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

Entanglement dynamics for three nitrogen-vacancy centers coupled to a whispering-gallery-mode microcavity

Open Access Open Access

Abstract

The solid-state qubits based on diamond nitrogen-vacancy centers (NVC) are promising for future quantum information processing. We investigate the dynamics of entanglement among three NVCs coupled to a microtoroidal cavity supporting two counter-propagating whispering-gallery-modes (WGMs) in the presence of Rayleigh scattering. Our results indicate that the maximal entanglement among all the NVCs could be achieved through adjusting several key parameters, such as the scattering-induced coupling between the WGMs, the distance between the NVCs, and the NVC-WGM coupling strengths, as well as the frequency detuning between the NVC and cavity. We show that entanglement of the NVCs displays a series of damped oscillations under various experimental situations, which reflects the intricate interplay and competition between the Rayleigh scattering and the NVC-WGM coupling. The quantum dynamics of the system is obtained via solutions to the corresponding microscopic master equation, which agrees well with the numerical simulation results using the phenomenological master equation. The feasibility of our proposal is supported by the application of currently available experimental techniques.

© 2015 Optical Society of America

1. Introduction

Over recent years, increasing interest has focused on the strong coherent interactions between light and matter in the context of cavity quantum electrodynamics (CQED), which is of central concern in various applications in quantum information processing (QIP), quantum computation, and fundamental studies in quantum optics [15]. Among the platforms of CQED, the hybrid system consisting of whispering gallery mode (WGM) microcavities (with very small volumes Vm 100 μm3 [6] and high-Q factors Q ≥ 108 ~ 1010 [6, 7]) and solid-state-based emitters becomes quite promising thanks to the experimental advance in the fabrication of optical microcavities and in manipulation of solid-state qubits [810]. Especially, the nitrogen-vacancy center (NVC) consisting of a substitutional nitrogen atom and an adjacent vacancy in diamond has gained widespread interest since the NVC is one of the promising building blocks for room-temperature quantum computation and becomes one of the excellent candidates for QIP owing to extremely long electronic and nuclear spin lifetimes as well as the capability for coherent manipulation in an optical fashion, such as fast initialization, high-fidelity information storage [1115] and well qubit readout [1619].

In contrast to the standing modes in the conventional Fabry-Perot cavity, the WGM cavity [2028], such as the microtoroidal [29, 30], microcylinders [31], microdisks [32, 33], and microspheres [34, 35], could support two counter-propagating traveling modes (also called propagating twin modes: the clockwise (cw) and counter-clockwise (ccw)) with the degenerate frequency and the same field distribution function, and they could couple to each other with a scattering strength due to the scattering of imperfection, which has been observed experimentally [3639]. In addition, the WGMs in microcavities travel around the curved boundary and are confined by continuous total internal reflection [40], which enables the NVCs in the vicinity of the cavity to interact with the two WGMs via the evanescent field [41, 42]. Therefore, coherently coupling the NVCs to microcavities or even nanocavities offers predominant conditions for reaching strong coupling regime and provides an excellent platform for QIP applications [20, 4356]. Experimentally, the efficient coupling of a NVC to a microcavity has been demonstrated [35, 5760]. Recent experimental progresses about the nanocrystal-microsphere system also provide experimental evidence for strong coupling between the NVCs and the WGM of silica microsphere [57] or polystyrene microsphere [58], respectively. Based on the hybrid NVC-cavity systems [43,44], much research has been recently devoted to the generation of entanglement between separated NVCs by coupling to the same quantized cavity mode using a probabilistic way [53,54], deterministic and robust method [52], the Raman transition method or adiabatic passage technique [4850, 55], dissipation-assisted steady entanglement [51], and the common mediation of the structured environment [56].

Most of these above-mentioned studies are mostly focused on the dynamics of bipartite entanglement measures between a pair of NVCs. The multiqubit entanglement dynamics of the NVCs has not been extensively explored because the multiqubit dynamics itself is particularly complicated and well-accepted measures for entanglement are still missing [6164]. However, it was generally agreed that entanglement in the systems with more than two parties exhibits a richer dynamical behavior, and more interesting features could be expected [6569]. In this sense such studies not only provide information of how entanglement evolves in time but also suggest ways toward practical purposes with entanglement [61, 62]. Here another important issue is how to develop efficient methods for generating highly entangled states among multiqubits. Therefore, it would be interesting to conduct theoretical and experimental studies on the dynamics of hybrid systems exhibiting a high degree of multi-partite entanglement, where the multiqubit-entanglement dynamics could be controlled by adjusting the tunable parameters. Meanwhile, the related question arises as to what extent entanglement generation and dynamics in such a multipartite system can be obtained and controlled. Therefore, it is desirable to investigate the preparation and dynamics of the entangled states among several distant NVCs, which are prerequisites for realization of large-scale spin-based quantum networks [12, 13, 70]. Furthermore, the WGMs carry phase factors, which plays a central role in the multi-scatterer system, and the Rayleigh scattering brings significant influences in the entanglement generation and photon transportation [21, 22]. However, there have not been well-done studies so far on the effect of the phase factors and the Rayleigh scattering on the multiqubit-entanglement dynamics. We will study below the relation between these physical mechanism and the entanglement generation, and explore the possibility to obtain maximal entanglement among all the NVCs through tuning the key parameters related to the Rayleigh scattering.

In the present paper we investigate the multiparticle solid-state CQED system consisting of a high-Q microtoroidal cavity and three separated diamond nanocrystals with each containing a single NVC. The dissipative effects associated with both the WGM and NVCs are taken into account, and the entanglement is characterized by using the concept of the lower bound of concurrence (LBC) [61, 62, 71]. Our results indicate that the maximum LBC among all the NVCs could be achieved through accurately adjusting the tunable parameters d, g and Δ, where d is the distance between the separated NVCs, g is the scattering strength between the twin modes of the WGMs, and Δ is the frequency detuning between the NVC and cavity. We also show that the system displays a series of damped oscillations under various experimental situations, where the dynamics of the system reflects the intricate balance and competition between the Jaynes-Cummings (JC) type of NVC-WGM couplings and the Rayleigh scattering process of the WGMs. Our further study reveals that using a realistic microscopic model combined with our detailed analysis could find a way to extract the optimal experimental parameters for maximal entanglement of the NVCs. In addition, we resort to the perspective of normal modes of the microcavity to clarify the physical picture behind the multiqubit-entanglement dynamics. For our purpose, the exact entanglement dynamics of the NVCs under dissipation is investigated analytically using the method of microscopic master equation [7276], which agrees well with the numerical simulation result using the phenomenological Markovian master equation [7779]. Therefore, the present system provides a platform to generate multiparticle quantum entanglement among three or more NVCs embedded in distant diamond nanocrystals, which may be another route toward building a distributed QIP architecture based on the increasingly-developed nanoscale solid-state technology.

The remainder of this paper is organized as follows. We present our system architecture in details in Sec. 2. Sec. 3 is devoted to the dynamics evolution of entanglement among the NVCs using two different methods to solve the quantum master equation employed in our model. Finally, we compare the results of both the approaches and conclude our findings in Sec. 4. Some details of deductions can be found in Appendixes.

2. System Hamiltonian

The system under consideration consists of three separated NVCs located on the surface of a microtoroidal cavity, where the two counter-propagating WGMs acw and accw with the degenerate frequency ωc and decay rate κm (m = cw,ccw), could couple to each other through the process of Rayleigh scattering [43, 44], and the position of each NVC can be controlled accurately [39, 58, 80, 81], illustrated in Fig. 1. Each NVC (usually treated as an electron spin-1) is negatively charged with two unpaired electrons located at the vacancy, where the introduction of an external magnetic field B0 along the quantized symmetry axis of the NVC could lift the degeneracy of the levels |3A,ms = ±1⟩ with an level splitting Deg [82]. Here we choose |A2=(|E,ms=+1+|E+,ms=1)/2 (one of the six excited states defined by the method of group theory) as the excited state |e⟩ and the ground state |g⟩ is encoded into the state |3A,ms=−1⟩ with the orbital state |E0⟩. For |E±⟩ and |E0⟩, the angular momentum projection along the NVC axis are ±1 and 0, respectively. According to the total angular momentum conservation, the dipole transition between the excited state and ground state will be accompanied by absorbing or releasing photons with σ+ circular polarization [13, 55, 56] and frequency ω0. For the jth NVC, the spontaneous emission rate is Γj and the pure dephasing rate is γj. In our scheme, the WGM is coupled to the transition |g↔ |e⟩ with the coupling strength Gj, and the NVCs are fixed and separated by the distance dij (between the i-th NVC and j-th NVC). Taking the rotating wave approximation under the condition ωc, ω0Gj, the Hamiltonian of the whole system can be written as (in units of ħ = 1) [20, 23, 43, 44],

