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

Collective dynamics and entanglement of two distant atoms embedded into single-negative index material

Open Access Open Access

Abstract

We study the dynamics of two two-level atoms embedded near to the interface of paired meta-material slabs, one of negative permeability and the other of negative permittivity. This combination generates a strong surface plasmon field at the interface between the meta-materials. It is found that the symmetric and antisymmetric modes of the two-atom system couple to the plasmonic field with different Rabi frequencies. Including the Ohmic losses of the materials we find that the Rabi frequencies exhibit threshold behaviour which distinguish between the non-Markovian (memory preserving) and Markovian (memoryless) regimes of the evolution. Moreover, it is found that significantly different dynamics occur for the resonant and an off-resonant couplings of the plasmon field to the atoms. In the case of the resonant coupling, the field does not appear as a dissipative reservoir to the atoms. We adopt the image method and show that the dynamics of the two atoms coupled to the plasmon field are analogous to the dynamics of a four-atom system in a rectangular configuration. A large and long living entanglement mediated by the plasmonic field in both Markovian and non-Markovian regimes of the evolution is predicted. We also show that a simultaneous Markovian and non-Markovian regime of the evolution may occur in which the memory effects exist over a finite evolution time. In the case of an off-resonant coupling of the atoms to the plasmon field, the atoms interact with each other by exchanging virtual photons which results in the dynamics corresponding to those of two atoms coupled to a common reservoir. In addition, the entanglement is significantly enhanced.

© 2017 Optical Society of America

1. Introduction

The radiative properties of emitters (e.g. atoms or quantum dots) located inside a dielectric or conducting material can be significantly modified compared to those in vacuum. The modification results from the variation of the density of modes of the EM field which can be adjusted by changing the geometric shape or space period structure of the material [1–3]. The radiative properties of an emitter can also be modified by locating the emitter close to the surface of a dielectric or conducting material [4–9]. In this case, the modification originates from the presence of electromagnetic modes propagating along the surface, called surface plasmon polaritons, or shortly surface plasmons (SP) [10–17]. A new category of materials has been proposed, so-called meta-materials [18–21], characterized by specifically designed geometrical structures which drastically modify the density of the EM modes, so the field propagation, also yielding to the SP field [22–27]. Owing to the high local density of modes, the radiative properties of the emitters can suffer a sharply changed when couple to the surface field [28–34], which has potential applications and attracts a great deal of attention. For example, when quantum dots are placed at a distance about several tens of nanometers above two dimensional metal surface, strong coupling generated between the quantum dots and the collective mode would results in the Rabi oscillation [35]. In the structure composed of zero index and left hand materials, maximum quantum interference and a suppression of the atomic decay rate can be achieved between Zeeman levels [36] due to the anisotropy of the EM modes [37]. It has been shown that by changing the strength of the driving field and adjusting position of the quantum dot, the plasmon modes on the surface of a metal-nanoparticles material can introduce asymmetrical features [38] into the fluorescence spectrum [39].

In this paper, we consider the dynamics of two independent atoms coupled to a SP field generated at the interface of two meta-materials, one of a negative permeability (MN) and the other of a negative permittivity (EN). Our focus is on how the plasmon field induced at the interface between the materials changes the dynamics of the atoms, in particular, the population transfer and entanglement. The mathematical approach we adopt here is based on the Green’s function method [40]. The remarkably simple analytical expressions are derived for the probability amplitudes valid for an arbitrary initial state, arbitrary strengths of the coupling constants of the atoms to the plasmon field, and arbitrary distances between the atoms. We find a number of interesting general results in both strong and weak coupling regimes of the atoms to the SP field. In particular, we find a threshold behavior of the coupling constants which separate the non-Markovian behavior of the system from the Markovian one [41]. The Markovian evolution is usually attributed to a weak coupling of an atom to the field. We show that the collective effects may result in the Markovian evolution even in the limit of a strong coupling of the atoms to the field. Inversely, a non-Markovian evolution can be seen even in the regime of a weak coupling.

The system which we discuss here is closely related to systems, for example, involving photonic crystal waveguides [42–45], nanorod resonators [46], nano-wires [47], metal nanoparticles [48], a thin membrane made of a negative left handed material [49], and a quantum dotgraphene structure [50]. In these systems the coupling between emitters is also mediated by surface polaritons, but the calculations are not specifically oriented towards studying the strong coupling effects. The calculations are limited to the weak coupling regime resulting in a Markovian dynamics of the emitters. In this case, the incoherent spontaneous exchange of photons may occur resulting in collective damping of the emitters [51, 52]. The spontaneous exchange of the excitation can lead to entanglement between the emitters which may exist over distances much larger than the resonant wavelength. The study of entanglement between distant atoms and controlled the transmission of information between them are vital to the development of quantum information technology [53, 54].

The plan of this paper is as follows. In Sec. 2 we introduce the model and present the explicit analytic expressions for the time dependence of the probability amplitudes of the symmetric and antisymmetric combinations of the probability amplitudes of the atoms. We assume that a single excitation is present initially in the system and demonstrate how the evolution of the system can be simply understood in terms of the evolution of the atoms and their corresponding images. We then demonstrate in Sec. 3 the collective behavior of the atoms in both Markovian and non-Markovian regimes of the evolution. Entangled properties of the atoms are discussed in Sec. 4, where we calculate the concurrence for different initial states and different coupling strengths of the atoms to the plasmon field. The effect of an off-resonant coupling of the atoms to the plasmon field on the collective dynamics and entanglement is discussed in Sec. 5. We summarize our results in Sec. 6. The paper concludes with an Appendix in which we give details of the derivation of the integro-differential equations for the probability amplitudes and the calculations of the integral kernels. Both longitudinal and transverse parts of the Green function are considered in the evaluation of the kernels.

2. Atoms interacting with the plasmon field

Perfectly conducting materials are known to generate a strong SP field which is refined in a short regime near the surface [55–57]. The plasmon field can also be produced at the surface of a meta-material with either negative permittivity (ε < 0) or negative permeability (μ < 0). However, the density of the SP field only originates from one or several discrete modes, which can be derived by applying the continuous conditions on the boundaries. Recently, Tan et al. [58] have shown that the density of the EM modes can be significantly enhanced at the interface between two meta-materials, one with negative ε and the other with negative μ. Especially, when the materials are perfectly paired, i.e. ε1 = −ε2 and μ1 = −μ2, the effective permittivity and permeability, defined as

εr=(d1ε1+d2ε2)/(d1+d2),μr=(d1μ1+d2μ2)/(d1+d2)
are both zero when the slabs have the same thickness. Hence, the structure such constructed can be treated as the zero-index meta-material [59]. In this case, the band gap disappears and the density of modes of the EM field becomes continuous. Consequently, a large density of the modes exists at the interface between the two perfectly paired negative meta-material slabs, which can be treated as optical topological material [60].

In this paper, we consider a system composed of two identical atoms located at a distance z0 from the interface between two different negative index material slabs, μ-negative (MN) and ε-negative (EN) slabs, as shown in Fig. 1. For a potential experimental realization of a paired zero-index materials one could consider a structure fabricated by the composite right/left-handed transmission line, in which both the effective permittivity and permeability are zero in the microwave region (ω ~ 2.0 GHz) [61]. An impedance-matched zero index material of resonant frequencies ranging from optical to near infrared region (ω ~ 200 THz), can also be realized by using purely dielectric high index rods or cut wire pairs [62, 63].

We assume that the atoms are located in the MN slab and the distance between the atoms is large so the direct interaction between the atoms is weak and can be ignored, especially when compared to the coupling strength with the plasmon field. Each atom is represented by its ground state |gl〉 and excited state |el〉 (l = 1, 2), the atomic transition frequency ωa, and the atomic transition dipole moment pl. The atoms interact with an electromagnetic field via a dipole interaction according to the Hamiltonian [64]

H^=H^0+H^I,
where
H^0=12ωa(σ^z1+σ^z2)+λ=e,mdr0dωωf^λ(r,ω)f^λ(r,ω)
is the unperturbed Hamiltonian of the atoms and the field, and
H^I=l=1,2[pl0dωE^(+)(rl,ω)σ^l+H.c.]
is the interaction of the atoms with the field. Here, f^λ(r,ω) and f^λ(r,ω) are the creation and annihilation operators which can be viewed as collective excitations of the electromagnetic field, λ = e represents noise polarization of the EN material, λ = m represents noise magnetization of the MN material [9], σ^l(σ^l) and σ^z are the raising (lowering) and the energy difference operators of atom l. The positive frequency part of the electric field operator at the position rl of the lth atom is given by
E^(+)(rl,ω)=iπε0ωcdr{ωc[ε(r,ω)]G(rl,r,ω)f^e(r,ω)+[κ(r,ω)]×G(rl,r,ω)f^m(r,ω)},
where ℑ[ε(r, ω)] is the imaginary part of permittivity, ℑ[κ(r, ω)] is the imaginary part of reciprocal of permeability (κ(r, ω) = 1(r, ω)), respectively, and G(rl,r,ω) is the Green function of the field, which characterizes the density of the field modes at the location rl.

 figure: Fig. 1

Fig. 1 (Color online) Schematic illustration of the system showing the electric field (yellow arrows) of an electromagnetic wave and dislocated surface charges at the interface between MN and EN slabs. The z axis is taken normal to the interface with its origin at the interface. The slabs have thickness d1 and d2, respectively, and are assumed to have infinite extents in the xy plane. Two atoms are embedded in the MN slab at fixed positions (x1, 0, z0) and (x2, 0, z0), where z0 is the distance of the atoms from the interface between the materials. The atomic transition dipole moments p1 and p2 are parallel to each other and oriented in the xz plane.

Download Full Size | PDF

For the permittivity and permeability of the slabs, we assume that ε1 and μ2 are positive constants, but ε2 and μ1 are negative and strongly depend on frequency of electromagnetic field, and are given by the following expressions [59, 65]

ε2(ω)ε0=1+ωep2ωeo 2ω2iωγe,μ1(ω)μ0=1+ωmp2ωmo 2ω2iωγm,
where ωep and ωmp are plasmon frequencies of the electric and magnetic materials, respectively, ωeo and ωmo are resonance frequencies of the materials, and the parameters γe and γm represent Ohmic losses of the materials. For clarity of the notation we have omitted the spatial argument. It is clear from Eq. (6) that in the frequency region above the resonance, ω > ωeo (ω > ωmo) and ω < ωep (ω < ωmp), the EN (MN) slab is a single-negative material. Thus, the structure of a meta-material can be designed [66, 67].

