understanding and suppressing decoherence,” Phys. Rev. A 68(1), 012310 (2003). [CrossRef]  

89. G. Bester, S. Nair, and A. Zunger, “Pseudopotential calculation of the excitonic fine structure of million-atom self-assembled In1−xGaxAs/GaAs quantum dots,” Phys. Rev. B 67(16), 161306 (2003). [CrossRef]  

90. J. R. Petta, A. C. Johnson, J. M. Taylor, E. A. Laird, A. Yacoby, M. D. Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard, “Coherent manipulation of coupled electron spins in semiconductor quantum dots,” Science 309(5744), 2180–2184 (2005). [CrossRef]  

91. A. Greilich, D. Yakovlev, A. Shabaev, A. L. Efros, I. Yugova, R. Oulton, V. Stavarache, D. Reuter, A. Wieck, and M. Bayer, “Mode locking of electron spin coherences in singly charged quantum dots,” Science 313(5785), 341–345 (2006). [CrossRef]  

92. J. Elzerman, R. Hanson, L. W. Van Beveren, B. Witkamp, L. Vandersypen, and L. P. Kouwenhoven, “Single-shot read-out of an individual electron spin in a quantum dot,” Nature 430(6998), 431–435 (2004). [CrossRef]  

93. M. Kroutvar, Y. Ducommun, D. Heiss, M. Bichler, D. Schuh, G. Abstreiter, and J. J. Finley, “Optically programmable electron spin memory using semiconductor quantum dots,” Nature 432(7013), 81–84 (2004). [CrossRef]  

94. D. Press, K. De Greve, P. L. McMahon, T. D. Ladd, B. Friess, C. Schneider, M. Kamp, S. Höfling, A. Forchel, and Y. Yamamoto, “Ultrafast optical spin echo in a single quantum dot,” Nat. Photonics 4(6), 367–370 (2010). [CrossRef]  

95. W. Langbein, P. Borri, U. Woggon, V. Stavarache, D. Reuter, and A. Wieck, “Radiatively limited dephasing in InAs quantum dots,” Phys. Rev. B 70(3), 033301 (2004). [CrossRef]  

96. D. Brunner, B. D. Gerardot, P. A. Dalgarno, G. Wüst, K. Karrai, N. G. Stoltz, P. M. Petroff, and R. J. Warburton, “A coherent single-hole spin in a semiconductor,” Science 325(5936), 70–72 (2009). [CrossRef]  

97. M. Atatüre, J. Dreiser, A. Badolato, and A. Imamoglu, “Observation of faraday rotation from a single confined spin,” Nat. Phys. 3(2), 101–106 (2007). [CrossRef]  

98. J. Berezovsky, M. Mikkelsen, N. Stoltz, L. Coldren, and D. Awschalom, “Picosecond coherent optical manipulation of a single electron spin in a quantum dot,” Science 320(5874), 349–352 (2008). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) A singly charged QD in a double-sided optical micropillar cavity. The two distributed Bragg reflectors (DBRs) and the transverse index guiding provide the three-dimensional confinement of light. The two DBRs of the microcavity are made partially reflective (double sided) and symmetric, and the cross section is made circular. (b) The relevant energy levels and the optical spin selection rule for $X^-$ transitions due to the Pauli exclusion principle and the conservation of total spin angular momentum (see text).
Fig. 2.
Fig. 2. Schematic diagram of the error-detected circuit unit constructed with a double-sided QD-cavity system and some linear-optical elements. Here, C is an optical circulator and D is a single-photon detector. BS represents a 50:50 beam splitter, which performs the spatial-mode Hadamard operation [$\left | {{i_1}} \right \rangle \leftrightarrow \left ( {\left | {{j_1}} \right \rangle + \left | {{j_2}} \right \rangle } \right )/\sqrt 2$ and $\left | {{i_2}} \right \rangle \leftrightarrow \left ( {\left | {{j_1}} \right \rangle - \left | {{j_2}} \right \rangle } \right )/\sqrt 2$]. H (H$_{j1}$ or H$_{j2}$) represents a half-wave plate oriented at ${22.5^ \circ }$, which performs the polarization Hadamard operation [$\left | R \right \rangle \leftrightarrow \left ( {\left | R \right \rangle + \left | L \right \rangle } \right )/\sqrt 2$ and $\left | L \right \rangle \leftrightarrow \left ( {\left | R \right \rangle - \left | L \right \rangle } \right )/\sqrt 2$]. M (M$_{j1}$ or M$_{j2}$) is a mirror.
Fig. 3.
Fig. 3. Quantum circuit for the implementation of a failure-heralded and fidelity-robust P-SWAP gate with the error-detected circuit unit. Here, CPBS represents a circularly polarizing beam splitter in the basis $\left \{ {\left | R \right \rangle ,\left | L \right \rangle } \right \}$, which transmits the right-circularly polarized photon component and reflects the left-circularly polarized photon component. X represents a half wave plate oriented at ${45^ \circ }$, which performs the polarization bit-flip operation ${\sigma _x} = \left | R \right \rangle \left \langle L \right | + \left | L \right \rangle \left \langle R \right |$. UBS represents an unbalanced beam splitter with the transmission coefficient $T$. S is an optical switch. Other devices are the same as that in Fig. 2.
Fig. 4.
Fig. 4. Schematic diagram of our complete and faithful HBSA scheme for hyperentangled two-photon systems in the polarization and spatial-mode DOFs.
Fig. 5.
Fig. 5. Calculated success probability of the P-SWAP gate as a function of $g/\kappa$ and $\kappa _s/\kappa$. Here, $\omega = {\omega _c} = {\omega _{{X^ - }}}$ is assumed and $\gamma = 0.1\kappa$ is taken by considering the practical QD-cavity parameters.

