Next Article in Journal
Sparse Estimation Based on a New Random Regularized Matching Pursuit Generalized Approximate Message Passing Algorithm
Next Article in Special Issue
On Two-Distillable Werner States
Previous Article in Journal
Quantum Coherent Three-Terminal Thermoelectrics: Maximum Efficiency at Given Power Output
Previous Article in Special Issue
Many-Body-Localization Transition in the Strong Disorder Limit: Entanglement Entropy from the Statistics of Rare Extensive Resonances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multiphoton Controllable Transport between Remote Resonators

1
School of Physics, Beijing Institute of Technology, Beijing 100081, China
2
Center for Emergent Matter Science (CEMS), The Institute of Physical and Chemical Research (RIKEN), Wako-shi, Saitama 351-0198, Japan
3
State Key Laboratory of Low-Dimensional Quantum Physics and Department of Physics, Tsinghua University, Beijing 100084, China
4
Tsinghua National Laboratory for Information Science and Technology, Tsinghua University, Beijing 100084, China
*
Author to whom correspondence should be addressed.
Entropy 2016, 18(6), 179; https://doi.org/10.3390/e18060179
Submission received: 17 February 2016 / Revised: 4 May 2016 / Accepted: 9 May 2016 / Published: 27 May 2016
(This article belongs to the Special Issue Quantum Information 2016)

Abstract

:
We develop a novel method for multiphoton controllable transport between remote resonators. Specifically, an auxiliary resonator is used to control the coherent long-range coupling of two spatially separated resonators, mediated by a coupled-resonator chain of arbitrary length. In this manner, an arbitrary multiphoton quantum state can be either transmitted through or reflected off the intermediate chain on demand, with very high fidelity. We find, on using a time-independent perturbative treatment, that quantum information leakage of an arbitrary Fock state is limited by two upper bounds, one for the transmitted case and the other for the reflected case. In principle, the two upper bounds can be made arbitrarily small, which is confirmed by numerical simulations.

Graphical Abstract

1. Introduction

The realization of quantum information transport between remote parties is of central importance for any scalable quantum information processing. To this end, one straightforward approach is to use quantum channels to connect these spatially separated parties. Owing to high speed transmission and negligible interaction with the environment, in the form of flying qubits, photons are the natural candidates for long-distance quantum communication [1,2,3,4,5,6,7]. Alternatively, for short-distance quantum communication inside a quantum computer, the majority of promising channels rely upon the use of solid-state-based devices, including nuclear spins in nuclear magnetic resonance (NMR) [8,9], electron spins of nitrogen-vacancy (NV) colour centers in diamond [10,11,12,13,14,15], and flux qubits in superconductors [16,17,18,19]. Moreover, coupled-resonator arrays (CRAs), being currently explored in various physical systems such as superconducting transmission line resonators [20,21,22,23,24], toroidal microresonators [25,26,27,28] and plasmonic nanoparticle arrays [29], have been attracted much attention in recent years. A particular advantage of such arrays is the full addressability of individual resonators, which allows each of the resonators to act as a quantum network node [30,31,32]. Indeed, utilizing these CRAs of having the same fundamental hardware to process quantum information can also avoid a quantum interface between the quantum register and the quantum channel.
In addition to offering an effective platform to simulate quantum many-body phenomena such as Mott-superfluid and topological effects [33,34,35,36], the CRAs have been previously considered for controllable transport of photons by making use of the photon-atom scattering [22,37,38,39,40,41]. Despite such substantial developments, prior work on controllable photon transport has typically focused upon either single photons [22,38,42,43,44] or nearby CRAs [45,46,47,48,49]. However, the ability to transport multiphoton quantum states is a key requirement for encoding a high-dimensional Hilbert space, which is applicable, for example, to universal quantum computation [50,51,52]; at the same time, quantum information stored in the multiphoton fields needs to be controllably transported between two distant quantum registers to carry out quantum network operations. For these reasons, developing a quantum channel capable of controllably transporting multiphoton states is thus of both fundamental and practical importance.
In this paper, we propose and analyze a multiphoton controllable transport protocol, where we use an auxiliary resonator coupled to one resonator of a coupled-resonator chain, which serves as a quantum channel to connect two remote resonators. The physical essence underlying our method is that this auxiliary resonator is employed to control the coherent long-range interaction between the two boundary resonators. To be specific, in the case when the auxiliary resonator is absent, the two boundary resonators could be only coupled to a specific eigenmode of the intermediate chain, within the weak-coupling regime. In this case, the time evolution can swap arbitrary bosonic quantum states of the two boundary resonators, yielding an effective photon transport channel (EPTC). In contrast, when the auxiliary resonator is coupled to the intermediate chain, the specific eigenmode could be split, such that the two boundary resonators are decoupled from the intermediate chain in the large-detuning limit. Photons are therefore reflected back, remaining unchanged. As opposed to prior work, the proposed model is capable of controlling the coherent transport of an arbitrary multiphoton quantum state between two distant resonators over an arbitrarily long range.