If the field was initially at t = 0 in the vacuum state and the atoms shared a single excitation, the wave function of the system at time t > 0, written in the interaction picture is of the form

|Ψ(t)=C1(t)eiωat|{0}|e1,g2+C2(t)eiωat|{0}|g1,e2+λ=e,m0dωeiωtdrCλ(r,ω,t)|1λ(r,ω)|g1,g2,
where C1(t) is the probability amplitude of the state in which atom 1 is in its excited state |e1〉, atom 2 is in the ground state |g2〉, and the field is in the vacuum state |{0}〉. C2(t) is the probability amplitude of the state in which atom 1 is in its ground state |g1〉, atom 2 is in the excited state |e2〉, and the field is in the vacuum state |{0}〉. Cλ (r, ω, t) is the probability amplitude of the state in which both atoms are in their ground states, |g1〉, |g2〉, and there is an excitation of the medium-assisted field |1λ(r,ω)f^λ(r,ω)|{0}.

To study the dynamics of the atoms, we consider the wave function of the system whose the time evolution is governed by the Schrödinger equation. We may introduce symmetric and antisymmetric combinations of the probability amplitudes, Cs=(C1+C2)/2 and Ca=(C1C2)/2, corresponding to collective symmetric and antisymmetric states of the two-atom system. A straightforward but lengthly calculation (for details see Appendix) leads to the following equations of motion

C˙s(t)=0tdtKs(t,t)Cs(t),C˙a(t)=0tdtKa(t,t)Ca(t),
where
Kj(t,t)=Ωj2e(12γ+iδ)(tt),j=s,a.

Here, γγe = γm represents Ohmic losses in the materials, δ = ωsωa is the detuning of the atomic transition frequency from the plasmon field frequency, Ωs2=Ω02[1+U(x21,z0)] and Ωa2=Ω02[1U(x21,z0)] are coupling strengths (Rabi frequencies) of the symmetric and antisymmetric states to the plasmon field in which

Ω0={[3+4π2|[μ1(ωs)]|(2z0/λs)2]ωsΓA64π3(2z0/λs)3}1/2
is the coupling strength of the atoms to the surface plasmon field, and
U(x21,z0)=F[12,1,2;x212(2z0)2]+13+4π2|[μ1(ωs)]|(2z0/λs)2{F[32,2,2;x212(2z0)2]+2F[32,2,1;x212(2z0)2]3F[12,1,2;x212(2z0)2]3x212(2z0)2F[52,3,3;x212(2z0)2]}
determines the strength of the interaction between the atoms resulting from the coupling of the atoms with the same plasmon field. In Eq. (10), ΓA=ωa3pa2/(3ε0πc3) is the spontaneous emission rate of the atoms in free space, assumed the atoms are identical, pap1 = p2. The damping rate γγe = γm which appears in Eq. (9) represents the Ohmic losses of the materials. In a typical negative index material the Ohmic losses are small compared to the plasmon frequency, γ ≈ (10−4 ∼ 10−3)ωep [59, 68].

The function U(x21, z0) depends on the separation between the atoms, x21, and also their distance z0 from the interface. It determines the strength of the coupling between the atoms. In the limit of x21λs, U(x21, z0) ≈ 0, while for x21λs, U(x21, z0) ≈ 1. Thus, for large x21, the effects of the coupling between the atoms become negligible and the atoms evolve independently. Notice that at x21 = 0 the function U(x21, z0) is always unity independent of the value of z0, the distance of the atoms from the interface. However, the variation of U(x21, z0) with x21 depends strongly on z0. This is illustrated in Fig. 2, which shows U(x21, z0) as a function of x21 for several different values of z0. It is seen that for large z0, the function U(x21, z0) varies slowly with x21. In that case, the indirect coupling between the atoms which is provided by the plasmon field is effectively quite strong even at large separations. On the other hand, for small z0, (z0λs), the function U(x21, z0) is different from zero only over very small distances x21 and decays rapidly to zero as x21 increases. Thus, a strong coupling of the atoms to the plasmon field destroys the collective behavior of the atoms. In the physical terms, the location of the atoms at a small distance z0 from the interface leads to a strong spatial confinement (localization) of the surface plasmon fields around x1 and x2 resulting in a weak overlap of the surface fields excited by the atoms.

 figure: Fig. 2

Fig. 2 Dependence of U(x21, z0) on separation between the atoms x21s for several different distances of the atoms from the interface: z0 = 0.05λs (solid black line), z0 = 0.1λs (dashed red line), z0 = 0.25λs (dashed-dotted blue line), and z0 = 0.5λs (solid green line).

Download Full Size | PDF

From Eq. (9) it follows that the kernel Kj (t, t′) is a function only of the time difference tt′. Therefore, Eq. (8) can be solved exactly by Laplace transformation, and the solution is

Cj(t)=12Cj(0)e12iδt[(1+ueiθΩ˜j)e(14γΩ˜j)t+(1ueiθΩ˜j)e(14γ+Ω˜j)t],j=s,a,
where Ω˜j=(ueiθ)2Ωj2 is the effective Rabi frequency of the coupling of mode j to the SP field, u=γ2+4δ2/4 and θ = arctan (2δ/γ).

The result (12) shows a number of interesting features. Firstly, the presence of the Ohmic losses γ in the materials introduces a threshold effect that depending upon Ωj < u or Ωj > u, the effective Rabi frequencies Ω˜s and Ω˜a can be either purely real or purely imaginary. In the other words, the Rabi frequencies may contribute to either damping of the amplitudes or sinusoidal oscillations. Secondly, in the strong coupling limit of Ωj > u, the losses present in the system are only those of the Ohmic losses γ. Thirdly, below threshold each of the amplitudes Cs (t) and Ca(t) is damped with two different rates, a reduced (subradiant) rate, 14γΩ˜j, and an enhanced (superradiant) rate, 14γ+Ω˜j. The involvement of two rather than a single damping rate seems to contradicts our expectation since, according to Dicke [69] (see also Refs. [70–72]), each of the collective amplitudes of a two atom system should decay with a single rate only. The amplitude Cs (t) should decay with the enhanced rate while Ca(t) should decay with the reduced rate.

 figure: Fig. 3

Fig. 3 (Color online) Two atoms located at a distance z0 from the interface between two materials and their images located at a distance z0 behind the interface. The atoms are not directly coupled to each other, but can be coupled by the radiation reflected from the interface. A photon emitted by atom 1 and reflected from the interface towards atom 2 can be viewed as being emitted from the image of the atom 1.

Download Full Size | PDF

A qualitative understanding of the involvement of two decay rates in the evolution of both Cs(t) and Ca(t) may be obtained by considering the interaction of the atoms with the plasmon field as the interaction with images of the atoms located at a distance z0 behind the interface. As illustrated in Fig. 3, the radiation field emitted by either atom 1 or atom 2 and reflected from the interface in the direction normal to the interface can be regarded as the radiation from an image located at a distance z0 behind the interface.

If we introduce the notation C˜j(t)=Cj(t)exp[(12γ+iδ)t] and C˜jI(t)=iΩ00tdtC˜j(t), where C˜jI(t) can be interpreted as the probability amplitude of the image j, then it is straightforward to write Eq. (12) in the form

C˜s(t)=(γ+2iδ)8Ω˜s[D˜a(t)cosϕs+iD˜s(t)sinϕs],C˜a(t)=(γ+2iδ)8Ω˜a[G˜a(t)cosϕa+iG˜s(t)sinϕa],
where
D˜s(t)=iC˜s(t)sinϕs+C˜sI(t)cosϕs=iCs(0)e[14(γ+2iδ)Ω˜s]tsinϕs,D˜a(t)=C˜s(t)cosϕsiC˜sI(t)sinϕs=Cs(0)e[14(γ+2iδ)+Ω˜s]tcosϕs,G˜s(t)=iCa(0)e[14(γ+2iδ)Ω˜a]tsinϕa,G˜a(t)=Ca(0)e[14(γ+2iδ)+Ω˜a]tcosϕa,
are symmetric and antisymmetric superpositions of the probability amplitudes of the atomic and image states, with
cos2ϕj=12+2Ω˜j(γ+2iδ),j=s,a.

The reason for the involvement of two decay rates in Eq. (12) is now clear: The term decaying with the enhanced rate is associated with the decay of the symmetric superpositions involving the atomic and image states, D˜s(t) and G˜s(t), whereas the term decaying with the reduced rate is associated with the decay of the antisymmetric superpositions D˜a(t) and G˜a(t). The slowest decay rate in the system, γa()=14γΩ˜a, is the decay rate of the antisymmetric superposition G˜a(t) whereas the fastest decay rate, γs(+)=14γ+Ω˜s, is the decay rate of the symmetric superposition D˜s(t).

We may conclude that the interaction of the atoms with the surface plasmon field can be viewed as the interaction between the atoms and their corresponding images. This also shows that the dynamics of a system composed of two atoms coupled to a plasmon field is analogous to the dynamics of a four-atom system in a rectangular configuration. Thus, one can obtain detailed and exact solutions for the dynamics of four interacting atoms in a rectangular configuration by considering the dynamics of the simpler system, two atoms coupled to a SP field.

3. Markovian and non-Markovian regimes of the evolutions

We have already seen that the Rabi frequencies of the collective states are altered by the interaction U(x21, z0), with Ωs enhanced and Ωa reduced by U(x21, z0). This results in two effective Rabi frequencies Ω˜s and Ω˜a which exhibit a threshold effect that depending upon Ωj < u, (b) Ωj > u, the effective Rabi frequencies can be purely real or imaginary. Since Ωs ≠ Ωa, the threshold effects occur at different u. Therefore, we can distinguish between three regions of Ωs and Ωa: (a) Ωs < u and Ωa < u, (b) Ωs > u and Ωa < u, (c) Ωs > u and Ωa > u. Physically, the threshold values of Ω˜s and Ω˜a separate what we can identify as the non-Markovian (memory preserved) regime from the Markovian (memoryless) regime of the evolution [73–76]. In the case (a), in which Ω˜s and Ω˜a are real, the Rabi frequencies contribute to the decay rates, which is a manifestation of a Markovian evolution. In the case (b), the dynamics of the system are partly Markovian and partly non-Markovian. The symmetric state undergoes a non-Markovian whereas the antisymmetric state undergoes a Markovian evolutions. In the case (c), the dynamics of the system are fully non-Markovian that the amplitudes of both symmetric and antisymmetric states undergo an oscillatory evolution, which is a manifestation of a non-Markovian evolution. A non-Markovian evolution is a reversible (coherent) process characterized by a continuous oscillation of an initial excitation between the atoms and the field. On the other hand, the Markovian evolution is a irreversible process of a flow of the excitation to the field resulting in the dissipation of an initial excitation. In experimental practice the non-Markovian evolution could be realized by locating the atoms at distances z0 from the interface such that the resulting magnitudes of Ωs and Ωa would be large enough to overcome the losses in the materials determined by the parameter u. The Markovian type evolution, on the other hand, could be achieved by locating the atoms at distances z0 such that the resulting magnitudes of Ωs and Ωa would be smaller than the losses in the materials.