Tables (1)

Tables Icon

Table 1. The relationship between the final measurement outcomes of the two photons in the spatial-mode and polarization DOFs and the initial hyperentangled Bell states of the two-photon system. The products of two single-photon hybrid entangled states after the successful P-SWAP operations are also contained in the table, where | ϕ ± A ( B ) p , s = ( | R a 1 ( b 1 ) ± | L a 2 ( b 2 ) ) A ( B ) / 2 and | ψ ± A ( B ) p , s = ( | R a 2 ( b 2 ) ± | L a 1 ( b 1 ) ) A ( B ) / 2 represent the single-photon hybrid entangled states.

Equations (17)

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

d a ^ d t = [ i ( ω c ω ) + κ + κ s 2 ] a ^ g σ ^ κ a ^ i n κ a ^ i n + H ^ , d σ ^ d t = [ i ( ω X ω ) + γ 2 ] σ ^ g σ ^ z a ^ + G ^ , a ^ r = a ^ i n + κ a ^ , a ^ t = a ^ i n + κ a ^ ,
r ( ω ) = 1 + t ( ω ) , t ( ω ) = κ [ i ( ω X ω ) + γ 2 ] [ i ( ω X ω ) + γ 2 ] [ i ( ω c ω ) + κ + κ s 2 ] + g 2 .
r 0 ( ω ) = i ( ω c ω ) + κ s 2 i ( ω c ω ) + κ + κ s 2 , t 0 ( ω ) = κ i ( ω c ω ) + κ + κ s 2 .
| L , r | R , + t | L , , | R , r 0 | L , + t 0 | R , , | R , r | L , + t | R , , | L , r 0 | R , + t 0 | L , , | R , r | L , + t | R , , | L , r 0 | R , + t 0 | L , , | L , r | R , + t | L , , | R , r 0 | L , + t 0 | R , .
| Φ 1 = D | R | i 1 | + + T | L | i 2 | ,
| Ψ 1 = D | R | i 1 | + T | L | i 2 | + .
| φ A = ( α 1 | R + β 1 | L ) A ( χ 1 | a 1 + δ 1 | a 2 ) A , | φ B = ( α 2 | R + β 2 | L ) B ( χ 2 | b 1 + δ 2 | b 2 ) B ,
| Ω 1 = T 2 2 [ α 1 α 2 ( | R R + | R L + | L R + | L L ) A B | + α 1 β 2 ( | R R | R L + | L R | L L ) A B | + β 1 α 2 ( | R R + | R L | L R | L L ) A B | + β 1 β 2 ( | R R | R L | L R + | L L ) A B | ] ( χ 1 | a 1 + δ 1 | a 2 ) A ( χ 2 | b 1 + δ 2 | b 2 ) B .
| Ω 2 = T 4 2 [ ( α 2 | R + β 2 | L ) A ( α 1 | R + β 1 | L ) B | + ( α 2 | R β 2 | L ) A ( α 1 | R β 1 | L ) B | ] ( χ 1 | a 1 + δ 1 | a 2 ) A ( χ 2 | b 1 + δ 2 | b 2 ) B .
| Ω 3 = T 4 ( α 2 | R + β 2 | L ) A ( χ 1 | a 1 + δ 1 | a 2 ) A ( α 1 | R + β 1 | L ) B ( χ 2 | b 1 + δ 2 | b 2 ) B ,
| φ A B p , s = | ε A B p | ξ A B s ,
| ϕ ± A B p = 1 2 ( | R R ± | L L ) A B , | ψ ± A B p = 1 2 ( | R L ± | L R ) A B .
| ϕ ± A B s = 1 2 ( | a 1 b 1 ± | a 2 b 2 ) A B , | ψ ± A B s = 1 2 ( | a 1 b 2 ± | a 2 b 1 ) A B .
| ϕ + p | ϕ + A B s 1 2 ( | R a 1 + | L a 2 ) A 1 2 ( | R b 1 + | L b 2 ) B .
| ϕ + p | ϕ + A B s | a 2 , R A | b 2 , R B , | ϕ p | ϕ + A B s | a 2 , R A | b 2 , L B , | ψ + p | ϕ + A B s | a 2 , R A | b 1 , R B , | ψ p | ϕ + A B s | a 2 , R A | b 1 , L B , | ϕ + p | ϕ A B s | a 2 , L A | b 2 , R B , | ϕ p | ϕ A B s | a 2 , L A | b 2 , L B , | ψ + p | ϕ A B s | a 2 , L A | b 1 , R B , | ψ p | ϕ A B s | a 2 , L A | b 1 , L B , | ϕ + p | ψ + A B s | a 1 , R A | b 2 , R B , | ϕ p | ψ + A B s | a 1 , R A | b 2 , L B , | ψ + p | ψ + A B s | a 1 , R A | b 1 , R B , | ψ p | ψ + A B s | a 1 , R A | b 1 , L B , | ϕ + p | ψ A B s | a 1 , L A | b 2 , R B , | ϕ p | ψ A B s | a 1 , L A | b 2 , L B , | ψ + p | ψ A B s | a 1 , L A | b 1 , R B , | ψ p | ψ A B s | a 1 , L A | b 1 , L B .
| ϕ + A B p | a 1 b 1 1 2 ( | a 2 , R + | a 2 , L ) A | b 2 , R B , | ϕ A B p | a 1 b 1 1 2 ( | a 2 , R + | a 2 , L ) A | b 2 , L B , | ψ + A B p | a 1 b 1 1 2 ( | a 2 , R + | a 2 , L ) A | b 1 , R B , | ψ A B p | a 1 b 1 1 2 ( | a 2 , R + | a 2 , L ) A | b 1 , L B .
η = | T | 8 .
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.