H=H0+H1+H2,H0=j=13ω02σjz+ωcacwacw+ωcaccwaccw,H1=j=13Gj(eikd1jσj+acw+eikd1jσj+accw)+H.C.,H2=j=13gj(acwacw+accwaccw)+j=13gj(e2ikd1jaccwacw+H.C).,
where d11 = 0, σj+=|ejgj| and σjz=|ejej||gjgj| are usual Pauli operators for the j-th NVC. am(am) is the creation (annihilation) operator for the WGM with the wave vector kcw = kccw = k. We have considered the phase factors e±ikd1j because the interaction between the NVC and field varies when the NVC’s location in the field changes. The Hamiltonian H0 describes the free evolution of the system consisting of the NVCs and WGMs. The Hamiltonian H1 describes the NVC-WGM dipole interaction with the coherent coupling strength Gmax=Gj=μ[ωc/(2ε0εsVc)]1/2fc(r)(j=1,2,3), where fc(r)=|E(r)/Emax| is the normalized field distribution function of the WGMs, Vc is the quantized volumes of the WGM interacting with the NVC, μ = 2.74 × 10−29Cm is the dipole moment of the NVC transition, and ε0(εs) is the electric permittivity of the vacuum (surrounding medium) [43, 44]. The Hamiltonian H2 describes the diamond nanocrystal-induced scattering into the co-propagating (m = m′) or counter-propagating (mm′) WGM fields with the strengths gjmm. Note that the scattering of light should be taken into consideration when the diamond nanocrystals are coupled to ultrahigh-Q microcavities, because the size of the diamond nanocrystals is much larger than the neutral atoms/ions used in conventional CQED systems. For a subwavelength nanocrystal, we can model the interaction between WGMs and NVC by the Weisskopf-Wigner semi-QED method [38], in which the electric field of the WGM’s m-mode induces a dipole moment at the location of the n-th NVC in terms of the dipole approximation and the polarization is described by Pn,m. The interaction energy Pn,mEn,m between m′ and m, with En,m the electric field of the m′-mode at the location of the n-th NVC, can be regarded as the scattering-induced interaction by the n-th NVC. Then the scattering-induced coupling strengths can be calculated by the relation gjmm=αfc2(r)ωc/(2Vc) in the case of the elastic Rayleigh scattering, where α = 4πR3 (εd − εs)/(εd + 2εs) is the polarizability for the spherical diamond nanocrystal with the electric permittivity εd = 2.42 and the radius R [43, 44]. Here we regard the NVCs as entities which are individually coupled to the WGM with the identical strengths Gj = G, and the scattering-induced coupling coefficients gj are also assumed to be equal for simplicity, namely, gj = g. One can find that the scattering-induced coupling coefficient g grows rapidly with the increase of R, and there is a phase interference induced by different distances d12 (d13) between NVC1 and NVC2 (NVC3) in the present multi-scatterer case, which is different from the single-scatterer case.

 figure: Fig. 1

Fig. 1 (a) Schematic of the system consisting of three NVCs and a microtoroidal cavity supporting two counter-propagating WGMs acw and accw. (b) Level diagram for the j-th NVC, where Δj is the frequency detuning of the NVC from the cavity. Deg = γeB0 is the level splitting induced by an external magnetic field B0 with γe the electron gyromagnetic ratio.

Download Full Size | PDF

3. Quantum dynamics of entanglement among three NVCs

In this section, we investigate the dynamics evolution of entanglement among the NVCs, where the main dissipative effects originating from the leakage of the WGM photons, the spontaneous emission and pure dephasing of the excited-state of NVCs are taken into account by two different methods, respectively. One of the methods analytically uses the microscopic master equation [7276], which goes back to the ideas of Davies on how to describe the system–reservoir interactions in Markovian master equations [77, 78]. This method (called the microscopic case) focuses on the quantum jumps between the eigenstates of the system Hamiltonian rather than the eigenstates of the field-free subsystems. The other (called the phenomenological case) exploring the dynamics of the system numerically uses the phenomenological master equation with a dissipative (Lindblad) term, which is employed as one of the major approaches in quantum optics and QIP (See the Appendix A).

Based on the above-mentioned two methods, we investigate the dynamics of entanglement among three NVCs. As we all know, for a general pure tripartite state |φH1H2H3, the concurrence can be expressed as [83, 84]

C3(|φ)=13[3Tr(ρ12)Tr(ρ22)Tr(ρ32)],
and ρi is the reduced density matrix obtained by tracing over the remaining subsystem j and k (i, j, k = 1, 2, 3 and ijk). Then it can be extended to the tripartite mixed state ρ by the convex roof [85, 86],
C3(ρ)=min{pi,φi}ipiC3(|φi)
where {|φi⟩} denotes the collection of all the possible pure states into which the mixed state ρ can be decomposed as ρ = ∑i pi |φi⟩ ⟨φi| with the positive pi (normalized already). However, it is indeed a difficult task to find the minimum because the optimization procedure involves a large number of free parameters. What’s worse, no numerical algorithm could guarantee to find the global minimum. Thus, to avoid finding the exact minimum, the lower bound to this measure is available [6163, 71] with the following definition
C¯3(ρ)=13j=16[(Cj12|3(ρ))2+(Cj31|2(ρ))2+(Cj23|1(ρ))2],
Cjkl|m(ρ)=max{0,λj,1kl|mn>1λj,nkl|m},
where λj,nkl|m represents the eigenvalues of the matrix ρ˜=ρ(Ljklσym)ρ*(Ljklσym) in a decreasing order with Ljkl(j=1,2,6) the six generators of the group SO(4) [87] acting on the qubits k,l and σym the Pauli matrix acting on the qubit m. If C3(ρ)=0, then the state ρ is separable. But, if C¯3(ρ)=0, it does not necessarily mean it is separable. Without doubt, C¯3(ρ)>0 implies an entangled state and a separable state always results in C¯3(ρ)=0. We may calculate numerically the LBC C¯3(ρ)num for ρnum (derived from the phenomenological case), and we may also obtain the analytical expression of the LBC for ρana (derived from the microscopic case) written as
C¯3(ρ)ana=max{0,85(ρegg,eggρgeg,geg+ρegg,eggρgge,gge+ρgeg,gegρgge,gge)}.

Here the LBC can be approximately determined by the analytical expression (Eq. (6)) with respect to the three key time-dependent density matrix elements ρegg,egg, ρgge,gge, and ρgeg,geg. One can find that the quantum dynamics based on the analytical method agrees well with the numerical simulation by the phenomenological case, as shown later.

More specifically, we investigate the dynamics of entanglement among the NVCs under the resonant case and detuning case, respectively, via the approach of microscopic master equation. Here we consider two different initial states of the system: (1) all the NVCs are in the ground state and one photon is prepared in one of the twin modes, i.e., ψ(0) = |g1,g2,g3⟩|1cw,0ccw⟩; (2) the first NVC is in the excited state, and the other two NVCs are in the ground state, without any photons in the WGM, i.e., Ψ(0) = |e1,g2,g3⟩|0cw,0ccw⟩.

3.1. The resonant case

The distinct advantages of the present system, such as the individual accessibility and high tunability of the parameters, make this NVC-WGM system an ideal tunable structure, where the scattering-induced coupling rates can be regulated by the distance between the NVCs or the radius of the diamond nanocrystal, and the dipole interaction between the NVC and the WGM can also be adjusted by changing the relative position of the NVC to the microcavity surface. Figure 2 shows the dependence of the LBC dynamics on the distances (d12 and d13) or the phase differences (θ12 = kd12 and θ13 = kd13) between any pair of NVCs in the resonant case (Δ = 0), where we assume that the distance between NVC1 and NVC2 is fixed at π2+nπ or , and the position of NVC3 is flexible. As shown in Fig. 2, the phase interference induced by the spatial distance dij between different NVCs leads to profile/frequency variations of the oscillation and the maximum values of the LBC. For a given value of θ12, as shown in Fig. 2, the behavior of the LBC shows periodic evolution with the value of θ13 changing; and for a given value of θ13, the LBC oscillates between zero and maximum due to the Rabi oscillations between the NVC and the quantized WGM field, the amplitude of which is decaying in time arise from the dissipation effects.

 figure: Fig. 2

Fig. 2 The LBC vs time t and the phase difference θ13 when the system is initially prepared in the state ψ(0). (a) θ12=π2+nπ, (b) θ12 = . The time parameter is dimensionless and we set G = 1. Other parameters are Δ = 0, g = 0.1, Γ = 0.05, γ = 0.01 and κ = 0.03.

Download Full Size | PDF

To get a clear picture of how the LBC evolves in the space of phase difference {θ12, θ13}, we plot the time evolution of the LBC in Fig. 3, where an interesting feature is that entanglement of the NVCs shows a periodic oscillation behavior along both the θ12 axis and the θ13 axis, and the period of the maximal value of the LBC is π. Furthermore, the period of the LBC becomes π/2 along the lines of θ12 = (n+1)π/2 or θ13 = (n+1)π/2, due to the fact that each NVC exerts the same influence on the two WGMs when θ12 = (n+1)π/2 or θ13 = (n+1)π/2. Therefore, the values of LBC are strongly dependent on the interposition of the NVCs, and a subwavelength displacement could result in an evident change of the entanglement. From another point of view, this promises the possibility of utilizing the position-sensitive response to realize the desired maximal entanglement of NVCs, i.e., high-degree LBC of the NVCs could be obtained by independently adjusting the inter-qubit distance so that some special values of phase differences can be taken. Figure 3(b) shows that the maximum of the LBC can reach 0.92 with an optimal choice of θ12 = 33π/50 + and θ13 = 7π/25 + , where the three NVCs are placed with almost equal intervals.

 figure: Fig. 3

Fig. 3 (a) The maximal values of the LBC in the parameter plane of {θ12, θ13} with Δ = 0, G = 1, g = 0.1, Γ = 0.05, γ = 0.01 and κ = 0.03, where the system is initially prepared in the state ψ(0). (b) The zooming-in plot of (a) with one period.