2. Physical Model and Effective Hamiltonian

The basic idea of our protocol is schematically illustrated in Figure 1a. To begin, we consider a quantum channel consisting of a coupled-resonator chain of N resonators, with a Hamiltonian
H 0 = κ i = 1 N 1 c i c i + 1 + H . c .
where c i ( c i ) represents the annihilation (creation) operator acting on the resonator i and obeys a boson commutation relation c i , c j = δ i j , and κ is the coupling constant between the nearest-neighbor resonators. Two distant resonators, labelled 0 and N + 1 , are coupled to the ends of the quantum channel, and the interaction Hamiltonian is correspondingly given by
V 1 = g 0 c 0 c 1 + c N + 1 c N + H . c .
where g 0 is the coupling strength between the two boundary resonators and the intermediate chain. In order to control the multiphoton coherent transport, we introduce an auxiliary control resonator, labelled N + 2 , to interact with the m-th resonator of the quantum channel through
V 2 = J 0 c N + 2 c m + H . c .
with a coupling strength J 0 . Here, we have assumed that all resonators have a common frequency ω, and the system is transformed into a frame rotating at ω. Therefore, the total Hamiltonian governing the system is H T = H 0 + V 1 + V 2 .
Upon following an orthogonal transformation [53,54],
f k = 2 N + 1 i = 1 N sin i k π N + 1 c i
the Hamiltonian H 0 is diagonalized to
H 0 = k = 1 N ε k f k f k
where ε k = 2 κ cos k π / N + 1 . Substituting Equation (4) into Equations (2) and (3), V 1 and V 2 are likewise transformed to
V 1 = k = 1 N g k c 0 f k + 1 k 1 c N + 1 f k + H . c .
and
V 2 = k = 1 N J k c N + 2 f k + H . c .
respectively. Here, we have defined
g k = g 0 2 N + 1 sin k π N + 1
and
J k = J 0 2 N + 1 sin m k π N + 1
By restricting our attention to odd N, it yields the existence of a single zero-energy mode in the intermediate chain corresponding to k = z N + 1 / 2 , such that this mode is in resonance with the two boundary resonators as well as with the auxiliary resonator. Under the assumption that g 0 and J 0 are much smaller than κ, off-resonant couplings to the boundary resonators and to the auxiliary resonator can be neglected owing to g z , J z | ε z ± 1 ε z | [55]. As a result, the full evolution dynamics is reduced to an effective model in which only the two boundary resonators, the auxiliary resonator and the zero-energy mode are involved, and accordingly H T is approximated as an effective Hamiltonian
H eff = g z c 0 f z + 1 z 1 c N + 1 f z + H . c . + J z c N + 2 f z + H . c .
which can be used to make a multiphoton quantum switch.

3. Multiphoton Controllable Transport and Fidelities

If the auxiliary resonator is uncoupled to the intermediate chain, J 0 = 0 , the effective Hamiltonian becomes
H eff = g z c 0 f z + 1 z 1 c N + 1 f z + H . c .
In this case, the two boundary resonators are coherently coupled by means of the zero-energy mode. In the Heisenberg picture, a straightforward calculation yields
c 0 t = c 0 + 1 2 1 + cos 2 g z t c 0 + 1 z 1 c N + 1 + i sin 2 g z t 2 f z
Choosing the evolution time t = τ π / 2 g z gives
c 0 τ = 1 z c N + 1
and in a similar manner, we have
c N + 1 τ = 1 z c 0
Equations (12) and (13) exhibit that the evolution for a specific time behaves as a swap operation between the two boundary resonators, as shown in Figure 1b. However, this zero-energy mode could, in the case when the auxiliary resonator is coupled to the intermediate channel, be split into two new modes separated by an energy gap 2 J z . It follows, on ensuring g 0 J 0 , that the two boundary resonators are significantly detuned from the new modes if m is odd, and thus are decoupled from the intermediate channel. In this case, the time evolution is referred to as an identity operation, leading to
c 0 t = c 0
and
c N + 1 t = c N + 1
Let us consider the controllable transport of an arbitrary Fock state | n , for example, from the left resonator to the right resonator. The more general treatment of an arbitrary superposition of Fock states are presented in the Appendix. Such a Fock state can be generated using a nonlinear quantum system as an intermediary between a classical radiation field and the resonator [56]. We start with an initial state of the total system,
| Φ 0 = | n 0 | 0 | 0 N + 1 = c 0 n n ! | 0 T
where | 0 = | 0 1 | 0 N | 0 N + 2 , and | 0 T = | 0 0 | 0 | 0 N + 1 is the vacuum state of all resonators. Under the time evolution, the total system freely evolves into a finial state,
| Φ t = c 0 t n n ! | 0 T
In the uncoupled case when J 0 = 0 , according to Equation (12), the finial state for time τ is | Φ τ = 1 n z | 0 0 | 0 | n N + 1 , which means that all photons are simultaneously transported from the left resonator to the right resonator as desired. In contrast, in the coupled case of g 0 J 0 κ , the finial state becomes | Φ t = | Φ 0 according to Equation (14), thus these photons are reflected back, remaining unchanged.
In order to characterize the quality of our protocol, we employ two fidelities with one transmission fidelity, F t = n | ρ N + 1 τ | n , and one reflection fidelity, F r = n | ρ 0 τ | n . Here, ρ 0 τ and ρ N + 1 τ are the reduced density matrices of the resonators 0 and N + 1 , respectively, at time t = τ . Together with an N + 1 × N + 1 coupling matrix A, the total Hamiltonian can be compactly expressed as
H T = i , j = 0 N + 2 A i j c i c j
where A i j is the coupling strength between two resonators i and j. Applying the Heisenberg equation of motion, c i t = i H 0 , c i t , one finds that
c i t = j = 0 N + 2 M i j c j
with M = exp i A t being a unitary evolution matrix. To calculate the two fidelities F t and F r , more explicitly, we rewrite c i t of Equation (19) as
c 0 t = M 0 , μ c μ + δ μ c δ μ
where δ μ = 1 | M 0 , μ | 2 and μ = 0 , , N + 2 . Here, we define a collective mode c δ μ as a normalized linear combination of all modes apart from c μ , resulting in c μ , c δ μ = 0 and c δ μ , c δ μ = 1 . In combination with Equations (17) and (20), we have
| Φ t = r = 0 n f μ r , n | r δ μ | n r μ
where
f μ r , n = C n r M 0 , μ n r δ μ r / 2
C n r = n ! r ! n r !
and | r δ μ is a Fock state of the collective mode c δ μ . The corresponding density matrix is
ρ t = | Φ t Φ t | = r , r = 0 n f μ r , n f μ * r , n | n r μ n r | | r δ μ r |
and then, by tracing out the variables of the collective mode, the reduced density matrix of the resonator μ is calculated as
ρ μ t = r = 0 n | f μ r , n | 2 | n r μ n r |
Thus, the transmission fidelity and the reflection fidelity are respectively given by
F t = | f N + 1 0 , n | 2 = | M 0 , N + 1 | 2 n
and
F r = | f 0 0 , n | 2 = | M 0 , 0 | 2 n
In Figure 2 we plot the transmission fidelity F t versus J 0 / κ with g 0 / κ = 0 . 01 (see Figure 2a), or 0 . 005 (see Figure 2b), as well as the reflection fidelity F r . In the two cases, we assume that a Fock state | n is initialized into the left boundary resonator, and further take n = 2 , 3 and 5 as three special examples to simulate numerically. It is shown that as J 0 = 0 , we obtain F t 1 while F r 0 , implying that photons are transmitted from the left resonator to the right resonator. In contrast, in the regime g 0 J 0 κ , photons are reflected back to the left resonator, according to F t 0 whereas F r 1 . Hence, a Fock state could, by tuning the coupling strength J 0 , be transmitted forward or reflected backward at will.