With the experimentally realistic parameters, for example, of the impedance matched structure fabricated by the transmission line [61], where the surface plasmon resonant frequency ωs ~ 1.8 GHz and the Ohmic losses of the slabs are of a value γ ~ 0.3 MHz, and by taking a typical value for the atomic damping rate ΓA ~ 4 × 10−2 MHz, one can easily verify using Eq. (10) that the Rabi frequencies Ωs and Ωa can be larger than γ/4 at distances from the interface z0 < 0.45λs.

3.1. Both Ωs and Ωa below threshold

Let us specialize Eq. (12) to the case of exact resonance, δ = 0, and first examine the situation when the coupling of the atoms to the plasmon field is weak, Ω0γ. In this case, both Ω˜s and Ω˜a could be below threshold. If Ωj < u, then Ω˜s and Ω˜a are real and we see from Eq. (12) that the Rabi frequencies contribute to the damping rates of the probability amplitudes. As we have already explained, this is the kind of behavior corresponding to a Markovian evolution. Since Ωs ≠ Ωa, we see that the weak coupling of the atoms to the plasmon field may result in the decay of the probability amplitudes C1(t) and C2(t) with four rates, two enhanced and two reduced rates.

Figure 4 shows the time evolution of the populations P1(t) = |C1(t)|2 and P2(t) = |C2(t)|2 for both Ω˜s and Ω˜a below threshold. At early times, the population P1 (t) decreases whereas P2(t) increases until the populations become equal. At that time, the populations began to decay monotonically. The rate they decay is equal to γa(), the slowest decay rate of the antisymmetric state. Since the atoms are very strongly coupled to each other, U(x21, z0) = 0.95, the decay rate γa()0.002γ. Consequently, the effective decay time of the populations can be very long. The decay of the populations is irreversible so the evolution of the system is Markovian.

 figure: Fig. 4

Fig. 4 Time evolution of the populations P1(t) = |C1(t)|2 (black solid line) and P2(t) = |C2(t)|2 (red dashed line) for δ = 0, Ω0 = 0.15γ, and U(x21, z0) = 0.95, corresponding to both Ωs and Ωa below the threshold of 0.25γ, Ωs = 0.21γ and Ωa = 0.033γ. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

3.2. Ωs above threshold and Ωa below threshold

At large values of U(x21, z0) it may happen that Ωa < u and simultaneously Ωs > u even if the atoms are weakly coupled to the plasmon field, i.e. Ω0 < u. For example, when U(x21, z0) ≈ 1, we have Ωa ≈ 0 and Ωs ≈ 2Ω0. Hence, Ωs can be larger than u even if Ω0 < u and at the same time Ωa can be smaller than u. For Ωs > u the Rabi frequency Ω˜s is purely imaginary, and then the time evolution of the probability amplitude Cs(t) takes the form

Cs(t)=Cs(0)e14(γ+2iδ)t[cosΩ¯st+ueiθΩ¯ssinΩ¯st],
where Ω¯s=Ωs2(ueiθ)2. The temporal evolution of the Cs(t) is sinusoidal whereas the temporal evolution of the amplitude Ca(t), which is below threshold, is exponential and is given in Eq. (12). In this case, the symmetric mode evolves in the non-Markovian regime whereas the antisymmetric mode evolves in the Markovian regime. It follows that in this case each atom evolves under the simultaneous influence of Markovian and non-Markovian mechanisms.

Figure 5 shows the evolution of the populations for Ωs above and Ωa below the threshold of γ/4. At early times, the oscillations of the populations with the Rabi frequency of the symmetric mode are clearly visible. Meanwhile, the populations oscillate with the Rabi frequency of the symmetric mode. In other words, the evolution of the populations is reversible but the reversibility occurs in a restricted time range t<1/γa(). Beyond t~1/γa() the populations decay monotonically that the evolution is irreversible. Thus, we can clearly distinguish between the non-Markovian and Markovian regimes of the evolutions. We see that the upper limit on time of the reversible evolution results from the presence of the interaction between the atoms. Clearly, it is a collective effect. Physically, it is a consequence of the fact that a large part of the population is trapped in the asymmetric state determined by the amplitude G˜a(t) thereby lowering the strength of the coupling of the atoms to the plasmon field. It is easy to see, since for small distances between the atoms U(x21, z0) ≈ 1, we have Ωa ≈ 0, which means that the antisymmetric states decouple from the plasmon field. This example also shows that the atoms when behaving collectively can be weakly coupled to the field even if individually they are strongly coupled to the field.

 figure: Fig. 5

Fig. 5 Time evolution of the populations P1(t) = |C1(t)|2 (black solid line) and P2(t) = |C2(t)|2 (red dashed line) for δ = 0, Ω0 = 0.5γ, and U(x21, z0) = 0.99 corresponding to Ωs above and Ωa below the threshold of 0.25γ, Ωs = 0.705γ and Ωa = 0.05γ. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

3.3. Both Ωs and Ωa above threshold

Above the thresholds, Ω˜s and Ω˜a are purely imaginary. The time evolution of the probability amplitudes is then given by

Cj(t)=Cj(0)e14(γ+2iδ)t(cosΩ¯jt+ueiθΩ¯jsinΩ¯jt),
where Ω¯j=Ωj2(ueiθ)2. In this case, the time evolution of the probability amplitudes becomes sinusoidal. Such dynamics reflect the reversible property of the system that the evolution is non-Markovian.

Figure 6 shows the time evolution of the populations P1(t) = |C1(t)|2 and P2(t) = |C2(t)|2 when the atoms are strongly coupled to the plasmon field with z = 0.05λs, δ = 0, but are weakly coupled to each other, U(x21, z0) = 0.1. Note the presence of two characteristic time scales of the oscillations associated with the presence of two slightly different Rabi frequencies. At short times, t1/Ω¯a, the initially excited atom 1 periodically exchange the excitation with the plasmon field at the Rabi frequency Ω¯s. The population of the atom 2 builds up with the oscillation of frequency Ω¯s. The amplitudes of the populations are modulated with frequency Ω¯a causing collapses and revivals of the atomic populations.

 figure: Fig. 6

Fig. 6 Time evolution of the populations (a) P1(t) = |C1(t)|2 and (b) P2(t) = |C2(t)|2 for δ = 0, Ω0 = 25γ, and U(x21, z0) = 0.1. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

When the atoms are close to each other the collapses and revivals of the populations are absent. Instead, a periodic localization of the excitation is observed. This is illustrated in Fig. 7, which shows the evolution of the populations for a small distance between the atoms at which U(x21, z0) = 0.8. The manner the populations oscillate is different for P1(t) and P2(t). We see a periodic localization of the excitation that even at long times the memory effects are still evident. The explanation of this feature follows from the observation that at small distances between the atoms, the time scale of the oscillations of the antisymmetric state is very large, approaches infinity when U(x21, z0) → 1. Therefore, the system effectively evolves with a single time scale determined by the Rabi frequency of the symmetric state, t ~ 1/Ωs.

 figure: Fig. 7

Fig. 7 Time evolution of the populations (a) P1(t) = |C1(t)|2 and (b) P2(t) = |C2(t)|2 for δ = 0, Ω0 = 25γ, and U(x21, z0) = 0.8. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

4. Evolution of entanglement between the atoms

Given the time evolution of the probability amplitudes, we now proceed to evaluate the concurrence, a measure of entanglement between two qubits [71, 72]. Following the definition of the concurrence, we find that in terms of the amplitudes of the symmetric and antisymmetric combinations, the concurrence is given by the following expression

C(t)=|[Cs(t)Ca(t)][Cs*(t)+Ca*(t)]|.

A positive value of the concurrence, C(t) > 0, indicates entanglement between the atoms, and C(t) = 1 corresponds to maximally entangled atoms. It is clear from Eq. (18) that the atoms are entangled whenever Cs(t) ≠ Ca(t). Otherwise, the atoms are separable. Thus, to examine the occurrence of entanglement between the atoms we must look at differences in the evolution of the amplitudes Cs(t) and Ca(t). If initially, Cs(0) = Ca(0), then according to the solutions of Eq. (12), the amplitudes will evolve differently only if U(x21, z0) ≠ 0. It then follows that the coupling between the atoms through the plasmon field is necessary to create entanglement between the atoms from an initial separable state.

The features of the concurrence for the three regions of Ωs and Ωa are illustrated in Figs. 811. Figure 8 shows the evolution of the concurrence in the case of a strong coupling of the atoms to the plasmon field at which Ωs and Ωa are above their thresholds and two different values of the interaction strength between the atoms, U(x21, z0) = 0.1 and U(x21, z0) = 0.8. The interaction creates a small difference between the frequencies of the oscillation of the symmetric and antisymmetric modes that Ω¯sΩ¯a. The frequency difference induces beating oscillations of the populations of the atoms, as was seen in Fig. 6, and one can see from Fig. 8 that these beating oscillations are rendered visible as beats in the concurrence.

 figure: Fig. 8

Fig. 8 Concurrence versus time for the case of above threshold with Ω0 = 25γ, δ = 0, (a) U(x21, z0) = 0.1 and (b) U(x21, z0) = 0.8. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Concurrence versus time for Ω0 = γ, δ = 0 and U(x21, z0) ≈ 1 corresponding to the case of Ωs above threshold but Ωa below the threshold. Frame (a) shows the concurrence for U(x21, z0) = 0.95 (solid black line) and U(x21, z0) = 0.99 (dashed red line). The atoms were initially in a separable state |Ψ(0)〉 = |e1〉 |g2〉. Frame (b) shows the concurrence for U(x21, z0) = 0.95 and two different initial states, the maximally entangled symmetric state |Ψ(0)〉 = |s〉 (solid black line) and the maximally entangled antisymmetric state |Ψ(0)〉 = |a〉 (dashed red line).

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 The time evolution of the populations of the atoms for the situation presented in Fig. 9. Frame (a) shows the populations P1(t)(solid black line) and P2(t) (dashed red line) for Ω0 = γ, δ = 0 and U(x21, z0) = 0.95. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉. Frame (b) shows the time evolution of the population P1(t) for two different initial states, |Ψ(0)〉 = |s〉 (solid black line) and |Ψ(0)〉 = |a〉 (dashed green line). Not shown is P2(t) since in this case P2(t) = P1(t).

Download Full Size | PDF

 figure: Fig. 11

Fig. 11 Concurrence versus time for the case of below threshold and two different initial states (a) |Ψ(0)〉 = |e1〉 |g2〉 and (b) |Ψ(0)〉 = |s〉 with Ω0 = 0.15γ, δ = 0 and different U(x21, z0): U(x21, z0) = 0.99 (solid black line), U(x21, z0) = 0.5 (dashed red line), U(x21, z0) = 0.25 (dashed-dotted blue line).

Download Full Size | PDF

Interesting features of the entanglement also appear when the symmetric mode evolves at Rabi frequency which is above the threshold, Ωs > u, and simultaneously the antisymmetric mode evolves at Rabi frequency which is below the threshold, Ωa < u. Under this circumstance, the probability amplitude of the symmetric mode is determined by Eq. (17) whereas the amplitude of the antisymmetric mode is given by Eq. (12).