Download Full Size | PDF

From the corresponding eigenvectors |ϕi of Eq. (10), we see that the time-dependent superposition of five harmonic wave functions ultimately determines the single-excitation dynamics of the system, and thereby the overall dynamics exhibits some complex interferences or competing effects. The remarkable result is the rich entanglement dynamics of the system reflecting the intricate balance and competition between the two different kinds of couplings. One of such couplings is the JC-type interaction between each NVC and WGM with the coupling strength G, and the other is the coupling between the WGMs induced by the Rayleigh scattering of the NVC with the coupling strength g, where these two tunable parameters (G and g) influence the system dynamics in different ways. Therefore the dynamical feature of the LBC can be characterized by two kinds of oscillation behavior with two periods depending on the related coupling strengths. When the system is initially prepared in the state ψ(0), the concrete physical process can be described in more details by following image. The excitation is initially located in the clockwise WGM acw: The energy is exchanged between three NVCs and acw with a rate G/2π. Then the excitation gradually flows to the counter-clockwise WGM accw with a rate g/2π through the Rayleigh scattering mechanism during the JC oscillation between the NVC and acw. At this stage, a similar JC oscillation between the NVC and the mode accw also occurs. As a result, three NVCs achieve a time-dependent entanglement during the system evolution.

An important aspect of the present system is the ability to tune the Rayleigh scattering strength. Figure 4 displays the influence from the Rayleigh scattering on the entanglement dynamics for different initial states ψ(0) and Ψ(0), respectively, where the phase difference is fixed. To better understand the physics underlying these figures, we consider some illustrations on following three different regimes. The first one is called the strong coupling regime with Gg, where the NVC-WGM dipole interaction is much larger than the Rayleigh scattering, implying that the NVC-WGM interface dominates the single-excitation evolution process. Thus the magnitude of large coupling strength G determines the relatively large speed of the energy-exchange between the NVC and the WGM. On the other hand, the excitation will ultimately tunnel between the twin modes acw and accw with a relatively smaller speed. The second one is the strong scattering regime with Gg, where the excitation-transfer between the twin modes dominates the evolution process. In between, lies the third one called the competition regime with G ≈ g, where the coupling strength is comparable to the scattering strength. From Fig. 4(a1) one can find that the large degree LBC can be achieved at some appropriate time points when the system is within the strong coupling regime and initialized in the state ψ(0). With the growth of the values of g, as shown in Fig. 4(a2) and Fig. 4(a3), the dynamics will exhibit two remarkable characters: the maximum of LBC gradually decreases and the LBC among the NVCs oscillates faster. The physical picture behind is that the excitation mainly transmits between the two twin modes acw and accw of the WGMs in the strong scattering regime. Therefore, both the maximal LBC among the NVCs and the periodic time inevitably reduce when g increases since the major populations with respect to the excitation almost resides in the twin modes, rather than the NVCs. The situation becomes different when the system is prepared in the state Ψ(0). Figures 4(b1–b3) indicate that the scattering process plays a positive role in the generation of entanglement of the NVCs, similar to the conclusions in the previous work [2022, 24]. The overall dynamics of the system is characterized by a fast oscillation behavior accompanied by a slow oscillation envelope, as shown in Fig. 4(b3).

 figure: Fig. 4

Fig. 4 The LBC vs time t and the coupling strength g for different initial states of the system (a) ψ(0) and (b) Ψ(0). Other parameters are θ12=π4+nπ, θ13=π8+nπ, Δ = 0, G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03. In the bottom six subgraphs, the curves of the LBC dynamics are plotted in (a1,b1) g = 0.1, (a2,b2) g = 1, (a3,b3) g = 5.

Download Full Size | PDF

3.2. The detuning case

We now turn our attention to the entanglement dynamics in the detuning case of Δ ≠ 0. We plot the dependence of the LBC on the phase difference θ13 (in Figs. 5(a,c)), and the scattering-induced coupling strength g (in Figs. 5(b,d)) when the system is initially prepared in the state ψ(0) under different detunings Δ, respectively. Compared with Fig. 2(b), Figures 5(a,c) presents the remaining periodicity of the LBC with respect to θ13. The maximal LBC decreases greatly and the oscillation frequency increases when the value of Δ increases. Moreover, in comparison with Fig. 4(a), there is a slight dropping of the maximal value of the LBC in Figs. 5(b,d) caused by the growth of Δ. Another interesting feature is that the optimal coupling strength gop corresponding to the maximal LBC augments with the growth of Δ. It implies that the LBC evolution can be optimized by appropriately tailoring those key parameters. This provides a useful and effective way to controlling the dynamics of entanglement among the NVCs.

 figure: Fig. 5

Fig. 5 The LBC vs time t and the phase difference θ13 under different detunings (a) Δ = 5 and (c) Δ = 10, where θ12 = , G = 1, g = 0.1, Γ = 0.05, γ = 0.01 and κ = 0.03. The LBC vs time t and the coupling strength g under different detunings (b) Δ = 5 and (d) Δ = 10, where θ12=π4π, θ13=π8+nπ, G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03.

Download Full Size | PDF

Figure 6 displays the time-dependent LBC on the detuning Δ under three different regimes with the fixed phase differences θ12 and θ13. In the strong coupling regime in Fig. 6(a), the large degree LBC can be obtained at a specific time, and the increase of the detuning Δ destroys the entanglement of the NVCs. When the coupling strength is comparable to the scattering strength, i.e., G = g, as shown in Fig. 6(b), the intricate balance and competition between these two kinds of couplings are more obvious than the case in Figs. 6(a,c). In the strong scattering regime (Fig. 6(c)), the maximal value of the LBC drops and the optimized values of Δ for the maximal entanglement are increased in comparison with in Figs. 6(a,b) which is consistent with the features in Figs. 5(b,d).

 figure: Fig. 6

Fig. 6 The LBC vs time t and the detuning Δ for different scattering strengths (a) g = 0.1, (b) g = 1, (c) g = 5. The other parameters are θ12=π4+nπ, θ13=π8+nπ, G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03.

Download Full Size | PDF

In order to clarify the relation between the detuning Δ and the coupling strength g as well as the entanglement dynamics itself, we resort to the perspective of normal modes of the micro-cavity, i.e., the use of the normal modes a1=(acw+accw)/2 and a2=(acwaccw)/2 to replace the cavity modes acw and accw. Then the total Hamiltonian is transformed to

H=j=13[ω0σejej+δ1a1a1+δ2a2a2+G˜1j(a1σj+a1σj+)]+j=23G˜2j(a2σj+a2σj+)+g˜(a1a2a2a1)
where G˜ij is the effective coherent coupling strength between the i-th normal mode ai and the j-th NVC with G˜11=2G, G˜12=2Gcos(kd12), G˜13=2Gcos(kd13), G˜22i2Gsin(kd12), and G˜23=i2Gsin(kd13). Eq. (7) indicates that three NVCs simultaneously couple to the normal mode with frequency δ1 = ωc + 2g [1 + cos2 (kd12) + cos2 (kd13)], but only NVC2 and NVC3 couple to the normal mode a2 with frequency δ2 = ωc + 2gsin2 (kd12) + sin2 (kd13). These two normal modes couple to each other with the effective scattering-induced coupling strength g˜=ig(sin(2kd12)+sin(2kd13)). One can find that the system is equivalent to the three NVCs coupled to a single normal mode a1 with frequency ωc + 6g in the case of kd12 = kd13 = since the normal mode a2 decouples to all NVCs.

In the detuning case, three NVCs couple to the normal mode a1 with frequency δ1=ω0Δ+2g[1+cos2(kd12)+cos2(kd13)]. NVC2 and NVC3 couple to the normal mode a2 with frequency δ2=ω0Δ+2g(sin2(kd12)+sin2(kd13)). When kd12 = kd13 = , the normal mode a2 decouples to all NVCs and the frequency of the normal mode a1 becomes ω0 − Δ+6g. The modes a1 and a2 are decoupled. Thus entanglement of the NVCs turns to be maximal once the condition Δ = 6g is well satisfied. To check whether this conclusion is correct, the density plots of the LBC versus Δ and g with kd12 = kd13 = is presented in Fig. 7(a), where the maximal LBC in or near the brightest line Δ = 6g is larger than in other regions. But the situation becomes more complicated if kd12 = kd13, where the normal mode a2 is not decoupled to all NVCs, and thus it is hard to get an accurate relation between Δ and g. Here we plot the time evolution of the LBC under three different cases respectively: (i) kd12=π2+nπ, kd13=π4+nπ, for which the resonance condition between the normal modes a1(2) and NVC is Δ = 3g, and the coupling strength between a1 and a2 is g. The results can be found in Fig. 7(b). (ii) kd12=π4+nπ, kd13=π8+nπ, where the resonance condition between the normal mode a1 and the NVCs is Δ = 4.7g, the resonance condition between the normal mode a2 and NVC2 (NVC3) is Δ = 1.3g, and the coupling strength between a1 and a2 is 1.7g. As plotted in Fig. 7(c), the large frequency difference between a1 and a2 splits the brightest line into two lines along different direction. (iii) kd12=3π4+nπ, kd13=3π8+nπ, where the resonance condition between the normal mode a1 and NVCs is Δ = 3.3g, the resonance condition between the normal mode a2 and NVC2 (NVC3) is Δ = 2.7g, and the coupling strength between a1 and a2 is 0.3g (See Fig. 7(d)).

 figure: Fig. 7

Fig. 7 The maximal LBC vs the coupling strength g and the detuning Δ in (a) θ12 = θ13 = , (b) θ12=π2+nπ and θ13=π4+nπ, (c) θ12=π4+nπ and θ13=π8+nπ, (d) θ12=3π4+nπ and θ13=3π8+nπ. Other parameters are G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03.

Download Full Size | PDF

4. Discussion and conclusion

We first compare the analytical results from the microscopic master equation with the results of numerical calculation using the phenomenological master equation. As shown in the Fig. 8 (a), the LBC dynamics obtained by two methods are almost completely coincident. We further define a new function δLBC=C¯3(ρ)anaC¯3(ρ)nun to visualize the slight difference between these two different methods, as shown in Fig. 8(b). In fact, the numerical simulation agrees well with the calculation result in Sec. 3.

 figure: Fig. 8

Fig. 8 (a) The LBC vs time t under the condition θ12=π4, θ13=π8, Δ = 0, g = 2, where the red-solid (blue-dashed) lines denote the analytical (numerical) results, respectively. (b) The slight difference between the LBC dynamics using these two different methods.