4. Quantum Information Leakage

Having explicitly demonstrated the controllable eigenmode-mediated transport of multiphoton information, we now calculate the quantum information leakage by making use of perturbation theory. We begin by considering the evolution matrix of Equation (19). In fact, the coupling matrix A is identical to the total Hamiltonian confined in a single-excitation subspace,
H S = g 0 | 0 1 | + | N + 1 N | + H . c . + κ j = 1 N 1 | j j + 1 | + H . c . + J 0 | N + 2 m | + H . c .
where | μ = c μ | 0 T ( μ = 0 , , N + 2 ) represents the μ-th basis of this single-excitation subspace. As a consequence, the evolution matrix M is identical to U t = e i H S t , so that M μ , μ = μ | U t | μ . By introducing an orthogonal transformation,
| k = 2 N + 1 j = 1 N sin j k π N + 1 | j
the Hamiltonian H S is transformed to
H S = k = 1 N g k | 0 k | + 1 k 1 | N + 1 k | + H . c . + k = 1 N ε k | k k | + k = 1 N J k | N + 2 k | + H . c .
Based upon such a Hamiltonian, we can obtain the leakage of quantum information for the uncoupled case of J 0 = 0 , and the coupled case of g 0 J 0 κ , respectively, as we will show in the following.
Let us first focus upon the former case, where the total system can be thought of as an EPTC being perturbatively coupled to a fictitious bosonic environment in the limit g 0 κ . The corresponding Hamiltonian could be divided into three parts, H S = H E + H z + V z . H E features the Hamiltonian for the EPTC:
H E = g z | e 1 e 2 | + 1 z 1 | e 3 e 2 | + H . c .
where, for convenient, we have used | e 1 , | e 2 , | e 3 to replace | 0 , | z , | N + 1 . The environment is determined by
H z = k z ε k | k k |
of having 2 N bosonic modes, and ε k = ε k . The part V z modelling the interaction between them is
V z = k z g k | e 1 k | + 1 k 1 | e 3 k | + H . c .
Assuming that the EPTC Hamiltonian can be diagonalized through a unitary transformation T, it results in i , j = 1 3 T q , i T j , q e i | H E | e j = λ q δ q , q , such that
H E = q = 1 3 λ q | q q |
with | q = i = 1 3 T i q | e i . Similarly,
V z = k z q = 1 3 G k q | q k | + H . c .
where
G k q = g k T q , 1 + 1 k 1 T q , 3
By performing a first-order perturbative treatment, V z has no effects on the eigenenergies λ q and ε k ; nevertheless, the eigenstates | q and | k are modified by
| q ˜ | q k z G k q * ε k | k
and
| k ˜ | k + q = 1 3 G k q ε k | q
respectively. Combining Equations (37) and (38), we find, up to second order,
| q 1 k z | G k q | 2 ε k 2 | q ˜ + k z G k q * ε k | k ˜ k z q q G k q * G k q ε k 2 | q
and hence, after an iteration, | q takes the form of
| q | q ˜ + k z G k q * ε k | k ˜ k z q = 1 3 G k q * G k q ε k 2 | q ˜
Since
q ˜ | q ˜ = δ q , q + k z G k q * G k q ε k 2 k ˜ | k ˜ = δ k k + q G k q * G k q ε k ε k q ˜ | k ˜ = 0
the matrix elements of U t in the energy space can be evaluated as
q | U t | q e i λ q t δ q , q k z G k q * G k q ε k 2 e i λ q t + k z G k q G k q * ε k 2 e i ε k t
Subsequently, after inversion back to the space spanned by | e i ( i = 1 , 2 , 3 ), we arrive at
e 1 | U t | e 3 e 1 | e i H E t | e 3 + k z 1 k 1 g k 2 ε k 2 e i ε k t k z g k 2 ε k 2 1 k 1 e 1 | e i H E t | e 1 + e 1 | e i H E t | e 3
Combining the evolution under H E leads to e 1 | e i H E τ | e 3 = 1 z and e 1 | e i H E τ | e 1 = 0 , and further we find
M 0 , N + 1 1 z 1 2 Δ t
for the evolution time τ. Here, we have defined
Δ t = k < z g k 2 ε k 2 1 1 k + z 1 cos ε k τ
The transmission fidelity is therefore modified by
F t ( 1 2 Δ t ) 2 n 1 4 n Δ t
The leakage of quantum information can be quantified using a transmission infidelity, σ t = 1 F t 4 n Δ t , and thus has an upper bound,
σ t 8 n k < z g k 2 ε k 2
We now consider the coupled case. In this case, the zero-energy mode of the intermediate chain is the only state coupled to the auxiliary resonator due to J 0 κ , which induces two new modes as previously mentioned. The two boundary resonators are coupled to such new modes in addition to the fictitious environment; however, this coupling to the environment can be neglected so long as g 0 J 0 . The Hamiltonian is therefore reduced to
H S = g z | e 1 e 2 | + 1 z 1 | e 3 e 2 | + H . c . + J z | e 4 e 2 | + H . c .
Here, we have used | e 1 , | e 2 , | e 3 , | e 4 to replace | 0 , | z , | N + 1 , | N + 2 for convenient. Using | γ 1 = | e 2 + | e 4 / 2 , | γ 2 = | e 2 | e 4 / 2 and | ± = | e 1 ± | e 3 / 2 , the Hamiltonian H S of Equation (48) is brought to
H S = g z 2 k = 1 , 2 | + + | γ k | + 1 z 1 | + | γ k | + H . c . + k = 1 , 2 J k | γ k γ k |
with J 1 = J 2 = J z . When 1 z 1 = 1 , we get
H S = g z k = 1 , 2 | + γ k | + H . c . + k = 1 , 2 J k | γ k γ k |
from which we can follow the same recipe as described in the uncoupled case to obtain
M 0 , 0 1 2 Δ r
where
Δ r = g z 2 2 J z 2 1 cos J z t
For 1 z 1 = 1 , it has the same result as mentioned in Equation (51). In direct analogy to the uncoupled case, the reflection fidelity is modified by
F r 1 2 Δ r 2 n 1 4 n Δ r
Correspondingly, the reflection infidelity is σ r = 1 F r 4 n Δ r , together with an upper bound,
σ r 4 n g z 2 J z 2
To confirm our calculation of quantum information leaking into the off-resonant couplings [57], we compare numerical results of the transmission infidelity to the analytical upper bound as depicted in Figure 3, as well as the reflection infidelity in Figure 4. It is found that this upper bound is in excellent agreement with the numerical results. In addition, the leakage of quantum information decreases with deceasing g 0 / κ , so that this leakage can, in principle, be made arbitrarily small.