Figure 9 shows the evolution of the concurrence for this special case. We see that the concurrence is zero only at the initial time t = 0. As time progresses the concurrence develops to a nonzero value. The concurrence never becomes zero as time develops, and thus no periodic entanglement quenching occurs. This feature is associated with the fact that with the Rabi frequency Ωa < u, the population of the antisymmetric state does not evolve in time leading to a trapping of a part of the atomic populations in their energy states. This is illustrated in Fig. 10, which shows the time evolution of the population P1(t) and P2(t) for the same parameters as in Fig. 9. We see from Fig. 10(a) that the initial population is periodically transferred between the atoms. However, the transfer is not complete that the populations of the atoms never become zero during the evolution. A part of the population is trapped in the atoms and is not transferred between them. Figure 10(b) shows the evolution of the population P1(t) for two initial maximally entangled states, |s〉 and |a〉. Since in this case P2(t) = P1(t), we clearly see that the concurrence, if starts from maximally entangled state, it follows the evolution of the population of the atoms.

We now turn to the case of a weak coupling of the atoms to the plasmon field. In this case, the interaction of the atoms with the field contributes to the dissipation of the excitation. Figure 11 shows the effect of increasing interaction strength between the atoms on the concurrence for a weak coupling of the atoms to the plasmon field, both Ωs and Ωa below threshold, which corresponds to a Markovian evolution of the system. In Fig. 11(a) the system starts from the separable state |e1〉 |g2〉, whereas in Fig. 11(b) the initial state of the system is the maximally entangled state |s〉. We see that even in the weak coupling regime, a large and long living entanglement can be created between the atoms. The entanglement created from the initial separable state increases with an increasing U(x21, z0) and attains the maximal value of C = 0.5 for U(x21, z0) ≈ 1. The behavior of the concurrence is similar to that noted in the decay of two atoms into a common Markovian reservoir [72].

When the system starts from a maximally entangled state, either |s〉 or |a〉, the initial entanglement always decays to zero with no entanglement present at long times, as illustrated in Fig. 11(b). This is readily understood if it is recalled that the symmetric and antisymmetric states evolve independently in time. Thus, if the population of the antisymmetric state was initially zero it will remain zero for all times. In this case, the system of two atoms effectively behave as a single two-level system with the upper state |s〉 and the ground state |g〉. Then, the initial population of the state |s〉 decays exponentially to the ground state with the rate 2Ωs2/γ.

5. Off-resonant coupling

To this end we have discussed the collective effects induced by the resonant interaction of the atoms with the plasmon field. We have established the importance of the images of the atoms in the atomic dynamics. Moreover, we have demonstrated the equivalence of the system with that of four interacting atoms. We now turn to the off-resonant case of the atomic transition frequencies strongly detuned from the plasmon frequency, δγ, Ωs. Under such condition, the effective Rabi frequency Ω˜s can be approximated by

Ω˜s14γ(12Ωs2δ2)+12iδ(1+2Ωs2δ2).

Thus, if in Eq. (12) the effective Rabi frequency is replaced by (19), we get, up to terms of order Ωs2/δ2,

Cs(t)Cs(0){e(γ2iδ)Ωs22δ2t+Ωs24δ2e12(γ+2iδ)t}.

A similar expression with sa gives Ca(t). We see that Cs(t) is composed of fast and slow oscillating terms varying in time with frequencies δ and Ωs2/δ, respectively. Of the two terms it is the one of the small magnitude (Ωs2/4δ2) arising from the presence of the images. Thus, the evolution of Cs(t) is well determined without much contribution of the images. It is particularly well seen from Eq. (15) that in the limit of δ ≫ Ωs, cos ϕ ≈ 1 (sin ϕ ≈ 0) so that the superposition amplitude D˜s(t)=0 and D˜a(t) is reduced to the atomic amplitude C˜s(t). In other words, two atoms significantly detuned from the plasmon field are coupled each other by exchanging virtual photons through a short interaction time with the plasmon field.

The above considerations are illustrated in Fig. 12, which shows the time evolution of the atomic populations for δ = 50γ. We see that the atoms exchange the population with frequency 2Ω02/δ. The fast oscillations seen in the early time of the evolution occur at frequency δ and can be attributed to the involvement of the images in the dynamics of the system. The presence of the fast oscillations only at very early times of the evolution is also consistent with energy-time uncertainty arguments. It is easy to understand. At short times the uncertainty of the energy of the atoms and the plasmon field is very large, so one cannot distinguish between ωa and ωs. This results in the presence of the fast oscillation of frequency δ. As time progresses, the frequencies become more distinguishable resulting in the disappearance of the fast oscillations.

 figure: Fig. 12

Fig. 12 Time evolution of the populations (a) P1(t) = |C1(t)|2 and (b) P2(t) = |C2(t)|2 for δ = 50γ, Ω0 = 25γ, and U(x21, z0) = 0.1. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

It is interesting to contrast the entanglement created at δ ≠ 0 with that created in the resonant case of δ = 0. We have seen in Sec. 4 that in the resonant case the maximal entanglement which can be created between the atoms from an initial separable state cannot exceed C = 0.5. For off-resonant case (δ ≠ 0), however, the entanglement can be significantly enhanced. This is shown in Fig. 13, which illustrates the time evolution of the concurrence for a large detuning δ. Clearly, at early times of the evolutions the concurrence is larger than 1/2, increases with an increasing U(x21, z0) and becoming as large as C = 1. This behavior can be explained in terms of the energy-time uncertainty relation. At short times a large uncertainty in the energy results in a large uncertainty in the localization of the excitation. We see that the increased possibility to distinguish between the frequencies of the atoms and the plasmon field results in an enhanced entanglement between the atoms.

 figure: Fig. 13

Fig. 13 Concurrence versus time for the case of above threshold with Ω0 = 25γ, δ = 50γ and different U(x21, z0): (a) U(x21, z0) = 0.1 and (b) U(x21, z0) = 0.95. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Download Full Size | PDF

6. Conclusions

We have studied the dynamics of two two-level atoms located near to the interface of two meta-materials: one of negative permeability and the other of negative permittivity. We have derived analytical expressions for the probability amplitudes of the atomic states valid for an arbitrary initial state, arbitrary strengths of the coupling constants of the atoms to the plasmon field, and arbitrary distances between the atoms. We have shown that the effect of the surface plasmon field is to produce several interesting features, such as (1), a threshold behavior of the Rabi frequencies of the atoms to the plasmon field which distinguishes between the non-Markovian and Markovian regimes of the evolutions. It has been found that the threshold behavior may lead to three different regimes of the evolution; fully Markovian, simultaneous Markovian and non-Markovian, and fully non-Markovian evolutions. The three regimes determines three different time scales of the evolution of the memory effects and entanglement. (2) In the case of the resonant coupling of the surface plasmon field to the atoms, the field does not appear as a common reservoir to the atoms. We have adopted the image method and showed that in the resonant case the dynamics of the two atoms are analogous to those of the dynamics of a system of four atoms in a rectangular configuration. (3) In the limit of a strong detuning of the plasmon field frequency from the atomic transition frequencies the dynamics resemble those of two atoms coupled to a common reservoir.

Appendix

In this Appendix we give details of the derivation of the integro-differential Eqs. (25) and (26) for the probability amplitudes of the atomic states. Equations of motion for the probability amplitudes are obtained from the Schrödinger equation, which have the following forms

C˙1(t)=1πε00dωω2c2ei(ωωa)tdr{[ε(r,ω)]p1G(r1,r,ω)Ce(r,ω,t)+[κ(r,ω)]p1[G(r1,r,ω)×]Cm(r,ω,t)},
C˙2(t)=1πε00dωω2c2ei(ωωa)tdr{[ε(r,ω)]p2G(r2,r,ω)Ce(r,ω,t)+[κ(r,ω)]p2[G(r2,r,ω)×]Cm(r,ω,t)},
C˙e(r,ω,t)=ei(ωωa)tπε0ω2c2[ε(r,ω)][G*(r1,r,ω)p1*C1(t)+G*(r2,r,ω)p2*C2(t)],
C˙m(r,ω,t)=ei(ωωa)tπε0ωc[κ(r,ω)]×[G*(r1,r,ω)p1*C1(t)+G*(r2,r,ω)p2*C2(t)].

Integrating Eqs. (23) and (24), and substituting the solutions into Eqs. (21) and (22), we obtain

C˙1(t)=0tdtK11(t,t)C1(t)+0tdtK12(t,t)C2(t),
C˙2(t)=0tdtK22(t,t)C2(t)+0tdtK21(t,t)C1(t),
in which
Kij(t,t)=1πε0c20dωω2ei(ωωa)(tt)pi[G(ri,rj,ω)]pj*,i,j=1,2.

The parameter Kii(t, t′) is the integral kernel determined by the imaginary part of the one-point Green function, G(ri,ri,ω), whereas Kij(t, t′)(ij) is the integral kernel determined by the two-point Green function, G(ri,rj,ω).

The kernels are determined by expressions pi[G(ri,rj,ω)]pj*. In order to evaluate these expressions we have to specify the orientation of the atomic dipole moments. We assume that the atomic dipole moments are parallel to each other and have the same x, z components so that pi[G(ri,rj,ω)]pj* becomes

pi[G(ri,rj,ω)]pj*=|pi||pj|(x¯+z¯)[G(ri,rj,ω)](x¯+z¯)=|pi||pj|{[G(ri,rj,ω)]xx+[G(ri,rj,ω)]zz+[G(ri,rj,ω)]xz+[G(ri,rj,ω)]zx}.

Since Gxz (ri, rj, ω) = −Gzx (ri, rj, ω), the term pi[G(ri,rj,ω)]pj* therefore becomes independent of the off-diagonal elements. Thus, with the choice of the orientation of the atomic dipole moments given by Eq. (28), nonvanishing contributions to pi[G(ri,rj,ω)]pj* can come only from the diagonal x and z components of the Green function. Although the choice of the dipole polarization in the xz plane affects the contribution of the diagonal elements of the Green function, however, it has no effect on the contribution of the off diagonal elements since independent of the atomic polarization all off-diagonal elements involving the y component are zero.