Download Full Size | PDF

Next we survey the relevant experimental parameters. Under practical situations, the transition of the NVCs at the wave-length 637 nm has a decay rate with Γ/2π = 13 MHz and a pure dephasing rate with γ/2π = 2 ~ 11 MHz [50, 88]. Due to the cavity coupling that enhances the photonic emission into the zero-phonon line, we consider Vc ~ 200 μm3, and fc(r)=0.47, yielding the maximal coupling strength Gmax/2π ≈ 180 MHz [33, 43, 44, 59]. In addition, for a WGM cavity with the quality factor Q = 108, we can obtain the related cavity decay rate κ/2π = ωc/2πQ = 4.7 MHz, which is about 0.03Gmax [43, 44]. For the NVC, the electron spin relaxation time T1 of the diamond NVC ranges from 6 ms at room temperature to seconds at cryogenic temperature in realistic experiments. The dephasing time T2 = 350 μs induced by the nuclear spin fluctuation inside the NVC has been reported [89]. A latest experimental progress [90] with isotopically pure diamond sample has demonstrated a longer dephasing time, i.e., T2 = 2 ms. This implies that the influence from the intrinsic damping and dephasing of the NVCs is possibly negligible in the specially treated NVC-WGM system.

In summary, we have investigated the entanglement dynamics of three NVCs coupled to a WGM cavity supporting counter-propagating twin modes. We have shown that the system displays a series of damped oscillations, which reflects the intricate balance and competition between the JC type NVC-WGM coupling and the WGM Rayleigh scattering. Our study has also provided a way to extracting the optimal experimental parameters for maximal entanglement of the NVCs. Moreover, straightforward extension of our idea to more than three NVCs is possible. The good agreement between the analytical and numerical results made our results credible and convincing. Therefore, the present study are valuable for building the future full-scale quantum information processor based on the increasingly-developed nanoscale solid-state technology.

APPENDIX A: The quantum master equation

In this section we present the details of our treatments for the quantum master equation by both microscopic and phenomenological considerations. The comparison between the two treatments can make our results more credible, and particularly the analytical expressions help further understand the physics in entanglement.

4.1. The microscopic case

We first describe the microscopic case using the method of the microscopic master equation. For this system described by the Hamiltonian H, there exist two invariant subspaces spanned by the following basic vectors

1{|1=|e1g2g30cw0ccw,|2=|g1e2g30cw0ccw,|3=|g1g2e30cw0ccw,|4=|g1g2g31cw0ccw,|5=|g1g2g30cw1ccw},
2{|6=|g1g2g30cw0ccw},
where gj (ej) denotes the ground (excited) state of the j-th NVC, 0cw(ccw) (1cw(ccw)) denotes the vacuum state (one-photon Fock state) of the cw (ccw) WGM. One can find that the ground state |6⟩ is a dark state, i.e., with no evolution in time. In addition, the system initially in the single-excitation manifold ∀1 will remain in that subspace. With these basic vectors, one can rewrite the Hamiltonian H in a matrix with less dimensions as follows,
H=(ω0200GG00ω020eikd12Geikd12G000ω02eikd13Geikd13G0Geikd12Geikd13G3ω02+ωc+3gg(1+e2ikd12+e2ikd13)0Geikd12Geikd13Gg(1+e2ikd12+e2ikd13)3ω02+ωc+3g0000003ω02)

Here the corresponding eigenvectors of H can be expressed by |ϕi=j=15cij|j with i = 1,2,3,4,5, and |ϕ6⟩ = |6⟩, and the forms of the coefficients cij are presented in Appendix B.

For this NVC-WGM system at zero temperature, the master equation for the density operator ρ(t) is given by

ρ˙(t)=i[H,ρ]+ω¯>0,m={cw,ccw}κm(ω¯)×[Am(ω¯)ρ(t)Am(ω¯)12{Am(ω¯)Am(ω¯),ρ(t)}]+ω¯>0,n={1,2,3}Γn(ω¯)×[n(ω¯)ρ(t)n+(ω¯)12{n+(ω¯)n(ω¯),ρ(t)}]+ω¯>0,n={1,2,3}γn(ω¯)×[nz(ω¯)ρ(t)nz(ω¯)12{nz(ω¯)nz(ω¯),ρ(t)}],
where Am(ω¯) and n(ω¯) are the Davies operators given by Am(ω¯)=|ϕiϕi|Am|ϕjϕj| with Am=am+am, n(ω¯)=|ϕiϕi|n|ϕjϕj| with n=σn+σn+ and nz(ω¯)=n+(ω¯)n(ω¯)n(ω¯)n+(ω¯). And ω¯=λjλi must be non-negative with λi the eigenvalues of H. The sum on m and n is over all the dissipation channels and the dissipative parameters κm(ω¯), Γn(ω¯) and γn(ω¯) are the Fourier transform of the correlation functions of the environment [79]. Under the condition of the single excitation, we find that only the terms such as 6|σ1|1, 6|σ2|2, 6|σ3|3, 6|acw|4, 6|accw|5 and their Hermitian conjugations are non-zero. Therefore, we can use these Davies operators to simplify the master equation (Eq. (11)) as
ρ˙(t)=i[H,ρ]+i=15Γ˜i[|ϕ6ϕi|ρ(t)|ϕiϕ6|12{|ϕiϕi|,ρ(t)}]+i=15γ˜i[(|ϕiϕi||ϕ6ϕ6|)ρ(t)(|ϕiϕi||ϕ6ϕ6|)12{(|ϕiϕi|+|ϕ6ϕ6|),ρ(t)}],
where the definition of the parameters Γ˜i and γ˜i(1,2,3,4,5) and some other details of our deductions are presented in Appendix B.

Now we can calculate the time-dependent density matrix of the total system by solving a set of differential equations under the given initial state |ψ(0)⟩. For uniformity, we should transform the initial state into the representation of the Hamiltonian H through the unitary transformation operator U˜ with

U˜=(c11c12c13c14c150c21c22c23c24c250c31c32c33c34c350c41c42c43c44c450c51c52c53c54c550000001).

After tracing out the freedom degree of the WGM and transforming back to the original basis {|e⟩, |g⟩}, we can obtain the reduced density matrix ρana(t) with respect to the three NVCs. The detailed derivation is presented below in Appendix B.

4.2. The phenomenological case

The entanglement dynamics of the NVCs can also be numerically explored using the phenomenological master equation. Different from the microscopic case, the calculation in the phenomenological case is not made in the eigenstates of the system Hamiltonian, but in a much larger space whose dimension grows exponentially with the number of qubits.

In order to model this open (dissipative) system, coupling to the environment (in the limit of weak coupling to a Markovian bath) can be considered in a phenomenological master equation with a dissipative (Lindblad) term on the right-hand side

ρ˙=i[Heff,ρ]+ρ,
where
ρ=m=cw,ccwκm(amρam+12am+amρ12ρam+am)+j=13Γj(σjρσj+12σj+σjρ12ρσj+σj)+j=13γj(σjzρσjz12σjzσjzρ12ρσjzσjz).

Here κm is the damping rate of the twin modes of the cavity, and Γj and γj are the spontaneous emission rate and pure dephasing rate of the j-th NVC respectively. For simplicity, we set κcw = κccw = κ, Γ1 = Γ2 = Γ3 = Γ and γ1 = γ2 = γ3 = γ in the following calculation. In Eq. (14), the Hamiltonian Heff in the interacting picture with the unitary transformation U=eiH0t reads

Heff=j=13[eiΔtGj(eikd1jσj+acw+eikd1jσj+accw)+gje2ikd1jaccwaccw+H.C]+j=13gj(acwacw+accwaccw),
where Δ = ω0ωc is the detuning between the WGM mode and the transition of the NVC.

In order to study the dynamics of entanglement among all the NVCs, we numerically solve Eq. (14) with the given initial state |ψ(0)⟩, and then trace over the twin modes of the cavity to obtain the time-dependent reduced density matrix ρnum of the NVCs.

APPENDIX B: Analytical solutions to the microscopic master equation

In this appendix, we give detailed deductions for the analytical solutions to the microscopic master equation.

Firstly, we use Eqs. (8) and (9) to rewrite the Hamiltonian H into the matrix form shown in Eq. (10), and we can obtain the corresponding eigenvalues λ1 = 0, λi=e2ikd122ikd13Ai1, and λ6 = −ω0 with i = 2,3,4,5. The corresponding eigenvectors of H (Eq. (10)) can be expressed by |ϕi=j=15cij|j and |φ6⟩ = |6⟩ with the forms of the coefficients cij as

c11=eikd13(e2ikd13e2ikd12)/(e2ikd121),c12=eikd12ikd13(1e2ikd13)/(e2ikd121),ci1=G{ξ[(e2ikd12+e2ikd132)η+ΔAi1]+Ai12}qAi1,ci2=eikd12G{ξ[(e2ikd12+e2ikd132)η+ΔAi1]+Ai12}qAi1,ci3=eikd13G{ξ[(e2ikd13+e2ikd122)η+ΔAi1]+Ai12}qAi1,ci4={ξ[3ηΔAi1]Ai12}/ξq,c14=c15=c16=ci6=0,c13=ci5=1,
where i = 2,3,4, ξ=e2ikd12+2ikd13, η=e2ikd12+2ikd13G2+gAi1, and q=(1+e2ikd12+e2ikd13)(e2ikd12+2ikd13G2+gAi1). Here A1, A2, A3, A4 are the four roots of the equation
x4+ax3+bx2+cx+d=0,
with a = 2ξ (Δ − 3g), b = ξ2 [2εg2 − 6G2 − 6gΔ + Δ2], c = 2ξ3G2 [2εg − 3Δ], d = 2ξ4εG4 and ε = 3 − cos(2kd12) − cos(2kd13) − cos (2kd12 − 2kd13).