5. Conclusions

In this paper, we have proposed a new approach for multiphoton controllable transport between two remote resonators being coupled to the ends of a coupled-resonator chain of arbitrary length. This manner essentially enables a coherent long-range interaction between the two spatially separated resonators, in this case, the pure Hamiltonian evolution for a specific time is referred to as a swap operation of the two boundary resonators. As a result, an arbitrary multiphoton quantum state can be transported with quantum information leakage arbitrarily close to zero. However, if an auxiliary resonator is harnessed to coupled to one resonator of the intermediate chain, this coherent long-range interaction will be eliminated, so the two boundary resonators are decoupled from the intermediate chain, yielding that the time evolution functions as an identity operation. Thus, an arbitrary multiphoton quantum state can be reflected back with quantum information leakage also arbitrarily close to zero. Our approach potentially allows for realizing controllable transport of an arbitrary dimensional quantum state or even a coherent state (see the Appendix), and can also be directly generalized to quantum networks consisting of at least three quantum registers. In fact, although we have discussed the specific case of a coupled-resonator chain, such a description is consistent for quantum coupled spin systems of having been widely studied [53,54,55,58,59,60,61,62]. Hence, the present approach could be applied to the realization of scalable quantum devices, for example, quantum switches.

Acknowledgments

This work was supported by the National Natural Science Foundation of China under Grant No. 11175094 and 91221205, the National Basic Research Program of China under Grant No. 2015CB921002, and the Basic Research Fund of the Beijing Institute of Technology under Grant No. 20141842005.

Author Contributions

Wei Qin completed the calculation and prepared the figures. Wei Qin and Guilu Long wrote the main manuscript text. Guilu Long supervised the whole project. Both authors have read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix

In this appendix, we will discuss the controllable transport of an arbitrary d-dimensional multiphoton state | ψ , which is a linear superposition of Fock states, | ψ = n = 0 d 1 α n | n . When preparing this state in the left resonator, the initial state of the total system then is
| Φ 0 = | ψ | 0 | 0 N + 1 = n = 0 d 1 α n c 0 n n ! | 0 T
Under the total Hamiltonian H T , according to Equation (21), the time evolution results in
| Φ t = n = 0 d 1 r = 0 n α n f μ r , n | r δ μ | n r μ
so the corresponding density matrix could be expressed as
ρ t = n , n = 0 d 1 r = 0 n r = 0 n α n α n * f μ r , n f μ * r , n | n r μ n r | | r δ μ r |
Consequently, by tracing out the variables of the collective mode, the reduced density matrix of the resonator μ is evaluated as
ρ μ t = n , n = 0 d 1 r = 0 min { n , n } α n α n * f μ r , n f μ * r , n | n r μ n r |
After a straightforward computation, the transmission and reflection fidelities are given, respectively, by
F t ψ | P ρ N + 1 τ P | ψ = n , n = 0 d 1 r = 0 min { n , n } 1 n + n z α n α n r α n * α n r * f N + 1 r , n f N + 1 * r , n
along with P = exp { i z π c N + 1 c N + 1 } to ensure the right phase evolution, and
F r ψ | ρ 0 τ | ψ = n , n = 0 d 1 r = 0 min { n , n } α n α n r α n * α n r * f 0 r , n f 0 * r , n
In order to quantify information transport more precisely, we need to calculate the average transmission and reflection fidelities over all initial pure states. Characterizing a d-dimensional pure state by means of the Hurwitz-parametrization method [63,64], then yields
F t = 1 d d + 1 n = 0 d 1 r = 0 n | f N + 1 r , n | 2 + n , n = 0 d 1 1 n + n z f N + 1 0 , n f N + 1 * 0 , n
and
F r = 1 d d + 1 n = 0 d 1 r = 0 n | f 0 r , n | 2 + n , n = 0 d 1 f 0 0 , n f 0 * 0 , n
In Figure A1 we plot the average transmission fidelity and the average reflection fidelity. It is found that they exhibit similar behaviors to what we have observed in Figure 2, except that, in the limit J 0 g 0 , the average transmission fidelity is close to 1 / d instead. This difference arises from the fact that the vacuum component of the superposition state | ψ has an average population probability of 1 / d after we average over all initial pure states. Hence, controllable transport can be achieved not only for the Fock states but also for arbitrary multiphoton quantum states.
We now consider the quantum information leakage. Specifically, we use an average transmission infidelity σ t = 1 F t for the uncoupled case, and an average reflection infidelity σ r = 1 F r for the coupled case, to quantify the quantum information leakage. It follows, on combining Equations (22), (45) and (A7), that the average transmission infidelity σ t is
σ t = 2 d d 1 d + 1 Δ t
thereby having an upper bound,
σ t 4 d d 1 d + 1 k < z g k 2 ε k 2
In a similar way, the average reflection infidelity σ r is
σ r = 2 d d 1 d + 1 Δ r
and also has an upper bound
σ r 2 d d 1 d + 1 g z 2 J z 2
As described in the main text, we compare numerical results of the average transmission and reflection infidelities to their analytical upper bounds as shown in Figure A2 and Figure A3, respectively. We find, as expected, that the quantum information leakage for controllable transport of arbitrary multiphoton states is limited by two upper bounds which can be made arbitrary small, in direct analogy to the cases of Fock states in Figure 3 and Figure 4.
Figure A1. (color online) Average transmission fidelity F t and average reflection fidelity F r as a function of J 0 / κ with either (a) g 0 / κ = 0 . 01 or (b) 0 . 005 for d = 2 , 3 and 5, corresponding to solid black lines, dashed red lines and dot-dashed blue lines, respectively. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Figure A1. (color online) Average transmission fidelity F t and average reflection fidelity F r as a function of J 0 / κ with either (a) g 0 / κ = 0 . 01 or (b) 0 . 005 for d = 2 , 3 and 5, corresponding to solid black lines, dashed red lines and dot-dashed blue lines, respectively. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Entropy 18 00179 g005
Figure A2. (color online) Numerical results of the average transmission infidelity, σ t = 1 F t , as a function of g 0 / κ with either (a) d = 3 or (b) d = 5 in the uncoupled case of J 0 = 0 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Figure A2. (color online) Numerical results of the average transmission infidelity, σ t = 1 F t , as a function of g 0 / κ with either (a) d = 3 or (b) d = 5 in the uncoupled case of J 0 = 0 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Entropy 18 00179 g006
Figure A3. (color online) Numerical results of the average reflection infidelity, σ r = 1 F r , as a function of g 0 / κ with either (a) d = 3 or (b) d = 5 in the coupled case of J 0 / κ = 0 . 1 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Figure A3. (color online) Numerical results of the average reflection infidelity, σ r = 1 F r , as a function of g 0 / κ with either (a) d = 3 or (b) d = 5 in the coupled case of J 0 / κ = 0 . 1 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Entropy 18 00179 g007