We now proceed to evaluate the imaginary part of the Green function G(r,ri,ω). The Green function when evaluated at an arbitrary space point r, distance R = |rri| from the atom located at ri, can be written as [29]

G(r,ri,ω)=c2ω2ε1(+k12I)eik1RR,
where k1=ε1μ1ω/c represents the wave number, ε1 and μ1 are permittivity and permeability of the MN slab in which the atoms are located, and I=x¯x¯+y¯y¯+z¯z¯ is the unit dyadic. The presence of the boundaries results in the field inside the MN material consisting of waves propagating in both the +z and −z directions. Therefore, G(k,ω,z,z0) can be written in terms of functions Uq±(k,ω,z) defined by imposing boundary conditions in the z direction
G(k,ω,z,z0)=iμ12(2π)2d2kξqeiβ1d1β1Dq[Uq+(k,ω,z)Uq(k,ω,z0)Θ(zz0)+Uq(k,ω,z)Uq+(k,ω,z0)Θ(z0z)]eik(ρρ0),
that the functions Uq±(k,ω,z) describe the electric field in MN slab, with unit strength incident from its upper side (by taking symbol ’−’) or lower side (by taking symbol ’+’), that can be categorized into TM (q = p) and TE (q = s) types of even (ξp = 1) and odd (ξs = −1) symmetries in the ±z directions, and Θ(z) is the unit step function. The forms of the functions Uq±(k,ω,z) for the field inside the MN material can be represented in terms of a sum of incident and reflected waves as
Uq+(k,ω,z)=eq+(k)eiβ1(zd1)+r+qeq(k)eiβ1(zd1),Uq(k,ω,z)=eq(k)eiβ1z+rqeq+(k)eiβ1z,
where ep±(k)=(β1k¯+kz¯)/k1 and es±(k)=k¯×z¯ are orthonormal polarization vectors of the electric field of p and s polarized waves, respectively; k¯ is the unit vector in the direction of k(k=kk¯), x¯, y¯ and z¯ are unit vectors in the Cartesian coordinates, r±q are reflection coefficients of the waves propagating in the ±z directions, and Dq=1rqr+qe2iβ1d1 results from summing the geometrical series due to the multiple reflections from the boundaries between different materials.

We now proceed to evaluate the components of the Green function, which are defined by the following equation

Gnm(r,ri,ω)=n(G(r,ri,ω))m,n,m=x,y,z,n,m=x¯,y¯,z¯.

To evaluate the components of the Green function, we use the polar representation for k,k=k(cosϕx¯+sinϕy¯) and, for simplicity, we assume that rri has only x component so that we can write the dot product in the form form k · (rri) = αi cosϕ, where αi = k |rri|. When we apply the explicit forms of the polarization vectors, we arrive at the following expressions for the diagonal components

Gxx(r,ri,ω)=iμ14πdkk{β1k12[J1(αi)αiJ2(αi)]R(p)(z)+J1(αi)β1αiR+(s)(z)},Gyy(r,ri,ω)=iμ14πdkk{β1J1(αi)k12αiR(p)(z)+1β1[J1(αi)αiJ2(αi)]R+(s)(z)},Gzz(r,ri,ω)=iμ14πdkk3β1k12J0(αi)R+(p)(z),
and for the off-diagonal components
Gxy(r,ri,ω)=Gyx(r,ri,ω)=0,Gyz(r,ri,ω)=Gzy(r,ri,ω)=0,Gxz(r,ri,ω)=Gzx(r,ri,ω)=μ14πdkk2J1(αi)k12Dp[r+peiβ1(z+z02d1)rpeiβ1(z+z0)],
where
R±(q)(z)=1Dq[eiβ1(zz0)±rqeiβ1(z+z0)±r+qeiβ1(z+z02d1)+r+qrqeiβ1(zz02d1)].

The reflection coefficients appearing in Eqs. (33)(35) are determined from the boundary conditions, and have the form

rijp=(βiεjβjεi)/(βiεj+βjεi),rijs=(βiμjβjμi)/(βiμj+βjμi),ij=±,
where ij indicates the direction of the propagating wave, from the material i to j. For a multiple-reflection case, the reflection coefficient is given by
rijkq=(rijq+rjkqe2iβjdj)/(1rjiqrjkqe2iβjdj).

In the MN and EN slabs, the electromagnetic fields are evanescent and the perpendicular components of the wave vectors have a large imaginary part. The evanescent waves in the slabs split as the tunneling (k < ω/c) and plasmon (k > ω/c) modes. Both the electric fields of tunneling and plasmon modes are localized in the slabs and decay exponentially in the z direction, with the localization lengths l = 1/(2|βj|). For sufficiently thick slabs, the exponential functions are negligible with an increasing distance away from the interface. At a short distance from the interface, the terms rqeiβj(z+z0) are the dominant factors in the imaginary part of the Green function. For small Ohmic losses of the materials, the imaginary parts of βj are larger than |εjμjω/c|. Thus, the magnitudes of e2iβjdj are much smaller than one for the thickness of the slabs much larger than the localized length. Therefore, we can approximate rijkq, as given in Eq. (37), by rijq and when this result is inserted into Eq. (36), we obtain for the reflection coefficients

r+pr10p=(β1ε0β0ε1)/(β1ε0+β0ε1)(ε0ε1)/(ε0+ε1),
rp=(r12p+r20pe2iβ2d2)/(1r20p+r21pe2iβ2d2)(ε2ε1)/(ε1+ε2).

We would like to point out that the reflection coefficient at the interface between the EN and MN materials, Eq. (39), differs significantly from the reflection coefficient at the interface between an ordinary dielectric or a metal material [77,78]. According to Eq. (36), the dispersion relation for TM-polarized mode propagating along the interface can be approximated by ε1 β2 + ε2 β1 = 0. When ε1μ1 = ε2μ2 and ε1 = −Re[ε2], i.e. two slabs are perfectly paired, the dispersion relation reduces to ε1 + ε2 = 0. In this case, the propagation of the plasmon mode of frequency ωs=ωep/1+ε1 is independent of the parallel component of the wave vector. This property is different from that of the plasmon mode propagating at the interface of ordinary materials, where the resonant plasmon frequency depends on k [77].

Hence, adopting the results of Eq. (6), and assuming the frequency region of ωepωωeo, the reflection coefficients at the interface between the EN and MN slabs take the form

rp=1ε1ωs(Δω12iγ)(ε1+1)(Δω2+14γ2),
rs=1+μ1ωs(Δω12iγ)(μ1+1)(Δω2+14γ2),
where Δω = ωωs and, for simplicity, we have assumed that the dissipation parameters of the two slabs are equal, γe = γmγ.

If we now substitute Eqs. (40) and (41) into Eqs. (33), we can perform the integration and arrive at the analytical expressions for the imaginary parts of the components of the Green function. Following the result (28), we evaluate only the diagonal x and z components of the Green function. Thus, when setting the parameter values ε1 = μ2 = 2 and Re[μ1] = −μ2, Re[ε2] = −ε1, the explicit expressions for the imaginary parts of the diagonal z component of the one- and two-point Green functions are

[Gzz(r1,r1,ω)]=γωs12πk2(Δω2+14γ2)(2z0)3,
and
[Gzz(r2,r1,ω)]=γωs12πk2(Δω2+14γ2)(2z0)3F[32,2,1;x212(2z0)2].
where k = ω/c, x21 = x2x1 is the distance between the atoms, and F(a, b, c; x) is the hypergeometrical function. Similarly, for the diagonal x component of the one- and two-point Green functions, we find
[Gxx(r1,r1,ω)]=[Gxx(r2,r2,ω)]=γωs{1ks2[μ1(ωs)](2z0)2}24πk2(Δω2+14γ2)(2z0)3,
[Gxx(r2,r1,ω)]=[Gxx(r1,r2,ω)]=γωs24πk2(Δω2+14γ2)(2z0)3×{F[32,2,2;x212(2z0)2]3x212(2z0)2F[52,3,3;x212(2z0)2][μ1(ωs)](2z0ks)2F[12,1,2;x212(2z0)2]}.

These expressions show that the imaginary parts of the Green function, when evaluated as a function of ω, are of the form of a Lorentzian centered at the plasmon frequency ωs and possesses a bandwidth γ/2. Thus, for small γ we expect that the largest contributions to the field come from ωωs. Therefore, when substituting Eqs. (42)(45) into Eq. (27), we can replace ω2 by ωs2(ωa2) and extend the lower limit in the integration over ω to −∞. Then after a simple algebra we obtain the analytical expressions for the kernels given in Eq. (9).

Funding

National Natural Science Foundation of China (Grant No. 61275123 and No.11474119).

Acknowledgments

We acknowledge the financial support by National Natural Science Foundation of China (61275123 and 11474119).

References and links

1. S. Scheel, L. Knöll, and D. G. Welsch, “Spontaneous decay of an excited atom in an absorbing dielectric,” Phys. Rev. A 60, 4094 (1999). [CrossRef]  

2. L.-W. Li, M.-S. Leong, T.-S. Yeo, and P.-S. Kooi, “Electromagnetic dyadic Green’s functions in spectral domain for multilayered cylinders,” J. Electro. Waves Appl. 14, 961 (2000). [CrossRef]  

3. W. Lukosz, “Theory of optical-environment-dependent spontaneous-emission rates for emitters in thin layers,” Phys. Rev. B 22, 3030 (1980). [CrossRef]  

4. A. G. Tudela, P. A. Huidobro, L. M. Moreno, C. Tejedor, and F. J. García-Vidal, “Reversible dynamics of single quantum emitters near metal-dielectric interfaces,” Phys. Rev. B 89, 041402 (2014). [CrossRef]  

5. D. E. Chang, A. S. Sørensen, P. R. Hemmer, and M. D. Lukin, “Quantum optics with surface plasmons,” Phys. Rev. Lett. 97, 053002 (2006). [CrossRef]   [PubMed]  

6. V. Karanikolas, C. A. Marocico, and A. L. Bradley, “Spontaneous emission and energy transfer rates near a coated metallic cylinder,” Phys. Rev. A 89, 063817 (2014). [CrossRef]  

7. J. M. Wylie and J. E. Sipe, “Quantum electrodynamics near an interface,” Phys. Rev. A 30, 1185 (1984). [CrossRef]  

8. K. C. Liu and T. F. George, “Spontaneous emission by two atoms with different resonance frequencies near a metal surface,” Phys. Rev. B 32, 3622 (1985). [CrossRef]  

9. S. Scheel, S. Y. Buhmann, C. Clausen, and P. Schneeweiss, “Directional spontaneous emission and lateral Casimir-Polder force on an atom close to a nanofiber,” Phys. Rev. A 92, 043819 (2015). [CrossRef]  

10. Y. Luo, A. I. Fernandez-Dominguez, A. Wiener, S. A. Maier, and J. B. Pendry, “Surface plasmons and nonlocality: A simple model,” Phys. Rev. Lett. 111, 093901 (2013). [CrossRef]   [PubMed]  

11. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424, 824 (2003). [CrossRef]   [PubMed]  

12. R. Ruppin, “Surface polaritons of a left-handed medium,” Phys. Lett. A 277, 61–64 (2000). [CrossRef]  

13. E. R-Gómez, N. Raigoza, S. B. Cavalcanti, C. A. A. de Carvalho, and L. E. Oliveira, “Plasmon polaritons in photonic metamaterial Fibonacci superlattices,” Phys. Rev. B 81, 153101 (2010). [CrossRef]  