In order to solve the microscopic master equation Eq. (11), we rewrite the master equation with eigenvalues and eigenvectors of the Hamiltonian H. By calculation we obtain

Acw(λiλ6)=ci4|ϕ6ϕi|,Accw(λiλ6)=ci5|ϕ6ϕi|,
and
1(λiλ6)=ci1|ϕ6ϕi|,2(λiλ6)=ci2|ϕ6ϕi|,3(λiλ6)=ci3|ϕ6ϕi|.

Substituting Eqs. (19)(20) into Eq. (11), the master equation can be rewritten as Eq. (12) with

Γ˜i=j=13|cij|2Γj(λiλ6)+|ci4|2γcw(λiλ6)+|ci5|2γccw(λiλ6),γ˜i=j=13|cij|4γj(λjλ6).

At this stage, we just need to solve a set of different equations with the initial state ρ(0) = |ψ(0)⟩ ⟨ψ(0)| or ρ(0) = |Ψ(0)⟩ ⟨Ψ(0)| represented in the vector bases |ϕi⟩ (i = 1,2,…,6). Then the elements of the density matrix are given by

ρii(t)=ρii(0)eΓ˜it,ρij(t)=ρij(0)e[2i(λiλj)(Γ˜i+Γ˜j+γ˜i+γ˜j)]t/2,ρi6(t)=ρi6(0)e[i(λiλ6)Γ˜i2j=15(γ˜j+3γ˜i)2]t,ρ66(t)=j=15ρjj(0)(1eΓ˜jt)+ρ66(0).

Here the subscripts i and j in Eqs. (19)(22) are all running from 1 to 5, and the condition i < j should be satisfied because ω¯=λjλi must be non-negative.

Finally, tracing over the twin modes of the cavity, one can obtain the reduced density matrix ρana(t) with respect to three NVCs as

ρana(t)=i,j=16ϕi|ρ(t)|ϕj00|ϕiϕj|00+10|ϕiϕj|10+01|ϕiϕj|01+11|ϕiϕj|11],
where i = 1,2,3,4,5,6. Direct calculation yields the reduced density matrix ρana(t) as
ρana(t)=ρegg,geg|e1g2g3g1e2g3|+ρegg,gge|e1g2g3g1g2e3|+ρegg,ggg|e1g2g3g1g2g3|+ρgeg,egg|g1e2g3e1g2g3|+ρgeg,gge|g1e2g3g1g2e3|+ρgeg,ggg|g1e2g3g1g2g3|+ρgge,egg|g1g2e3e1g2g3|+ρgge,geg|g1g2e3g1e2g3|+ρgge,ggg|g1g2e3g1g2g3|+ρggg,egg|g1g2g3e1g2g3|+ρggg,geg|g1g2g3g1e2g3|+ρggg,gge|g1g2g3g1g2e3|+ρegg,egg|e1g2g3e1g2g3|+ρgeg,geg|g1e2g3g1e2g3|+ρgge,gge|g1g2e3g1g2e3|+ρggg,ggg|g1g2g3g1g2g3|,
where ρegg,egg=i,j=15ci1cj1*ρij, ρgeg,geg=i,j=15ci2cj2*ρij, ρgge,gge=i,j=15ci3cj3*ρij, ρegg,geg=i,j=15ci1cj2*ρij, ρegg,gge=i,j=15ci1cj3*ρij, ρgeg,gge=i,j=15ci2cj3*ρij, ρegg,ggg=i=15ci1ρi6, ρgeg,ggg=i=15ci2ρi6, ρgge,ggg=i=15ci3ρi6, and ρggg,ggg=ρ66+i,j=15(ci4cj4*+ci5cj5*)ρij.

Acknowledgments

W.L.Y. thanks Prof. Yun-Feng Xiao, Prof. Qiongyi He, and Dr. Zhangqi Yin for enlightening discussions. This work is supported partially by the National Fundamental Research Program of China under Grants No. 2012CB922102 and No. 2013CB921803, by the National Natural Science Foundation of China under Grants No. 11274351, No. 11104326 and No. 11274352.

References and links

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

2. Z. L. Xiang, S. Ashhab, J. Q. You, and F. Nori, “Hybrid quantum circuits: superconducting circuits interacting with other quantum systems,” Rev. Mod. Phys. 85, 623–653 (2013). [CrossRef]  

3. I. Buluta and F. Nori, “Quantum simulators,” Science 326, 108–111 (2009). [CrossRef]   [PubMed]  

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

5. Z. L. Xiang, X. Y. Lü, T. F. Li, J. Q. You, and F. Nori, “Hybrid quantum circuit consisting of a superconducting flux qubit coupled to a spin ensemble and a transmission-line resonator,” Phys. Rev. B 87, 144516 (2013). [CrossRef]  

6. D. K. Armani, T. J. Kippenberg, S. M. Spillane, and K. J. Vahala, “Ultra-high-Q toroid microcavity on a chip,” Nature (London) 421, 925–928 (2003). [CrossRef]  

7. M. L. Gorodetsky, A. A. Savchenkov, and V. S. Ilchenko, “Ultimate Q of optical microsphere resonators,” Opt. Lett. 21, 453–455 (1996). [CrossRef]   [PubMed]  

8. A. Badolato, K. Hennessy, M. Atatüre, J. Dreiser, E. Hu, P. M. Petroff, and A. Imamoğlu, “Deterministic coupling of single quantum dots to single nanocavity modes,” Science 308, 1158–1161 (2005). [CrossRef]   [PubMed]  

9. K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atatüre, S. Gulde, S. Fält, E. L. Hu, and A. Imamoğlu, “Quantum nature of a strongly coupled single quantum dot–cavity system,” Nature (London) 445, 896–899 (2007). [CrossRef]  

10. T. Schröder, A. W. Schell, G. Kewes, T. Aichele, and O. Benson, “Fiber-integrated diamond-based single photon source,” Nano Lett. 11, 198–202 (2011). [CrossRef]  

11. L. Childress, M. V. GurudevDutt, J. M. Taylor, A. S. Zibrov, F. Jelezko, J. Wrachtrup, P. R. Hemmer, and M. D. Lukin, “Coherent dynamics of coupled electron and nuclear spin qubits in diamond,” Science 314, 281–285 (2006). [CrossRef]   [PubMed]  

12. P. Neumann, R. Kolesov, B. Naydenov, J. Beck, F. Rempp, M. Steiner, V. Jacques, G. Balasubramanian, M. L. Markham, D. J. Twitchen, S. Pezzagna, J. Meijer, J. Twamley, F. Jelezko, and J. Wrachtrup, “Quantum register based on coupled electron spins in a room-temperature solid,” Nat. Phys. 6, 249–253 (2010). [CrossRef]  

13. E. Togan, Y. Chu, A. S. Trifonov, L. Jiang, J. Maze, L. Childress, M. V. Gurudev Dutt, A. S. Sørensen, P. R. Hemmer, A. S. Zibrov, and M. D. Lukin, “Quantum entanglement between an optical photon and a solid-state spin qubit,” Nature (London) 466, 730–734 (2010). [CrossRef]  

14. A. M. Zagoskin, J. R. Johansson, S. Ashhav, and F. Nori, “Quantum information processing using frequency control of impurity spins in diamond,” Phys. Rev. B 76, 014122 (2007). [CrossRef]  

15. X. Y. Lü, Z. L. Xiang, W. Cui, J. Q. You, and F. Nori, “Quantum memory using a hybrid circuit with flux qubits and nitrogen-vacancy centers,” Phys. Rev. A 88, 012329 (2013). [CrossRef]  

16. F. Jelezko, T. Gaebel, I. Popa, A. Gruber, and J. Wrachtrup, “Observation of coherent oscillations in a single electron spin,” Phys. Rev. Lett. 92, 076401 (2004). [CrossRef]   [PubMed]  

17. R. Hanson, F. M. Mendoza, R. J. Epstein, and D. D. Awschalom, “Polarization and readout of coupled single spins in diamond,” Phys. Rev. Lett. 97, 087601 (2006). [CrossRef]   [PubMed]  

18. V. Jacques, P. Neumann, J. Beck, M. Markham, D. Twitchen, J. Meijer, F. Kaiser, G. Balasubramanian, F. Jelezko, and J. Wrachtrup, “Dynamic polarization of single nuclear spins by optical pumping of nitrogen-vacancy color centers in diamond at room temperature,” Phys. Rev. Lett. 102, 057403 (2009). [CrossRef]   [PubMed]  

19. F. Shi, X. Rong, N. Y. Xu, Y. Wang, J. Wu, B. Chong, X. H. Peng, J. Kniepert, R. S. Schoenfeld, W. Harneit, M. Feng, and J. F. Du, “Room-temperature implementation of the Deutsch-Jozsa algorithm with a single electronic spin in diamond,” Phys. Rev. Lett. 105, 040504 (2010). [CrossRef]   [PubMed]  

20. S. P. Liu, J. H. Li, R. Yu, and Y. Wu, “Achieving maximum entanglement between two nitrogen-vacancy centers coupling to a whispering-gallery-mode microresonator,” Opt. Express 21, 3501–3515 (2013). [CrossRef]   [PubMed]  

21. Y. F. Xiao, C. L. Zhou, B. B. Li, Y. Li, C. H. Dong, Z. F. Han, and Q. H. Gong, “High-Q exterior whispering-gallery modes in a metal-coated microresonator,” Phys. Rev. Lett. 105, 153902 (2010). [CrossRef]  

22. X. Yi, Y. F. Xiao, Y. C. Liu, B. B. Li, Y. L. Chen, Y. Li, and Q. H. Gong, “Multiple-Rayleigh-scatterer-induced mode splitting in a high-Q whispering-gallery-mode microresonator,” Phys. Rev. A 83, 023803 (2011). [CrossRef]  

23. G. Y. Chen, N. Lambert, C. H. Chou, Y. N. Chen, and F. Nori, “Surface plasmons in a metal nanowire coupled to colloidal quantum dots: scattering properties and quantum entanglement,” Phys. Rev. B 84, 045310 (2011). [CrossRef]  

24. J. S. Jin, C. S. Yu, P. Pei, and H. S. Song, “Positive effect of scattering strength of a microtoroidal cavity on atomic entanglement evolution,” Phys. Rev. A 81, 042309 (2010). [CrossRef]  

25. B. Peng, S. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. H. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity–time-symmetric whispering-gallery microcavities,” Nature Phys. 10, 394–398 (2014). [CrossRef]  

26. B. Peng, S. K. Özdemir, W. J. Chen, F. Nori, and L. Yang, “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nature Comms. 10, 1038 (2014).

27. B. Peng, S. K. Özdemir, S. Rotter, H. Yilmaz, M. Liertzer, F. Monifi, C. M. Bender, F. Nori, and L. Yang, “Loss-induced suppression and revival of lasing,” Science 346, 328–332 (2014). [CrossRef]   [PubMed]  

28. H. Jing, S. K. Özdemir, X. Y. Lü, J. Zhang, L. Yang, and F. Nori, “PT -Symmetric Phonon Laser,” Phys. Rev. Lett. 113, 053604 (2014). [CrossRef]  

29. T. G. McRae and W. P. Bowen, “Time-delayed entanglement from coherently coupled nonlinear cavities,” Phys. Rev. A 80, 010303 (2009). [CrossRef]  

30. Z. Q. Yin and Y. J. Han, “Generating EPR beams in a cavity optomechanical system,” Phys. Rev. A 79, 024301 (2009). [CrossRef]  

31. Y. D. Yang, Y. Z. Huang, and Q. Chen, “High-Q TM whispering-gallery modes in three-dimensional microcylinders,” Phys. Rev. A 75, 013817 (2007). [CrossRef]  

32. K. Srinivasan and O. Painter, “Mode coupling and cavity–quantum-dot interactions in a fiber-coupled microdisk cavity,” Phys. Rev. A 75, 023814 (2007). [CrossRef]  

33. P. E. Barclay, K. M. C. Fu, C. Santori, and R. G. Beausoleil, “Chip-based microcavities coupled to nitrogen-vacancy centers in single crystal diamond,” Appl. Phys. Lett. 95, 191115 (2009). [CrossRef]  

34. B. Min, E. Ostby, V. Sorger, E. U. Avila, L. Yang, X. Zhang, and K. Vahala, “High-Q surface-plasmon-polariton whispering-gallery microcavity,” Nature (London) 457, 455–458 (2009). [CrossRef]  

35. M. Larsson, K. N. Dinyari, and H. Wang, “Composite optical microcavity of diamond nanopillar and silica microsphere,” Nano Lett. 9, 1447–1450 (2009). [CrossRef]   [PubMed]  

36. T. J. Kippenberg, S. M. Spillane, and K. J. Vahala, “Modal coupling in traveling-wace resonators,” Opt. Lett. 27, 1669–1671 (2002). [CrossRef]  

37. D. S. Weiss, V. Sandoghdar, J. Hare, V. L. Seguin, J. M. Raimond, and S. Haroche, “Splitting of high-Q Mie modes induced by light backscattering in silica microspheres,” Opt. Lett. 20, 1835–1837 (1995). [CrossRef]   [PubMed]  

38. A. mazzei, S. Gotzinger, L. de S. Menezes, G. Zumofen, O. Benson, and V. Sandoghdar, “Controlled coupling of counterpropagating whispering-gallery modes by a single Rayleigh scatterer: a classical problem in a quantum optical light,” Phys. Rev. Lett. 99, 173603 (2007). [CrossRef]   [PubMed]  

39. B. B. Li, W. R. Clements, X. C. Yu, K. Shi, Q. H. Gong, and Y. F. Xiao, “Single nanoparticle detection using split-mode microcavity Raman lasers,” PNAS 111, 14657–14662 (2014). [CrossRef]   [PubMed]  

40. K. J. Vahala, “Optical microcavities,” Nature (London) 424, 839–846 (2003). [CrossRef]  

41. J. R. Buck and H. J. Kimble, “Optimal sizes of dielectric microspheres for cavity QED with strong coupling,” Phys. Rev. A 67, 033806 (2003). [CrossRef]  

42. D. V. Strekalov and N. Yu, “Generation of optical combs in a whispering gallery mode resonator from a bichromatic pump,” Phys. Rev. A 79, 041805 (2009). [CrossRef]  

43. Y. C. Liu, Y. F. Xiao, B. B. Li, X. F. Jiang, Y. Li, and Q. H. Gong, “Coupling of a single diamond nanocrystal to a whispering-gallery microcavity: photon transport benefitting from Rayleigh scattering,” Phys. Rev. A 84, 011805 (2011). [CrossRef]  

44. X. C. Yu, Y. C. Liu, M. Y. Yan, W. L. Jin, and Y. F. Xiao, “Coupling of diamond nanocrystals to a high-Q whispering-gallery microresonator,” Phys. Rev. A 86, 043833 (2012). [CrossRef]  

45. C. H. Su, A. D. Greentree, W. J. Munro, K. Nemoto, and L. C. L. Hollenberg, “High-speed quantum gates with cavity quantum electrodynamics,” Phys. Rev. A 78, 062336 (2008). [CrossRef]  

46. C. H. Su, A. D. Greentree, and L. C. L. Hollenberg, “High-performance diamond-based single-photon sources for quantum communication,” Phys. Rev. A 80, 052308 (2009). [CrossRef]  

47. C. H. Su, A. D. Greentree, and L. C. L. Hollenberg, “Towards a picosecond transform-limited nitrogen-vacancy based single photon,” Opt. Express 16, 6240–6250 (2008). [CrossRef]   [PubMed]  

48. W. L. Yang, Z. Q. Yin, Z. Y. Xu, M. Feng, and J. F. Du, “One-step implementation of multiqubit conditional phase gating with nitrogen-vacancy centers coupled to a high-Q silica microsphere cavity,” Appl. Phys. Lett. 96, 241113 (2010). [CrossRef]  

49. W. L. Yang, Z. Y. Xu, M. Feng, and J. F. Du, “Entanglement of separate nitrogen-vacancy centers coupled to a whispering-gallery mode cavity,” New J. Phys. 12, 113039 (2010). [CrossRef]  

50. P. B. Li, S. Y. Gao, and F. L. Li, “Quantum-information transfer with nitrogen-vacancy centers coupled to a whispering-gallery microresonator,” Phys. Rev. A 83, 054306 (2011). [CrossRef]  

51. P. B. Li, S. Y. Gao, H. R. Li, S. L. Ma, and F. L. Li, “Dissipative preparation of entangled states between two spatially separated nitrogen-vacancy centers,” Phys. Rev. A 85, 042306 (2012). [CrossRef]  

52. J. Wolters, J. Kabuss, A. Knorr, and O. Benson, “Deterministic and robust entanglement of nitrogen-vacancy centers using low-Q photonic-crystal cavities,” Phys. Rev. A 89, 060303 (2014). [CrossRef]  

53. S. C. Benjamin, B. W. Lovett, and J. M. Smith, “Prospects for measurement-based quantum computing with solid state spins,” Laser Photonics Rev. 3, 556–574 (2009). [CrossRef]  

54. Q. Chen, W. L. Yang, M. Feng, and J. F. Du, “Entangling separate nitrogen-vacancy centers in a scalable fashion via coupling to microtoroidal resonators,” Phys. Rev. A 83, 054305 (2011). [CrossRef]  

55. W. L. Yang, Z. Q. Yin, Z. Y. Xu, M. Feng, and C. H. Oh, “Quantum dynamics and quantum state transfer between separated nitrogen-vacancy centers embedded in photonic crystal cavities,” Phys. Rev. A 84, 043849 (2011). [CrossRef]  

56. W. L. Yang, J. H. An, C. J. Zhang, M. Feng, and C. H. Oh, “Preservation of quantum correlation between separated nitrogen-vacancy centers embedded in photonic-crystal cavities,” Phys. Rev. A 87, 022312 (2013). [CrossRef]  

57. Y. S. Park, A. K. Cook, and H. Wang, “Cavity QED with diamond nanocrystals and silica microspheres,” Nano Lett. 6, 2075–2079 (2006). [CrossRef]   [PubMed]  

58. S. Schietinger, Y. Schröder, and O. Benson, “One-by-one coupling of single defect centers in nanodiamonds to high-Q modes of an optical microresonator,” Nano Lett. 8, 3911–3915 (2008). [CrossRef]   [PubMed]  

59. P. E. Barclay, C. Santori, K. M. Fu, R. G. Beausoleil, and O. Painter, “Coherent interference effects in a nanoassembled diamond NV center cavity-QED system,” Opt. Express 17, 8081–8097 (2009). [CrossRef]   [PubMed]  

60. B. J. M. Hausmann, B. Shields, Q. M. Quan, P. Maletinsky, M. McCutcheon, J. F. Choy, T. M. Babinec, A. Kubanek, A. Yacoby, M. D. Lukin, and M. Lončar, “Integrated diamond networks for quantum nanophotonics,” Nano Lett. 12, 1578–1582 (2012). [CrossRef]   [PubMed]  