References and Notes

  1. Bouwmeester, D.; Pan, J.W.; Mattle, K.; Eibl, M.; Weinfurter, H.; Zeilinger, A. Experimental quantum teleportation. Nature 1997, 390, 575–579. [Google Scholar] [CrossRef]
  2. Naik, D.S.; Peterson, C.G.; White, A.G.; Berglund, A.J.; Kwiat, P.G. Entangled state quantum cryptography: Eavesdropping on the ekert protocol. Phys. Rev. Lett. 2000, 84, 4733–4736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Tittel, W.; Brendel, J.; Zbinden, H.; Gisin, N. Quantum cryptography using entangled photons in energy-time bell states. Phys. Rev. Lett. 2000, 84, 4737–4740. [Google Scholar] [CrossRef] [PubMed]
  4. Duan, L.M.; Lukin, M.D.; Cirac, J.I.; Zoller, P. Long-distance quantum communication with atomic ensembles and linear optics. Nature 2001, 414, 413–418. [Google Scholar] [CrossRef] [PubMed]
  5. Deng, F.G.; Long, G.L.; Liu, X.S. Two-step quantum direct communication protocol using the Einstein-Podolsky-Rosen pair block. Phys. Rev. A 2003, 68. [Google Scholar] [CrossRef]
  6. Salemian, S.; Mohammadnejad, S. An error-free protocol for quantum entanglement distribution in long-distance quantum communication. Chin. Sci. Bull. 2011, 56, 618–625. [Google Scholar] [CrossRef]
  7. Zheng, C.; Long, G.F. Quantum secure direct dialogue using Einstein-Podolsky-Rosen pairs. Sci. China Phys. Mech. Astron. 2014, 57, 1238–1243. [Google Scholar] [CrossRef]
  8. Zhang, J.F.; Long, G.L.; Zhang, W.; Deng, Z.W.; Liu, W.Z.; Lu, Z.H. Simulation of Heisenberg XY interactions and realization of a perfect state transfer in spin chains using liquid nuclear magnetic resonance. Phys. Rev. A 2005, 72. [Google Scholar] [CrossRef]
  9. Feng, G.R.; Xu, G.F.; Long, G.L. Experimental realization of nonadiabatic holonomic quantum computation. Phys. Rev. Lett. 2013, 110. [Google Scholar] [CrossRef] [PubMed]
  10. Childress, L.; Dutt, M.V.G.; Taylor, J.M.; Zibrov, A.S.; Jelezko, F.; Wrachtrup, J.; Hemmer, P.R.; Lukin, M.D. Coherent dynamics of coupled electron and nuclear spin qubits in diamond. Science 2006, 314, 281–285. [Google Scholar] [CrossRef] [PubMed]
  11. Zagoskin, A.M.; Johansson, J.R.; Ashhab, S.; Nori, F. Quantum information processing using frequency control of impurity spins in diamond. Phys. Rev. B 2007, 76. [Google Scholar] [CrossRef]
  12. Bermudez, A.; Jelezko, F.; Plenio, M.B.; Retzker, A. Electron-mediated nuclear-spin interactions between distant nitrogen-vacancy centers. Phys. Rev. Lett. 2011, 107. [Google Scholar] [CrossRef] [PubMed]
  13. Yao, N.Y.; Jiang, L.; Gorshkov, A.V.; Maurer, P.C.; Giedke, G.; Cirac, J.I.; Lukin, M.D. Scalable architecture for a room temperature solid-state quantum information processor. Nat. Commun. 2012, 3. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, P.F.; Ju, C.Y.; Shi, F.Z.; Du, J.F. Optimizing ultrasensitive single electron magnetometer based on nitrogen-vacancy center in diamond. Chin. Sci. Bull. 2013, 58, 2920–2923. [Google Scholar] [CrossRef]
  15. Doherty, M.W.; Manson, N.B.; Delaney, P.; Jelezko, F.; Wrachtrup, J.; Hollenberg, L.C.L. The nitrogen-vacancy colour centre in diamond. Phys. Rep. 2013, 528, 1–45. [Google Scholar] [CrossRef]
  16. Liu, Y.X.; Wei, L.F.; Tsai, J.S.; Nori, F. Controllable coupling between flux qubits. Phys. Rev. Lett. 2006, 96. [Google Scholar] [CrossRef] [PubMed]
  17. Niskanen, A.O.; Harrabi, K.; Yoshihara, F.; Nakamura, Y.; Lloyd, S.; Tsai, J.S. Quantum coherent tunable coupling of superconducting qubits. Science 2007, 316, 723–726. [Google Scholar] [CrossRef] [PubMed]
  18. Ashhab, S.; Niskanen, A.O.; Harrabi, K.; Nakamura, Y.; Picot, T.; De Groot, P.C.; Harmans, C.J.P.M.; Mooij, J.E.; Nori, F. Interqubit coupling mediated by a high-excitation-energy quantum object. Phys. Rev. B 2008, 77. [Google Scholar] [CrossRef]
  19. Xiong, W.; Jin, D.Y.; Jing, J.; Lam, C.H.; You, J.Q. Controllable coupling between a nanomechanical resonator and a coplanar-waveguide resonator via a superconducting flux qubit. Phys. Rev. A 2015, 92. [Google Scholar] [CrossRef]
  20. Sillanpää, M.A.; Park, J.I.; Simmonds, R.W. Coherent quantum state storage and transfer between two phase qubits via a resonant cavity. Nature 2007, 449, 438–442. [Google Scholar] [CrossRef] [PubMed]
  21. Majer, J.; Chow, J.M.; Gambetta, J.M.; Koch, J.; Johnson, B.R.; Schreier, J.A.; Frunzio, L.; Schuster, D.I.; Houck, A.A.; Wallraff, A.; et al. Coupling superconducting qubits via a cavity bus. Nature 2007, 449, 443–447. [Google Scholar] [CrossRef] [PubMed]
  22. Zhou, L.; Gong, Z.R.; Liu, Y.X.; Sun, C.P.; Nori, F. Controllable scattering of a single photon inside a one-dimensional resonator waveguide. Phys. Rev. Lett. 2008, 101. [Google Scholar] [CrossRef] [PubMed]
  23. Nataf, P.; Ciuti, C. Protected quantum computation with multiple resonators in ultrastrong coupling circuit QED. Phys. Rev. Lett. 2011, 107. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, C.P.; Su, Q.P.; Nori, F. Entanglement generation and quantum information transfer between spatially-separated qubits in different cavities. New J. Phys. 2013, 15. [Google Scholar] [CrossRef]
  25. Armani, D.K.; Kippenberg, T.J.; Spillane, S.M.; Vahala, K.J. Ultra-high-Q toroid microcavity on a chip. Natrue 2003, 421, 925–928. [Google Scholar] [CrossRef] [PubMed]
  26. Peng, B.; Özdemir, Ş.K.; Lei, F.; Monifi, F.; Gianfreda, M.; Long, G.L.; Fan, S.; Nori, F.; Bender, C.M.; Yang, L. Parity-time-symmetric whispering-gallery microcavities. Nat. Phys. 2014, 10, 394–398. [Google Scholar] [CrossRef]