14. A. Delga, J. Feist, J. Bravo-Abad, and F. J. Garcia-Vidal, “Quantum emitters near a metal nanoparticle: strong coupling and quenching,” Phys. Rev. Lett. 112, 253601 (2014). [CrossRef]   [PubMed]  

15. J. X. Zhang, L. Zhang, and W. Xu, “Surface plasmon polaritons: physics and applications,” J. Phys. D: Appl. Phys. 45, 113001 (2012). [CrossRef]  

16. Z. H. Han and S. I. Bozhevolnyi, “Radiation guiding with surface plasmon polaritons,” Rep. Prog. Phys. 76, 016402 (2013). [CrossRef]  

17. M. S. Tame, K. R. McEnery, S. K. Özdemir, J. Lee, S. A. Maier, and M. S. Kim, “Quantum plasmonics,” Nat. Phys. 9, 329 (2013). [CrossRef]  

18. H. K. Yuan, U. K. Chettiar, W. S. Cai, A. V. Kildishev, A. Boltasseva, V. P. Drachev, and V. M. Shalaev, “A negative permeability material at red light,” Opt. Express 15, 1076 (2007). [CrossRef]   [PubMed]  

19. S. Linden, C. Enkrich, M. Wegener, J. F. Zhou, T. Koschny, and C. M. Soukoulis, “Magnetic response of metamaterials at 100 Terahertz,” Science 306, 1351 (2004). [CrossRef]   [PubMed]  

20. D. R. Smith, J. B. Pendry, and M. C. K. Wiltshire, “Metamaterials and negative refractive index,” Science 305, 788 (2004). [CrossRef]   [PubMed]  

21. L. W. Zhang, Y. W. Zhang, L. He, H. Q. Li, and H. Chen, “Experimental study of photonic crystals consisting of ε-negative and μ-negative materials,” Phys. Rev. E 74, 056615 (2006). [CrossRef]  

22. O. Hess, J. B. Pendry, S. A. Maier, R. F. Oulton, J. M. Hamm, and K. L. Tsakmakidis, “Active nanoplasmonic metamaterials,” Nat. Mater. 11, 573 (2012). [CrossRef]   [PubMed]  

23. M. I. Stockman, “Nanoplasmonics: past, present, and glimpse into future,” Opt. Express 19, 22029–22106 (2011). [CrossRef]   [PubMed]  

24. B. Stein, J. Y. Laluet, E. Devaux, C. Genet, and T. W. Ebbesen, “Surface plasmon mode steering and negative refraction,” Phys. Rev. Lett. 105, 266804 (2010). [CrossRef]  

25. Y. Li and N. Engheta, “Supercoupling of surface waves with ęÅ-near-zero metastructures,” Phys. Rev. B 90, 201107 (2014). [CrossRef]  

26. G. Dolling, C. Enkrich, M. Wegener, C. M. Soukoulis, and S. Linden, “Simultaneous negative phase and group velocity of light in a metamaterial,” Science 312, 892 (2006). [CrossRef]   [PubMed]  

27. M. Cuevas and R. A. Depine, “Radiation characteristics of electromagnetic eigenmodes at the corrugated interface of a left-handed material,” Phys. Rev. Lett. 103, 097401 (2009). [CrossRef]   [PubMed]  

28. G. X. Li, J. Evers, and C. H. Keitel, “Spontaneous emission interference in negative-refractive-index waveguides,” Phys. Rev. B 80, 045102 (2009). [CrossRef]  

29. H. T. Dung, S. Y. Buhmann, L. Knöll, D. G. Welsch, S. Scheel, and J. Kästel, “Electromagnetic-field quantization and spontaneous decay in left-handed media,” Phys. Rev. A 68, 043816 (2003). [CrossRef]  

30. X. D. Zeng, J. P. Xu, and Y. P. Yang, “Spontaneous emission interference enhancement with a mu-negative meta-material slab,” Phys. Rev. A 84, 033834 (2011). [CrossRef]  

31. J. Barthes, A. Bouhelier, A. Dereux, and G. C. des Francs, “Coupling of a dipolar emitter into one-dimensional surface plasmon,” Sci. Rep. 3, 2734 (2013). [CrossRef]   [PubMed]  

32. A. V. Akimov, A. Mukherjee, C. L. Yu, D. E. Chang, A. S. Zibrov, P. R. Hemmer, H. Park, and M. D. Lukin, “Generation of single optical plasmons in metallic nanowires coupled to quantum dots,” Nature 450, 402 (2007). [CrossRef]   [PubMed]  

33. K. J. Russell, T. L. Liu, S. Y. Cui, and E. L. Hu, “Large spontaneous emission enhancement in plasmonic nanocavities,” Nat. Photonics 6, 459 (2012). [CrossRef]  

34. Q. Cheng, W. X. Jiang, and T. J. Cui, “Spatial power combination for omnidirectional radiation via anisotropic metamaterials,” Phys. Rev. Lett. 108, 213903 (2012). [CrossRef]   [PubMed]  

35. A. G. Tudela, P. A. Huidobro, L. M. Moreno, C. Tejedor, and F. J. García-Vidal, “Theory of strong coupling between quantum emitters and propagating surface plasmons,” Phys. Rev. Lett. 110, 126801 (2013). [CrossRef]  

36. G. S. Agarwal, “Anisotropic vacuum-induced Interference in decay channels,” Phys. Rev. Lett. 84, 5500 (2000). [CrossRef]   [PubMed]  

37. G. Song, J. P. Xu, and Y. P. Yang, “Quantum interference between Zeeman levels near structures made of left-handed materials and matched zero-index metamaterials,” Phys. Rev. A 89, 053830 (2014). [CrossRef]  

38. C. Szymanowski, C. H. Keitel, B. J. Dalton, and P. L. Knight, “Switching between Rayleigh-like and Lorentzian lineshapes of the dispersion in driven two-level atoms,” J. Mod. Opt. 42, 985 (1995). [CrossRef]  

39. R. C. Ge, C. V. Vlack, P. Yao, J. F. Young, and S. Hughes, “Accessing quantum nanoplasmonics in a hybrid quantum dot–metal nanosystem: Mollow triplet of a quantum dot near a metal nanoparticle,” Phys. Rev. B 87, 205425 (2013). [CrossRef]  

40. M. S. Tomaš, “Green function for multilayers: Light scattering in planar cavities,” Phys. Rev. A 51, 2545 (1995). [CrossRef]  

41. H. P. Breuer, D. Faller, B. Kappler, and F. Petruccione, “Non-Markovian dynamics in pulsed- and continuous-wave atom lasers,” Phys. Rev. A 60, 3188 (1999). [CrossRef]  

42. A. Gonzalez-Tudela, D. Martin-Cano, E. Moreno, L. Martin-Moreno, C. Tejedor, and F. J. Garcia-Vidal, “Entanglement of two qubits mediated by one-dimensional plasmonic waveguides,” Phys. Rev. Lett. 106, 020501 (2011). [CrossRef]   [PubMed]  

43. C. Gonzalez-Ballestero, E. Moren, and F. J. Garcia-Vidal, “Generation, manipulation, and detection of two-qubit entanglement in waveguide QED,” Phys. Rev. A 89, 042328 (2014). [CrossRef]  

44. D. Martõn-Cano, A. Gonzalez-Tudela, L. Martõn-Moreno, F. J. Garcia-Vidal, C. Tejedor, and E. Moreno, “Dissipation-driven generation of two-qubit entanglement mediated by plasmonic waveguides,” Phys. Rev. B 84, 235306 (2011). [CrossRef]  

45. J. D. Hood, A. Goban, A. Asenjo-Garcia, M. Lu, S.-P. Yu, D. E. Chang, and H. J. Kimble, “Atom-atom interaction around the band edge of a photonic crystal waveguide,” arXiv: 1603.02771 (2016).

46. R. C. Ge and S. Hughes, “Quantum dynamics of two quantum dots coupled through localized plasmons: An intuitive and accurate quantum optics approach using quasinormal modes”, Phys. Rev. B 92, 205420 (2015). [CrossRef]  

47. 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]  

48. M. Otten, R. A. Shah, N. F. Scherer, M. Min, M. Pelton, and S. K. Gray, “Entanglement of two, three, or four plasmonically coupled quantum dots,” Phys. Rev. B 92, 125432 (2015). [CrossRef]  

49. J. P. Xu, M. Al-Amri, Y. P. Yang, S. Y. Zhu, and M. S. Zubairy, “Entanglement generation between two atoms via surface modes,” Phys. Rev. A 84, 032334 (2011). [CrossRef]  

50. W. Fang, Q. L. Wu, S. P. Wu, and G.-x. Li, “Enhancement of squeezing in resonance fluorescence of a driven quantum dot close to a graphene sheet,” Phys. Rev. A 93, 053831 (2016). [CrossRef]  

51. H. X. Zheng and H. U. Baranger, “Persistent quantum beats and long-distance entanglement from waveguide-mediated interactions,” Phys. Rev. Lett. 110, 113601 (2013). [CrossRef]   [PubMed]  

52. D. Dzsotjan, A. S. Sørensen, and M. Fleischhauer, “Quantum emitters coupled to surface plasmons of a nanowire: A Green’s function approach,” Phys. Rev. B 82, 075427 (2010). [CrossRef]  

53. C. H. Bennett, “Quantum information and computation,” Phys. Today 48(10), 24 (1995). [CrossRef]  

54. S. Haroche and J. M. Raimond, Exploring the quantum (Oxford University Press, 2006). [CrossRef]  

55. J. B. Pendry, “Extremely low frequency plasmons in metallic mesostructures,” Phys. Rev. Lett. 76, 4773 (1996). [CrossRef]   [PubMed]  

56. A. Huck, S. Smolka, P. Loudahl, A. S. Sørenson, A. Boltasseva, J. Janousek, and U. L. Andersen, “Demonstration of quadrature-squeezed surface plasmons in a gold waveguide,” Phys. Rev. Lett. 102, 246802 (2009). [CrossRef]   [PubMed]  

57. R. Ruppin, “Surface polaritons of a left-handed material slab,” J. Phys. Condens. Matter 13, 1811 (2001). [CrossRef]  

58. W. Tan, Y. Sun, H. Chen, and S. Q. Shen, “Photonic simulation of topological excitations in metamaterials,” Scientific Reports 4, 3842 (2014). [CrossRef]   [PubMed]  

59. H. T. Jiang, H. Chen, and S. Y. Zhu, “Rabi splitting with excitons in effective (near) zero-index media,” Opt. Lett. 32, 1980 (2007). [CrossRef]   [PubMed]  

60. X. Huang, Y. Lai, Z. H. Hang, H. Zheng, and C. T. Chan, “Dirac cones induced by accidental degeneracy in photonic crystals and zero-refractive-index materials,” Nat. Mater. 10, 582 (2011). [CrossRef]   [PubMed]  