61. N. B. An, J. Kim, and K. Kim, “Nonperturbative analysis of entanglement dynamics and control for three qubits in a common lossy cavity,” Phys. Rev. A 82, 032316 (2010). [CrossRef]  

62. N. B. An, J. Kim, and K. Kim, “Entanglement dynamics of three interacting two-level atoms within a common structured environment,” Phys. Rev. A 84, 022329 (2011). [CrossRef]  

63. C. Eltschka, D. Braun, and J. Siewert, “Heat bath can generate all classes of three-qubit entanglement,” Phys. Rev. A 89, 062307 (2014). [CrossRef]  

64. J. I. de Vicente, T. Carle, C. Streitberger, and B. Kraus, “Complete set of operational measures for the characterization of three-qubit entanglement,” Phys. Rev. Lett. 108, 060501 (2012). [CrossRef]   [PubMed]  

65. S. Armstrong, M. Wang, R. Y. Teh, Q. H. Gong, Q. Y. He, J. Janousek, H. A. Bachor, M. D. Reid, and P. K. Lam, “Multipartite Einstein–Podolsky–Rosen steering and genuine tripartite entanglement with optical networks,” Nature Phys. 11, 167–172 (2015). [CrossRef]  

66. Q. Y. He and Z. Ficek, “Einstein-Podolsky-Rosen paradox and quantum steering in a three-mode optomechanical system,” Phys. Rev. A 89, 022332 (2014). [CrossRef]  

67. M. Wang, Q. H. Gong, Z. Ficek, and Q. Y. He, “Role of thermal noise in tripartite quantum steering,” Phys. Rev. A 90, 023801 (2014). [CrossRef]  

68. H. Jing, X. J. Liu, M. L. Ge, and M. S. Zhan, “Correlated quantum memory: Manipulating atomic entanglement via electromagnetically induced transparency,” Phys. Rev. A 71, 062336 (2005). [CrossRef]  

69. H. Jing, Y. G. Deng, and W. P. Zhang, “Quantum control of light through an atom-molecule dark state,” Phys. Rev. A 80, 025601 (2009). [CrossRef]  

70. P. Rabl, S. J. Kolkowitz, F. H. L. Koppens, J. G. E. Harris, P. Zoller, and M. D. Lukin, “A quantum spin transducer based on nanoelectromechanical resonator arrays,” Nat. Phys. 6, 602–608 (2010). [CrossRef]  

71. M. Li, S. M. Fei, and Z. X. Wang, “A lower bound of concurrence for multipartite quantum states,” J. Phys. A 42, 145303 (2009). [CrossRef]  

72. H. J. Briegel and B. G. Englert, “Quantum optical master equations: the use of damping bases,” Phys. Rev. A 47, 3311–3329 (1993). [CrossRef]   [PubMed]  

73. M. Scala, B. Militello, A. Messina, J. Piilo, and S. Maniscalco, “Microscopic derivation of the Jaynes-Cummings model with cavity losses,” Phys. Rev. A 75, 013811 (2007). [CrossRef]  

74. M. Wilczewski and M. Czachor, “Theory versus experiment for vacuum Rabi oscillations in lossy cavities,” Phys. Rev. A 79, 033836 (2009). [CrossRef]  

75. V. Eremeev, V. Montenegro, and M. Orszag, “Thermally generated long-lived quantum correlations for two atoms trapped in fiber-coupled cavities,” Phys. Rev. A 85, 032315 (2012). [CrossRef]  

76. R. Coto and M. Orszag, “Determination of the maximum global quantum discord via measurements of excitations in a cavity QED network,” J. Phys. B 47, 095501 (2014). [CrossRef]  

77. E. B. Davies, “Markovian master equations,” Commun. Math. Phys. 39, 91–110 (1974). [CrossRef]  

78. E. B. Davies, Quantum Theory of Open Systems (Academic, 1976).

79. H. P. Breuer and F. Petruccione, The Theory of Open Quantum Systems (Clarendon, Oxford, 2006).

80. J. Merlein, M. Kahl, A. Zuschag, A. Sell, A. Halm, J. Boneberg, P. Leiderer, A. Leitenstorfer, and R. Bratschisch, “Nanomechanical control of an optical antenna,” Nature Photonics 2, 230–233 (2008). [CrossRef]  

81. D. Ratchford, F. Shafiei, S. Kim, S. K. Gray, and X. Q. Li, “Manipulating coupling between a single semiconductor quantum dot and single cold nanoparticle,” Nano Lett. 11, 1049–1054 (2011). [CrossRef]   [PubMed]  

82. N. B. Manson, J. P. Harrison, and M. J. Sellars, “Nitrogen-vacancy center in diamond: model of the electronic structure and associated dynamics,” Phys. Rev. B 74, 104303 (2006). [CrossRef]  

83. X. H. Gao, S. M. Fei, and K. Wu, “Lower bounds of concurrence for tripartite quantum systems,” Phys. Rev. A 74, 050303 (2006). [CrossRef]  

84. M. Siomau and S. Fritzsche, “Entanglement dynamics of three-qubit states in noisy channels,” Eur. Phys. J. D 60, 397–403 (2010). [CrossRef]  

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

86. A. Uhlmann, “Entropy and optimal decompositions of states relative to a maximal commutative subalgebra,” Open Syst. Inf. Dyn. 5, 209–228 (1998). [CrossRef]  

87. Z. Q. Ma and X. Y. Gu, Problems and Solutions in Group Theory for Physicists (World Scientific, Singapore, 2004).

88. C. Santori, P. Tamarat, P. Neumann, J. Wrachtrup, D. Fattal, R. G. Beausoleil, J. Rabeau, P. Olivero, A. D. Greentree, S. Prawer, F. Jelezko, and P. Hemmer, “Coherent population trapping of single spins in diamond under optical excitation,” Phys. Rev. Lett. 97, 247401 (2006). [CrossRef]  

89. T. Gaebel, M. Domhan, I. Popa, C. Wittmann, P. Neumann, F. Jelezko, J. R. Rabeau, N. Stavrias, A. D. Greentree, S. Prawer, J. Twamley, P. Hemmer, and J. Wrachtrup, “Room-temperature coherent coupling of single spins in diamond,” Nat. Phys. 2, 408–413 (2006). [CrossRef]  

90. G. Balasubramanian, P. Neumann, D. Twitchen, M. Markham, R. Kolesov, N. Mizuochi, J. Isoya, J. Achard, J. Beck, J. Tissler, V. Jacques, P. R. Hemmer, F. Jelezko, and J. Wrachtrup, “Ultralong spin coherence time in isotopically engineered diamond,” Nat. Mater. 8, 383–387 (2009). [CrossRef]   [PubMed]  

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 (8)

Fig. 1
Fig. 1 (a) Schematic of the system consisting of three NVCs and a microtoroidal cavity supporting two counter-propagating WGMs acw and accw. (b) Level diagram for the j-th NVC, where Δj is the frequency detuning of the NVC from the cavity. Deg = γeB0 is the level splitting induced by an external magnetic field B0 with γe the electron gyromagnetic ratio.
Fig. 2
Fig. 2 The LBC vs time t and the phase difference θ13 when the system is initially prepared in the state ψ(0). (a) θ 12 = π 2 + n π, (b) θ12 = . The time parameter is dimensionless and we set G = 1. Other parameters are Δ = 0, g = 0.1, Γ = 0.05, γ = 0.01 and κ = 0.03.
Fig. 3
Fig. 3 (a) The maximal values of the LBC in the parameter plane of {θ12, θ13} with Δ = 0, G = 1, g = 0.1, Γ = 0.05, γ = 0.01 and κ = 0.03, where the system is initially prepared in the state ψ(0). (b) The zooming-in plot of (a) with one period.
Fig. 4
Fig. 4 The LBC vs time t and the coupling strength g for different initial states of the system (a) ψ(0) and (b) Ψ(0). Other parameters are θ 12 = π 4 + n π, θ 13 = π 8 + n π, Δ = 0, G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03. In the bottom six subgraphs, the curves of the LBC dynamics are plotted in (a1,b1) g = 0.1, (a2,b2) g = 1, (a3,b3) g = 5.
Fig. 5
Fig. 5 The LBC vs time t and the phase difference θ13 under different detunings (a) Δ = 5 and (c) Δ = 10, where θ12 = , G = 1, g = 0.1, Γ = 0.05, γ = 0.01 and κ = 0.03. The LBC vs time t and the coupling strength g under different detunings (b) Δ = 5 and (d) Δ = 10, where θ 12 = π 4 π, θ 13 = π 8 + n π, G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03.
Fig. 6
Fig. 6 The LBC vs time t and the detuning Δ for different scattering strengths (a) g = 0.1, (b) g = 1, (c) g = 5. The other parameters are θ 12 = π 4 + n π, θ 13 = π 8 + n π, G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03.
Fig. 7
Fig. 7 The maximal LBC vs the coupling strength g and the detuning Δ in (a) θ12 = θ13 = , (b) θ 12 = π 2 + n π and θ 13 = π 4 + n π, (c) θ 12 = π 4 + n π and θ 13 = π 8 + n π, (d) θ 12 = 3 π 4 + n π and θ 13 = 3 π 8 + n π. Other parameters are G = 1, Γ = 0.05, γ = 0.01 and κ = 0.03.
Fig. 8
Fig. 8 (a) The LBC vs time t under the condition θ 12 = π 4, θ 13 = π 8, Δ = 0, g = 2, where the red-solid (blue-dashed) lines denote the analytical (numerical) results, respectively. (b) The slight difference between the LBC dynamics using these two different methods.

Equations (24)

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

H = H 0 + H 1 + H 2 , H 0 = j = 1 3 ω 0 2 σ j z + ω c a c w a c w + ω c a c c w a c c w , H 1 = j = 1 3 G j ( e i k d 1 j σ j + a c w + e i k d 1 j σ j + a c c w ) + H . C . , H 2 = j = 1 3 g j ( a c w a c w + a c c w a c c w ) + j = 1 3 g j ( e 2 i k d 1 j a c c w a c w + H . C ) . ,
C 3 ( | φ ) = 1 3 [ 3 Tr ( ρ 1 2 ) Tr ( ρ 2 2 ) Tr ( ρ 3 2 ) ] ,
C 3 ( ρ ) = min { p i , φ i } i p i C 3 ( | φ i )
C ¯ 3 ( ρ ) = 1 3 j = 1 6 [ ( C j 12 | 3 ( ρ ) ) 2 + ( C j 31 | 2 ( ρ ) ) 2 + ( C j 23 | 1 ( ρ ) ) 2 ] ,
C j k l | m ( ρ ) = max { 0 , λ j , 1 k l | m n > 1 λ j , n k l | m } ,
C ¯ 3 ( ρ ) a n a = max { 0 , 8 5 ( ρ e g g , e g g ρ g e g , g e g + ρ e g g , e g g ρ g g e , g g e + ρ g e g , g e g ρ g g e , g g e ) } .
H = j = 1 3 [ ω 0 σ e j e j + δ 1 a 1 a 1 + δ 2 a 2 a 2 + G ˜ 1 j ( a 1 σ j + a 1 σ j + ) ] + j = 2 3 G ˜ 2 j ( a 2 σ j + a 2 σ j + ) + g ˜ ( a 1 a 2 a 2 a 1 )
1 { | 1 = | e 1 g 2 g 3 0 c w 0 c c w , | 2 = | g 1 e 2 g 3 0 c w 0 c c w , | 3 = | g 1 g 2 e 3 0 c w 0 c c w , | 4 = | g 1 g 2 g 3 1 c w 0 c c w , | 5 = | g 1 g 2 g 3 0 c w 1 c c w } ,
2 { | 6 = | g 1 g 2 g 3 0 c w 0 c c w } ,
H = ( ω 0 2 0 0 G G 0 0 ω 0 2 0 e i k d 12 G e i k d 12 G 0 0 0 ω 0 2 e i k d 13 G e i k d 13 G 0 G e i k d 12 G e i k d 13 G 3 ω 0 2 + ω c + 3 g g ( 1 + e 2 i k d 12 + e 2 i k d 13 ) 0 G e i k d 12 G e i k d 13 G g ( 1 + e 2 i k d 12 + e 2 i k d 13 ) 3 ω 0 2 + ω c + 3 g 0 0 0 0 0 0 3 ω 0 2 )
ρ ˙ ( t ) = i [ H , ρ ] + ω ¯ > 0 , m = { c w , c c w } κ m ( ω ¯ ) × [ A m ( ω ¯ ) ρ ( t ) A m ( ω ¯ ) 1 2 { A m ( ω ¯ ) A m ( ω ¯ ) , ρ ( t ) } ] + ω ¯ > 0 , n = { 1 , 2 , 3 } Γ n ( ω ¯ ) × [ n ( ω ¯ ) ρ ( t ) n + ( ω ¯ ) 1 2 { n + ( ω ¯ ) n ( ω ¯ ) , ρ ( t ) } ] + ω ¯ > 0 , n = { 1 , 2 , 3 } γ n ( ω ¯ ) × [ n z ( ω ¯ ) ρ ( t ) n z ( ω ¯ ) 1 2 { n z ( ω ¯ ) n z ( ω ¯ ) , ρ ( t ) } ] ,
ρ ˙ ( t ) = i [ H , ρ ] + i = 1 5 Γ ˜ i [ | ϕ 6 ϕ i | ρ ( t ) | ϕ i ϕ 6 | 1 2 { | ϕ i ϕ i | , ρ ( t ) } ] + i = 1 5 γ ˜ i [ ( | ϕ i ϕ i | | ϕ 6 ϕ 6 | ) ρ ( t ) ( | ϕ i ϕ i | | ϕ 6 ϕ 6 | ) 1 2 { ( | ϕ i ϕ i | + | ϕ 6 ϕ 6 | ) , ρ ( t ) } ] ,
U ˜ = ( c 11 c 12 c 13 c 14 c 15 0 c 21 c 22 c 23 c 24 c 25 0 c 31 c 32 c 33 c 34 c 35 0 c 41 c 42 c 43 c 44 c 45 0 c 51 c 52 c 53 c 54 c 55 0 0 0 0 0 0 1 ) .
ρ ˙ = i [ H e f f , ρ ] + ρ ,
ρ = m = c w , c c w κ m ( a m ρ a m + 1 2 a m + a m ρ 1 2 ρ a m + a m ) + j = 1 3 Γ j ( σ j ρ σ j + 1 2 σ j + σ j ρ 1 2 ρ σ j + σ j ) + j = 1 3 γ j ( σ j z ρ σ j z 1 2 σ j z σ j z ρ 1 2 ρ σ j z σ j z ) .
H e f f = j = 1 3 [ e i Δ t G j ( e i k d 1 j σ j + a c w + e i k d 1 j σ j + a c c w ) + g j e 2 i k d 1 j a c c w a c c w + H . C ] + j = 1 3 g j ( a c w a c w + a c c w a c c w ) ,
c 11 = e i k d 13 ( e 2 i k d 13 e 2 i k d 12 ) / ( e 2 i k d 12 1 ) , c 12 = e i k d 12 i k d 13 ( 1 e 2 i k d 13 ) / ( e 2 i k d 12 1 ) , c i 1 = G { ξ [ ( e 2 i k d 12 + e 2 i k d 13 2 ) η + Δ A i 1 ] + A i 1 2 } q A i 1 , c i 2 = e i k d 12 G { ξ [ ( e 2 i k d 12 + e 2 i k d 13 2 ) η + Δ A i 1 ] + A i 1 2 } q A i 1 , c i 3 = e i k d 13 G { ξ [ ( e 2 i k d 13 + e 2 i k d 12 2 ) η + Δ A i 1 ] + A i 1 2 } q A i 1 , c i 4 = { ξ [ 3 η Δ A i 1 ] A i 1 2 } / ξ q , c 14 = c 15 = c 16 = c i 6 = 0 , c 13 = c i 5 = 1 ,
x 4 + a x 3 + b x 2 + c x + d = 0 ,
A c w ( λ i λ 6 ) = c i 4 | ϕ 6 ϕ i | , A c c w ( λ i λ 6 ) = c i 5 | ϕ 6 ϕ i | ,
1 ( λ i λ 6 ) = c i 1 | ϕ 6 ϕ i | , 2 ( λ i λ 6 ) = c i 2 | ϕ 6 ϕ i | , 3 ( λ i λ 6 ) = c i 3 | ϕ 6 ϕ i | .
Γ ˜ i = j = 1 3 | c i j | 2 Γ j ( λ i λ 6 ) + | c i 4 | 2 γ c w ( λ i λ 6 ) + | c i 5 | 2 γ c c w ( λ i λ 6 ) , γ ˜ i = j = 1 3 | c i j | 4 γ j ( λ j λ 6 ) .
ρ i i ( t ) = ρ i i ( 0 ) e Γ ˜ i t , ρ i j ( t ) = ρ i j ( 0 ) e [ 2 i ( λ i λ j ) ( Γ ˜ i + Γ ˜ j + γ ˜ i + γ ˜ j ) ] t / 2 , ρ i 6 ( t ) = ρ i 6 ( 0 ) e [ i ( λ i λ 6 ) Γ ˜ i 2 j = 1 5 ( γ ˜ j + 3 γ ˜ i ) 2 ] t , ρ 66 ( t ) = j = 1 5 ρ j j ( 0 ) ( 1 e Γ ˜ j t ) + ρ 66 ( 0 ) .
ρ a n a ( t ) = i , j = 1 6 ϕ i | ρ ( t ) | ϕ j 00 | ϕ i ϕ j | 00 + 10 | ϕ i ϕ j | 10 + 01 | ϕ i ϕ j | 01 + 11 | ϕ i ϕ j | 11 ] ,
ρ a n a ( t ) = ρ e g g , g e g | e 1 g 2 g 3 g 1 e 2 g 3 | + ρ e g g , g g e | e 1 g 2 g 3 g 1 g 2 e 3 | + ρ e g g , g g g | e 1 g 2 g 3 g 1 g 2 g 3 | + ρ g e g , e g g | g 1 e 2 g 3 e 1 g 2 g 3 | + ρ g e g , g g e | g 1 e 2 g 3 g 1 g 2 e 3 | + ρ g e g , g g g | g 1 e 2 g 3 g 1 g 2 g 3 | + ρ g g e , e g g | g 1 g 2 e 3 e 1 g 2 g 3 | + ρ g g e , g e g | g 1 g 2 e 3 g 1 e 2 g 3 | + ρ g g e , g g g | g 1 g 2 e 3 g 1 g 2 g 3 | + ρ g g g , e g g | g 1 g 2 g 3 e 1 g 2 g 3 | + ρ g g g , g e g | g 1 g 2 g 3 g 1 e 2 g 3 | + ρ g g g , g g e | g 1 g 2 g 3 g 1 g 2 e 3 | + ρ e g g , e g g | e 1 g 2 g 3 e 1 g 2 g 3 | + ρ g e g , g e g | g 1 e 2 g 3 g 1 e 2 g 3 | + ρ g g e , g g e | g 1 g 2 e 3 g 1 g 2 e 3 | + ρ g g g , g g g | g 1 g 2 g 3 g 1 g 2 g 3 | ,
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.