  27. Peng, B.; Özdemir, Ş.K.; Rotter, S.; Yilmaz, H.; Liertzer, M.; Monifi, F.; Bender, C.M.; Nori, F.; Yang, L. Loss-induced suppression and revival of lasing. Science 2014, 346, 328–332. [Google Scholar] [CrossRef] [PubMed]
  28. Jing, H.; Özdemir, Ş.K.; Lu, X.Y.; Zhang, J.; Yang, L.; Nori, F. PT-Symmetric phonon laser. Phys. Rev. Lett. 2014, 113. [Google Scholar] [CrossRef] [PubMed]
  29. Lee, C.; Tame, M.; Noh, C.; Lim, J.; Maier, S.A.; Lee, J.; Angelakis, D.G. Robust-to-loss entanglement generation using a quantum plasmonic nanoparticle array. New J. Phys. 2013, 15, 083017. [Google Scholar] [CrossRef]
  30. Cirac, J.I.; Zoller, P.; Kimble, H.J.; Mabuchi, H. Quantum state transfer and entanglement distribution among distant nodes in a quantum network. Phys. Rev. Lett. 1997, 78, 3221–3224. [Google Scholar] [CrossRef]
  31. Serafini, A.; Mancini, S.; Bose, S. Distributed quantum computation via optical fibers. Phys. Rev. Lett. 2006, 96. [Google Scholar] [CrossRef] [PubMed]
  32. Kimble, H.J. The quantum internet. Nature 2008, 453, 1023–1030. [Google Scholar] [CrossRef] [PubMed]
  33. Hartmann, M.J.; Brandão, F.G.S.L.; Plenio, M.B. Quantum many-body phenomena in coupled cavity arrays. Laser Photonics Rev. 2008, 2, 527–556. [Google Scholar] [CrossRef]
  34. Hafezi, M.; Demler, E.A.; Lukin, M.D.; Taylor, J.M. Robust optical delay lines with topological protection. Nat. Phys. 2011, 7, 907–912. [Google Scholar] [CrossRef]
  35. Georgescu, I.M.; Ashhab, S.; Nori, F. Quantum simulation. Rev. Mod. Phys. 2014, 86, 153–185. [Google Scholar] [CrossRef]
  36. Douglas, J.S.; Habibian, H.; Hung, C.-L.; Gorshkov, A.V.; Kimble, H.J.; Chang, D.E. Quantum many-body models with cold atoms coupled to photonic crystals. Nat. Photonics 2015, 9, 326–331. [Google Scholar] [CrossRef]
  37. Wang, Z.H.; Li, Y.; Zhou, D.L.; Sun, C.P.; Zhang, P. Single-photon scattering on a strongly dressed atom. Phys. Rev. A 2012, 86. [Google Scholar] [CrossRef]
  38. Zhou, L.; Yang, L.P.; Li, Y.; Sun, C.P. Quantum routing of single photons with a cyclic three-level system. Phys. Rev. Lett. 2013, 111. [Google Scholar] [CrossRef] [PubMed]
  39. Lombardo, F.; Ciccarello, F.; Palma, G.M. Photon localization versus population trapping in a coupled-cavity array. Phys. Rev. A 2014, 89. [Google Scholar] [CrossRef] [Green Version]
  40. Hai, L.; Tan, L.; Feng, J.S.; Xu, W.B.; Wang, B. Single photon transport properties in coupled cavity arrays nonlocally coupled to a two-level atom in the presence of dissipation. Chin. Phys. B 2014, 23, 024202. [Google Scholar] [CrossRef]
  41. Lu, J.; Wang, Z.H.; Zhou, L. T-shaped single-photon router. Opt. Express 2015, 23, 22955–22962. [Google Scholar] [CrossRef] [PubMed]
  42. Xu, D.Z.; Ian, H.; Shi, T.; Dong, H.; Sun, C.P. Photonic Feshbach resonance. Sci. China Phys. Mech. Astron. 2010, 53, 1234–1238. [Google Scholar] [CrossRef]
  43. Zhu, W.; Wang, Z.H.; Zhou, D.L. Multimode effects in cavity QED based on a one-dimensional cavity array. Phys. Rev. A 2014, 90. [Google Scholar] [CrossRef]
  44. Qin, W.; Nori, F. Controllable single-photon transport between remote coupled-cavity arrays. Phys. Rev. A 2016, 93. [Google Scholar] [CrossRef]
  45. Longo, P.; Schmitteckert, P.; Busch, K. Few-photon transport in low-dimensional systems: interaction-induced radiation trapping. Phys. Rev. Lett. 2010, 104. [Google Scholar] [CrossRef] [PubMed]
  46. Longo, P.; Schmitteckert, P.; Busch, K. Few-photon transport in low-dimensional systems. Phys. Rev. A 2011, 83. [Google Scholar] [CrossRef]
  47. Roy, D. Few-photon optical diode. Phys. Rev. B 2010, 81. [Google Scholar] [CrossRef]
  48. Roy, D. Two-Photon scattering by a driven three-level emitter in a one-dimensional waveguide and electromagnetically induced transparency. Phys. Rev. Lett. 2011, 106. [Google Scholar] [CrossRef] [PubMed]
  49. Shi, T.; Chang, D.E.; Cirac, J.I. Multiphoton-scattering theory and generalized master equations. Phys. Rev. A 2015, 92. [Google Scholar] [CrossRef]
  50. Luo, M.X.; Wang, X.J. Universal quantum computation with qudits. Sci. China Phys. Mech. Astron. 2014, 57, 1712–1717. [Google Scholar] [CrossRef]
  51. Qin, W.; Wang, C.; Cao, Y.; Long, G.L. Multiphoton quantum communication in quantum networks. Phys. Rev. A 2014, 89. [Google Scholar] [CrossRef]
  52. Hua, M.; Tao, M.J.; Deng, F.G. Efficient generation of NOON states on two microwave-photon resonators. Chin. Sci. Bull. 2014, 59, 2829–2834. [Google Scholar] [CrossRef]
  53. Lieb, E.; Schultz, T.; Mattis, D. Two soluble models of an antiferromagnetic chain. Ann. Phys. 1961, 16, 407–466. [Google Scholar] [CrossRef]
  54. Christandl, M.; Datta, N.; Ekert, A.; Landahl, A.J. Perfect state transfer in quantum spin networks. Phys. Rev. Lett. 2004, 92. [Google Scholar] [CrossRef] [PubMed]
  55. Yao, N.Y.; Jiang, L.; Gorshkov, A.V.; Gong, Z.-X.; Zhai, A.; Duan, L.-M.; Lukin, M.D. Robust quantum state transfer in random unpolarized spin chains. Phys. Rev. Lett. 2011, 106. [Google Scholar] [CrossRef] [PubMed]
  56. Hofheinz, M.; Weig, E.M.; Ansmann, M.; Bialczak, R.C.; Lucero, E.; Neeley, M.; O’Connell, A.D.; Wang, H.; Martinis, J.M.; Cleland, A.N. Generation of Fock states in a superconducting quantum circuit. Nature 2008, 454, 310–314. [Google Scholar] [CrossRef] [PubMed]
  57. In order to plot the transmission and reflection infidelities, we numerically perform the exact diagonalization of the coupling matrix A, and then calculate these infidelities according to Equations (26) and (27), and the definitions of such infidelities. The upper bounds, however, are straightforwardly plotted according to Equations (47) and (54). In addition, we plot these quantities at a specific evolution time, t = τ π / 2 g z , to make the quantum state of either the transmitted or reflected photons remain unchanged