61. L. W. Zhang, Y. W. Zhang, L. He, H. Q. Li, and H. Chen, “Experimental study of photonic crystals consisting of ε-negative and μ-negative materials,” Phys. Rev. E 74, 056615 (2006). [CrossRef]  

62. P. Moitra, Y. Yang, Z. Anderson, I. I. Kravchenko, D. P. Briggs, and J. Valentine, “Realization of an all-dielectric zero-index optical metamaterial,” Nat. Photonics 7, 791–795 (2013). [CrossRef]  

63. G. Dolling, C. Enkrich, M. Wegener, J. F. Zhou, C. M. Soukoulis, and S. Linden, “Cut-wire pairs and plate pairs as magnetic atoms for optical metamaterials,” Opt. Lett. 30, 3198 (2005). [CrossRef]   [PubMed]  

64. S. Y. Buhmann and D.-G. Welsch, “Dispersion forces in macroscopic quantum electrodynamics,” Prog. Quant. Electr. 31, 51 (2007). [CrossRef]  

65. R. A. Shelby, D. R. Smith, S. C. Nemat-Nasser, and S. Schultz, “Microwave transmission through a two-dimensional, isotropic, left-handed metamaterial,” Appl. Phys. Lett. 78, 489 (2001). [CrossRef]  

66. I. Bergmair, B. Dastmalchi, M. Bergmair, A. Saeed, W. Hilber, G. Hesser, C. Helgert, E. Pshenay-Severin, T. Pertsch, E. B. Kley, U. Hübner, N. H. Shen, R. Penciu, M. Kafesaki, C. M. Soukoulis, K. Hingerl, M. Muehlberger, and R. Schoeftner, “Single and multilayer metamaterials fabricated by nanoimprint lithography,” Nano Technol. 22, 325301 (2011).

67. N. T. Tung, V. D. Lam, J. W. Park, M. H. Cho, J. Y. Rhee, W. H. Jang, and Y. P. Lee, “Single- and double-negative refractive indices of combined metamaterial structure,” J. App. Phys. 106, 053109 (2009). [CrossRef]  

68. G. Dolling, C. Enkrich, M. Wegener, C. M. Soukoulis, and S. Linden, “Low-loss negative-index metamaterial at telecommunication wavelengths,” Opt. Lett. 31, 1800 (2006). [CrossRef]   [PubMed]  

69. R. H. Dicke, “Coherence in spontaneous radiation processes,” Phys. Rev. 93, 99 (1954). [CrossRef]  

70. R. H. Lehmberg, “Radiation from an N-atom system. II. Spontaneous emission from a pair of atoms,” Phys. Rev. A 2, 889 (1970). [CrossRef]  

71. L. H. Sun, G. X. Li, and Z. Ficek, “Continuous variables approach to entanglement creation and processing,” Appl. Math. Inf. Sci. 4, 315 (2010).

72. Z. Ficek, “Quantum entanglement processing with atoms,” Appl. Math. Inf. Sci. 3, 375 (2009).

73. H. P. Breuer and F. Petruccione, Theory of Open Quantum Systems (Oxford University Press, 2002).

74. T. J. G. Apollaro, S. Lorenzo, C. Di Franco, F. Plastina, and M. Paternostro, “Competition between memory-keeping and memory-erasing decoherence channels,” Phys. Rev. A 90, 012310 (2014). [CrossRef]  

75. Z. X. Man, Y. J. Xia, and R. Lo Franco, “Harnessing non-Markovian quantum memory by environmental coupling,” Phys. Rev. A 92, 012315 (2015). [CrossRef]  

76. Z. X. Man, N. B. An, and Y. J. Xia, “Non-Markovian dynamics of a two-level system in the presence of hierarchical environments,” Opt. Express 23, 5763 (2015). [CrossRef]   [PubMed]  

77. I. Avrutsky, I. Salakhutdinov, J. Elser, and V. Podolskiy, “Highly confined optical modes in nanoscale metal-dielectric multilayers,” Phys. Rev. B 75, 241402 (2007). [CrossRef]  

78. A. Archambault, T. V. Teperik, F. Marquier, and J. J. Greffet, “Surface plasmon Fourier optics,” Phys. Rev. B 79, 195414 (2009). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (13)

Fig. 1
Fig. 1 (Color online) Schematic illustration of the system showing the electric field (yellow arrows) of an electromagnetic wave and dislocated surface charges at the interface between MN and EN slabs. The z axis is taken normal to the interface with its origin at the interface. The slabs have thickness d1 and d2, respectively, and are assumed to have infinite extents in the xy plane. Two atoms are embedded in the MN slab at fixed positions (x1, 0, z0) and (x2, 0, z0), where z0 is the distance of the atoms from the interface between the materials. The atomic transition dipole moments p1 and p2 are parallel to each other and oriented in the xz plane.
Fig. 2
Fig. 2 Dependence of U(x21, z0) on separation between the atoms x21s for several different distances of the atoms from the interface: z0 = 0.05λs (solid black line), z0 = 0.1λs (dashed red line), z0 = 0.25λs (dashed-dotted blue line), and z0 = 0.5λs (solid green line).
Fig. 3
Fig. 3 (Color online) Two atoms located at a distance z0 from the interface between two materials and their images located at a distance z0 behind the interface. The atoms are not directly coupled to each other, but can be coupled by the radiation reflected from the interface. A photon emitted by atom 1 and reflected from the interface towards atom 2 can be viewed as being emitted from the image of the atom 1.
Fig. 4
Fig. 4 Time evolution of the populations P1(t) = |C1(t)|2 (black solid line) and P2(t) = |C2(t)|2 (red dashed line) for δ = 0, Ω0 = 0.15γ, and U(x21, z0) = 0.95, corresponding to both Ωs and Ωa below the threshold of 0.25γ, Ωs = 0.21γ and Ωa = 0.033γ. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.
Fig. 5
Fig. 5 Time evolution of the populations P1(t) = |C1(t)|2 (black solid line) and P2(t) = |C2(t)|2 (red dashed line) for δ = 0, Ω0 = 0.5γ, and U(x21, z0) = 0.99 corresponding to Ωs above and Ωa below the threshold of 0.25γ, Ωs = 0.705γ and Ωa = 0.05γ. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.
Fig. 6
Fig. 6 Time evolution of the populations (a) P1(t) = |C1(t)|2 and (b) P2(t) = |C2(t)|2 for δ = 0, Ω0 = 25γ, and U(x21, z0) = 0.1. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.
Fig. 7
Fig. 7 Time evolution of the populations (a) P1(t) = |C1(t)|2 and (b) P2(t) = |C2(t)|2 for δ = 0, Ω0 = 25γ, and U(x21, z0) = 0.8. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.
Fig. 8
Fig. 8 Concurrence versus time for the case of above threshold with Ω0 = 25γ, δ = 0, (a) U(x21, z0) = 0.1 and (b) U(x21, z0) = 0.8. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.
Fig. 9
Fig. 9 Concurrence versus time for Ω0 = γ, δ = 0 and U(x21, z0) ≈ 1 corresponding to the case of Ωs above threshold but Ωa below the threshold. Frame (a) shows the concurrence for U(x21, z0) = 0.95 (solid black line) and U(x21, z0) = 0.99 (dashed red line). The atoms were initially in a separable state |Ψ(0)〉 = |e1〉 |g2〉. Frame (b) shows the concurrence for U(x21, z0) = 0.95 and two different initial states, the maximally entangled symmetric state |Ψ(0)〉 = |s〉 (solid black line) and the maximally entangled antisymmetric state |Ψ(0)〉 = |a〉 (dashed red line).
Fig. 10
Fig. 10 The time evolution of the populations of the atoms for the situation presented in Fig. 9. Frame (a) shows the populations P1(t)(solid black line) and P2(t) (dashed red line) for Ω0 = γ, δ = 0 and U(x21, z0) = 0.95. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉. Frame (b) shows the time evolution of the population P1(t) for two different initial states, |Ψ(0)〉 = |s〉 (solid black line) and |Ψ(0)〉 = |a〉 (dashed green line). Not shown is P2(t) since in this case P2(t) = P1(t).
Fig. 11
Fig. 11 Concurrence versus time for the case of below threshold and two different initial states (a) |Ψ(0)〉 = |e1〉 |g2〉 and (b) |Ψ(0)〉 = |s〉 with Ω0 = 0.15γ, δ = 0 and different U(x21, z0): U(x21, z0) = 0.99 (solid black line), U(x21, z0) = 0.5 (dashed red line), U(x21, z0) = 0.25 (dashed-dotted blue line).
Fig. 12
Fig. 12 Time evolution of the populations (a) P1(t) = |C1(t)|2 and (b) P2(t) = |C2(t)|2 for δ = 50γ, Ω0 = 25γ, and U(x21, z0) = 0.1. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.
Fig. 13
Fig. 13 Concurrence versus time for the case of above threshold with Ω0 = 25γ, δ = 50γ and different U(x21, z0): (a) U(x21, z0) = 0.1 and (b) U(x21, z0) = 0.95. The atoms were initially in the state |Ψ(0)〉 = |e1〉 |g2〉.

Equations (45)

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

ε r = ( d 1 ε 1 + d 2 ε 2 ) / ( d 1 + d 2 ) , μ r = ( d 1 μ 1 + d 2 μ 2 ) / ( d 1 + d 2 )
H ^ = H ^ 0 + H ^ I ,
H ^ 0 = 1 2 ω a ( σ ^ z 1 + σ ^ z 2 ) + λ = e , m d r 0 d ω ω f ^ λ ( r , ω ) f ^ λ ( r , ω )
H ^ I = l = 1 , 2 [ p l 0 d ω E ^ ( + ) ( r l , ω ) σ ^ l + H . c . ]
E ^ ( + ) ( r l , ω ) = i π ε 0 ω c d r { ω c [ ε ( r , ω ) ] G ( r l , r , ω ) f ^ e ( r , ω ) + [ κ ( r , ω ) ] × G ( r l , r , ω ) f ^ m ( r , ω ) } ,
ε 2 ( ω ) ε 0 = 1 + ω e p 2 ω e o   2 ω 2 i ω γ e , μ 1 ( ω ) μ 0 = 1 + ω m p 2 ω m o   2 ω 2 i ω γ m ,
| Ψ ( t ) = C 1 ( t ) e i ω a t | { 0 } | e 1 , g 2 + C 2 ( t ) e i ω a t | { 0 } | g 1 , e 2 + λ = e , m 0 d ω e i ω t d r C λ ( r , ω , t ) | 1 λ ( r , ω ) | g 1 , g 2 ,
C ˙ s ( t ) = 0 t d t K s ( t , t ) C s ( t ) , C ˙ a ( t ) = 0 t d t K a ( t , t ) C a ( t ) ,
K j ( t , t ) = Ω j 2 e ( 1 2 γ + i δ ) ( t t ) , j = s , a .
Ω 0 = { [ 3 + 4 π 2 | [ μ 1 ( ω s ) ] | ( 2 z 0 / λ s ) 2 ] ω s Γ A 64 π 3 ( 2 z 0 / λ s ) 3 } 1 / 2
U ( x 21 , z 0 ) = F [ 1 2 , 1 , 2 ; x 21 2 ( 2 z 0 ) 2 ] + 1 3 + 4 π 2 | [ μ 1 ( ω s ) ] | ( 2 z 0 / λ s ) 2 { F [ 3 2 , 2 , 2 ; x 21 2 ( 2 z 0 ) 2 ] + 2 F [ 3 2 , 2 , 1 ; x 21 2 ( 2 z 0 ) 2 ] 3 F [ 1 2 , 1 , 2 ; x 21 2 ( 2 z 0 ) 2 ] 3 x 21 2 ( 2 z 0 ) 2 F [ 5 2 , 3 , 3 ; x 21 2 ( 2 z 0 ) 2 ] }
C j ( t ) = 1 2 C j ( 0 ) e 1 2 i δ t [ ( 1 + u e i θ Ω ˜ j ) e ( 1 4 γ Ω ˜ j ) t + ( 1 u e i θ Ω ˜ j ) e ( 1 4 γ + Ω ˜ j ) t ] , j = s , a ,
C ˜ s ( t ) = ( γ + 2 i δ ) 8 Ω ˜ s [ D ˜ a ( t ) cos ϕ s + i D ˜ s ( t ) sin ϕ s ] , C ˜ a ( t ) = ( γ + 2 i δ ) 8 Ω ˜ a [ G ˜ a ( t ) cos ϕ a + i G ˜ s ( t ) sin ϕ a ] ,
D ˜ s ( t ) = i C ˜ s ( t ) sin ϕ s + C ˜ s I ( t ) cos ϕ s = i C s ( 0 ) e [ 1 4 ( γ + 2 i δ ) Ω ˜ s ] t sin ϕ s , D ˜ a ( t ) = C ˜ s ( t ) cos ϕ s i C ˜ s I ( t ) sin ϕ s = C s ( 0 ) e [ 1 4 ( γ + 2 i δ ) + Ω ˜ s ] t cos ϕ s , G ˜ s ( t ) = i C a ( 0 ) e [ 1 4 ( γ + 2 i δ ) Ω ˜ a ] t sin ϕ a , G ˜ a ( t ) = C a ( 0 ) e [ 1 4 ( γ + 2 i δ ) + Ω ˜ a ] t cos ϕ a ,
cos 2 ϕ j = 1 2 + 2 Ω ˜ j ( γ + 2 i δ ) , j = s , a .
C s ( t ) = C s ( 0 ) e 1 4 ( γ + 2 i δ ) t [ cos Ω ¯ s t + u e i θ Ω ¯ s sin Ω ¯ s t ] ,
C j ( t ) = C j ( 0 ) e 1 4 ( γ + 2 i δ ) t ( cos Ω ¯ j t + u e i θ Ω ¯ j sin Ω ¯ j t ) ,
C ( t ) = | [ C s ( t ) C a ( t ) ] [ C s * ( t ) + C a * ( t ) ] | .
Ω ˜ s 1 4 γ ( 1 2 Ω s 2 δ 2 ) + 1 2 i δ ( 1 + 2 Ω s 2 δ 2 ) .
C s ( t ) C s ( 0 ) { e ( γ 2 i δ ) Ω s 2 2 δ 2 t + Ω s 2 4 δ 2 e 1 2 ( γ + 2 i δ ) t } .
C ˙ 1 ( t ) = 1 π ε 0 0 d ω ω 2 c 2 e i ( ω ω a ) t d r { [ ε ( r , ω ) ] p 1 G ( r 1 , r , ω ) C e ( r , ω , t ) + [ κ ( r , ω ) ] p 1 [ G ( r 1 , r , ω ) × ] C m ( r , ω , t ) } ,
C ˙ 2 ( t ) = 1 π ε 0 0 d ω ω 2 c 2 e i ( ω ω a ) t d r { [ ε ( r , ω ) ] p 2 G ( r 2 , r , ω ) C e ( r , ω , t ) + [ κ ( r , ω ) ] p 2 [ G ( r 2 , r , ω ) × ] C m ( r , ω , t ) } ,
C ˙ e ( r , ω , t ) = e i ( ω ω a ) t π ε 0 ω 2 c 2 [ ε ( r , ω ) ] [ G * ( r 1 , r , ω ) p 1 * C 1 ( t ) + G * ( r 2 , r , ω ) p 2 * C 2 ( t ) ] ,
C ˙ m ( r , ω , t ) = e i ( ω ω a ) t π ε 0 ω c [ κ ( r , ω ) ] × [ G * ( r 1 , r , ω ) p 1 * C 1 ( t ) + G * ( r 2 , r , ω ) p 2 * C 2 ( t ) ] .
C ˙ 1 ( t ) = 0 t d t K 11 ( t , t ) C 1 ( t ) + 0 t d t K 12 ( t , t ) C 2 ( t ) ,
C ˙ 2 ( t ) = 0 t d t K 22 ( t , t ) C 2 ( t ) + 0 t d t K 21 ( t , t ) C 1 ( t ) ,
K i j ( t , t ) = 1 π ε 0 c 2 0 d ω ω 2 e i ( ω ω a ) ( t t ) p i [ G ( r i , r j , ω ) ] p j * , i , j = 1 , 2 .
p i [ G ( r i , r j , ω ) ] p j * = | p i | | p j | ( x ¯ + z ¯ ) [ G ( r i , r j , ω ) ] ( x ¯ + z ¯ ) = | p i | | p j | { [ G ( r i , r j , ω ) ] x x + [ G ( r i , r j , ω ) ] z z + [ G ( r i , r j , ω ) ] x z + [ G ( r i , r j , ω ) ] z x } .
G ( r , r i , ω ) = c 2 ω 2 ε 1 ( + k 1 2 I ) e i k 1 R R ,
G ( k , ω , z , z 0 ) = i μ 1 2 ( 2 π ) 2 d 2 k ξ q e i β 1 d 1 β 1 D q [ U q + ( k , ω , z ) U q ( k , ω , z 0 ) Θ ( z z 0 ) + U q ( k , ω , z ) U q + ( k , ω , z 0 ) Θ ( z 0 z ) ] e i k ( ρ ρ 0 ) ,
U q + ( k , ω , z ) = e q + ( k ) e i β 1 ( z d 1 ) + r + q e q ( k ) e i β 1 ( z d 1 ) , U q ( k , ω , z ) = e q ( k ) e i β 1 z + r q e q + ( k ) e i β 1 z ,
G n m ( r , r i , ω ) = n ( G ( r , r i , ω ) ) m , n , m = x , y , z , n , m = x ¯ , y ¯ , z ¯ .
G x x ( r , r i , ω ) = i μ 1 4 π d k k { β 1 k 1 2 [ J 1 ( α i ) α i J 2 ( α i ) ] R ( p ) ( z ) + J 1 ( α i ) β 1 α i R + ( s ) ( z ) } , G y y ( r , r i , ω ) = i μ 1 4 π d k k { β 1 J 1 ( α i ) k 1 2 α i R ( p ) ( z ) + 1 β 1 [ J 1 ( α i ) α i J 2 ( α i ) ] R + ( s ) ( z ) } , G z z ( r , r i , ω ) = i μ 1 4 π d k k 3 β 1 k 1 2 J 0 ( α i ) R + ( p ) ( z ) ,
G x y ( r , r i , ω ) = G y x ( r , r i , ω ) = 0 , G y z ( r , r i , ω ) = G z y ( r , r i , ω ) = 0 , G x z ( r , r i , ω ) = G z x ( r , r i , ω ) = μ 1 4 π d k k 2 J 1 ( α i ) k 1 2 D p [ r + p e i β 1 ( z + z 0 2 d 1 ) r p e i β 1 ( z + z 0 ) ] ,
R ± ( q ) ( z ) = 1 D q [ e i β 1 ( z z 0 ) ± r q e i β 1 ( z + z 0 ) ± r + q e i β 1 ( z + z 0 2 d 1 ) + r + q r q e i β 1 ( z z 0 2 d 1 ) ] .
r i j p = ( β i ε j β j ε i ) / ( β i ε j + β j ε i ) , r i j s = ( β i μ j β j μ i ) / ( β i μ j + β j μ i ) , i j = ± ,
r i j k q = ( r i j q + r j k q e 2 i β j d j ) / ( 1 r j i q r j k q e 2 i β j d j ) .
r + p r 1 0 p = ( β 1 ε 0 β 0 ε 1 ) / ( β 1 ε 0 + β 0 ε 1 ) ( ε 0 ε 1 ) / ( ε 0 + ε 1 ) ,
r p = ( r 1 2 p + r 2 0 p e 2 i β 2 d 2 ) / ( 1 r 2 0 p + r 2 1 p e 2 i β 2 d 2 ) ( ε 2 ε 1 ) / ( ε 1 + ε 2 ) .
r p = 1 ε 1 ω s ( Δ ω 1 2 i γ ) ( ε 1 + 1 ) ( Δ ω 2 + 1 4 γ 2 ) ,
r s = 1 + μ 1 ω s ( Δ ω 1 2 i γ ) ( μ 1 + 1 ) ( Δ ω 2 + 1 4 γ 2 ) ,
[ G z z ( r 1 , r 1 , ω ) ] = γ ω s 12 π k 2 ( Δ ω 2 + 1 4 γ 2 ) ( 2 z 0 ) 3 ,
[ G z z ( r 2 , r 1 , ω ) ] = γ ω s 12 π k 2 ( Δ ω 2 + 1 4 γ 2 ) ( 2 z 0 ) 3 F [ 3 2 , 2 , 1 ; x 21 2 ( 2 z 0 ) 2 ] .
[ G x x ( r 1 , r 1 , ω ) ] = [ G x x ( r 2 , r 2 , ω ) ] = γ ω s { 1 k s 2 [ μ 1 ( ω s ) ] ( 2 z 0 ) 2 } 24 π k 2 ( Δ ω 2 + 1 4 γ 2 ) ( 2 z 0 ) 3 ,
[ G x x ( r 2 , r 1 , ω ) ] = [ G x x ( r 1 , r 2 , ω ) ] = γ ω s 24 π k 2 ( Δ ω 2 + 1 4 γ 2 ) ( 2 z 0 ) 3 × { F [ 3 2 , 2 , 2 ; x 21 2 ( 2 z 0 ) 2 ] 3 x 21 2 ( 2 z 0 ) 2 F [ 5 2 , 3 , 3 ; x 21 2 ( 2 z 0 ) 2 ] [ μ 1 ( ω s ) ] ( 2 z 0 k s ) 2 F [ 1 2 , 1 , 2 ; x 21 2 ( 2 z 0 ) 2 ] } .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.