  58. Bose, S. Quantum communication through an unmodulated spin chain. Phys. Rev. Lett. 2003, 91. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, Y.; Zhou, D.L. Optimized quantum state transfer through an XY spin chain. Phys. Rev. A 2014, 89. [Google Scholar] [CrossRef]
  60. Qin, W.; Li, J.L.; Long, G.L. High-dimensional quantum state transfer in a noisy network environment. Chin. Phys. B 2015, 24, 040305. [Google Scholar] [CrossRef]
  61. Qin, W.; Wang, C.; Zhang, X.D. Protected quantum-state transfer in decoherence-free subspaces. Phys. Rev. A 2015, 91. [Google Scholar] [CrossRef]
  62. Fukuhara, T.; Hild, S.; Zeiher, J.; Schauß, P.; Bloch, I.; Endres, M.; Gross, C. Spatially resolved detection of a spin-entanglement wave in a bose-hubbard chain. Phys. Rev. Lett. 2015, 115. [Google Scholar] [CrossRef] [PubMed]
  63. Życzkowski, K.; Sommers, H. Induced measures in the space of mixed quantum states. J. Phys. A Math. Gen. 2001, 34, 7111–7125. [Google Scholar] [CrossRef]
  64. Yang, Z.; Gao, M.; Qin, W. Transfer of high-dimensional quantum state through an XXZ-Heisenberg quantum spin chain. Int. J. Mod. Phys. B 2015, 29. [Google Scholar] [CrossRef]
Figure 1. (color online) (a) Two remote resonators, labelled 0 and N + 1 , are coupled to the ends of a coupled-resonator chain of N resonators with g 0 , the coupling strength between the two boundary resonators and the intermediate chain, and κ, the coupling strength between the intrachain elements. To achieve multiphoton transport, we introduce an auxiliary resonator, labelled N + 2 , to interact with the m-th resonator of the intermediate chain, with strength J 0 ; (b) In uncoupled case of J 0 = 0 , the intermediate chain could, under the assumption that g 0 κ , coherently couple the two spatially separated resonators, such that photons are transported between the two boundary resonators after the time evolution; (c) However, when g 0 J 0 κ , the boundary resonators are decoupled from the intermediate chain owing to the large detunings. In this case, photons are reflected back and remains unchanged.
Figure 1. (color online) (a) Two remote resonators, labelled 0 and N + 1 , are coupled to the ends of a coupled-resonator chain of N resonators with g 0 , the coupling strength between the two boundary resonators and the intermediate chain, and κ, the coupling strength between the intrachain elements. To achieve multiphoton transport, we introduce an auxiliary resonator, labelled N + 2 , to interact with the m-th resonator of the intermediate chain, with strength J 0 ; (b) In uncoupled case of J 0 = 0 , the intermediate chain could, under the assumption that g 0 κ , coherently couple the two spatially separated resonators, such that photons are transported between the two boundary resonators after the time evolution; (c) However, when g 0 J 0 κ , the boundary resonators are decoupled from the intermediate chain owing to the large detunings. In this case, photons are reflected back and remains unchanged.
Entropy 18 00179 g001aEntropy 18 00179 g001b
Figure 2. (color online) Transmission fidelity F t and reflection fidelity F r as a function of J 0 / κ with either (a) g 0 / κ = 0 . 01 or (b) 0 . 005 for n = 2 , 3 and 5, corresponding to solid black lines, dashed red lines and dot-dashed blue lines, respectively. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Figure 2. (color online) Transmission fidelity F t and reflection fidelity F r as a function of J 0 / κ with either (a) g 0 / κ = 0 . 01 or (b) 0 . 005 for n = 2 , 3 and 5, corresponding to solid black lines, dashed red lines and dot-dashed blue lines, respectively. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Entropy 18 00179 g002
Figure 3. (color online) Numerical results of the transmission infidelity, σ t = 1 F t , as a function of g 0 / κ with either (a) n = 2 or (b) n = 5 in the uncoupled case of J 0 = 0 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Figure 3. (color online) Numerical results of the transmission infidelity, σ t = 1 F t , as a function of g 0 / κ with either (a) n = 2 or (b) n = 5 in the uncoupled case of J 0 = 0 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Entropy 18 00179 g003
Figure 4. (color online) Numerical results of the reflection infidelity, σ r = 1 F r , as a function of g 0 / κ with either (a) n = 2 or (b) n = 5 in the coupled case of J 0 / κ = 0 . 1 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Figure 4. (color online) Numerical results of the reflection infidelity, σ r = 1 F r , as a function of g 0 / κ with either (a) n = 2 or (b) n = 5 in the coupled case of J 0 / κ = 0 . 1 . The analytic upper bounds are represented by the bold red lines. Here, the evolution time t = τ and we choose N = 7 , m = 3 .
Entropy 18 00179 g004

Share and Cite

MDPI and ACS Style

Qin, W.; Long, G. Multiphoton Controllable Transport between Remote Resonators. Entropy 2016, 18, 179. https://doi.org/10.3390/e18060179

AMA Style

Qin W, Long G. Multiphoton Controllable Transport between Remote Resonators. Entropy. 2016; 18(6):179. https://doi.org/10.3390/e18060179

Chicago/Turabian Style

Qin, Wei, and Guilu Long. 2016. "Multiphoton Controllable Transport between Remote Resonators" Entropy 18, no. 6: 179. https://doi.org/10.3390/e18060179

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop