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

Optical 𝒫𝒯-symmetry and 𝒫𝒯-antisymmetry in coherently driven atomic lattices

Open Access Open Access

Abstract

We study an ensemble of ultracold atoms trapped in the one-dimensional optical lattices under the N level configuration to examine the absorption and dispersion properties modulated into different ways along the lattice direction z. We find these trapped atoms with a Gaussian density distribution in each period may exhibit either symmetry χp(z)=χp*(z) or antisymmetry χp(z)=χp*(z) of parity-time (𝒫𝒯) in terms of the probe susceptibility. Such intriguing twofold modulations of real (χ′ p) and imaginary (χ″ p) susceptibilities with a π/2 phase shift are attained by spatially modulating intensities or frequencies of one driving field in a suitable way. The 𝒫𝒯-symmetric or 𝒫𝒯-antisymmetric atomic lattices correspond in fact to complex photonic crystals and may be extended to develop functional devices like photonic diodes and transistors, a task impossible for real photonic crystals.

© 2016 Optical Society of America

1. Introduction

The development of metamaterial has drawn considerable attention and research effort in recent years, for the capability of bringing unusual and remarkable electromagnetic properties unseen in the natural world. Among the works, the transport properties of Hermitian lattice systems have always been a subject of intense investigation, while much less attention has been paid to non-Hermitian systems, until 1998 when Bender and Boettcher proposed that a non-Hermitian Hamiltonian exhibiting parity-time (𝒫𝒯) symmetry can also have entirely real energy eigen-spectrum in certain regions of parameter space [1,2]. Since then the concept of 𝒫𝒯-symmetry has been widely considered in fields like quantum field theories, non-Hermitian Anderson models, complex Lie algebras, lattice QCD theories, etc., within the theoretical study domain. On the other hand, due to the isomorphism between time-dependent Schrödinger equations and optical paraxial wave equations, optical systems provide the ideal test bed where the features of 𝒫𝒯-symmetry can be experimentally explored [3], and many intriguing properties, like double refraction and band merging [4, 5], power oscillations [6–8 ], nonreciprocal light propagation [9], coherent perfect absorbers [10–13 ], unidirectional invisibility [5,14–17 ], and so on, have been studied since then.

The study of 𝒫𝒯-symmetry on optical systems, either theoretically or experimentally, till this time has been carried out on coupled waveguides [6,18,19], photonic structures [5,20,21], transmission lines [13], whispering-gallery microcavities [22, 23], and optomechanical systems [24, 25]. However, there hasn’t been much attempt to utilize atomic clouds or vapors. Atomic media have their big advantages in all-optical tunable and reconfigurable features in real-time control as compared to solid-state systems. It is therefore of valuable importance to extend the study of 𝒫𝒯-symmetry to this area. Recently several works have come into sight [26–30 ] on this aspect. In mapping Schrödinger equations to paraxial wave propagation equations, the role of time variable t is cast to the propagation direction z, indicating that the 𝒫𝒯-symmetric potential V(x) manifests in the transversely modulated refractive index n(x). Many works have been done under this consideration, e.g., to study a pair of coupled waveguides with balanced gain and loss in the transverse plane of wave propagation [6,8,22,31–33 ]. Recently, the question has been asked about what if we implant the 𝒫𝒯-symmetric modulation of complex refractive index to the longitudinal direction z instead of the transverse plane x [34, 35]. Considering n=1+χp1+χp/2, 1D atomic lattices have been designed with the N level configuration to realize the 𝒫𝒯-antisymmetric modulation of probe susceptibility χp(z)=χp*(z) in the absence of optical gain [29]. This scheme has the obvious flexibility in real-time control, all-optical tuning and reconfiguration and thus is more desired in the generation, manipulation, and application of optical 𝒫𝒯-antisymmetry.

In this work, we propose a new scheme of 1D atomic lattices for realizing both 𝒫𝒯-symmetry χp(z)=χp*(z) and 𝒫𝒯-antisymmetry χp(z)=χp*(z) along the lattice direction z by exploiting or avoiding the Raman gain. Here 𝒫𝒯-symmetry and 𝒫𝒯-antisymmetry are achieved via a careful sinusoidal modulation of one driving field in terms of intensity (Rabi frequency) or frequency (atom-field detuning) in the N level configuration. But it is also essential to have a Gaussian distribution of atomic density in each period of dipole traps forming the 1D optical lattices. A 𝒫𝒯-symmetric or 𝒫𝒯-antisymmetric metamaterial can be regarded as a complex photonic crystal because both real (χ′ p) and imaginary (χ″ p) susceptibilities are modulated with a common period. The critical point is that such twofold spatial modulations exhibit a fixed π/2 phase shift [36] so that χ′ p and χ″ p are even and odd functions or vice versa of lattice position z. These results may be extended to develop functional devices like photonic diodes and transistors, an impossible task utilizing real photonic crystals with only χ′ p being nonvanishing and periodically modulated in space. The realization of such devices needs to completely break the optical reciprocity of light transport in a controlled way. Though 𝒫𝒯-symmetric or 𝒫𝒯-antisymmetric metamaterials support unidirectional light reflection [16, 29], it is rather involved to attain nonreciprocal light transmission so that temporal modulations may be further required in addition to spatial modulations.

2. Model and equations

We consider here a coupled system of 1D optical lattices composed of Gaussian-distributed bunches of cold 87Rb atoms, with each bunch seated at the bottom of a dipole trap along the z direction. All atoms are coherently driven into the N configuration by three coherent laser fields at frequencies (amplitudes) ωp (E p), ωc (E c) and ωd (E d), as shown in Fig. 1. The N-configuration consists of two ground levels |1〉 and |2〉 and two excited levels |3〉 and |4〉, where the weak probe field ωp interacts with transition |1〉 ↔ |4〉 while the strong pump fields ωc and ωd act upon transitions |1〉 ↔ |3〉 and |2〉 ↔ |4〉, respectively. The corresponding frequency detunings (Rabi frequencies) are defined as Δp = ωpω 41p = E p · d 14/), Δc = ωcω 31c = E c · d 13/), and Δd = ωdω 42d = E d · d 24/) with ωij being resonant transition frequencies and d ij relevant dipole moments.

 figure: Fig. 1

Fig. 1 (a) Diagram of a four-level N configuration of ultracold atoms, which are driven by a weak probe field Ωp, a strong coupling field Ωc, and a strong driving field Ωd. (b) An ensemble of N-type ultracold atoms trapped at the bottoms of dipole traps forming 1D optical lattices in the z direction. The atomic density in each dipole trap of spatial period a is assumed to exhibit a Gaussian distribution.

Download Full Size | PDF

With the rotating-wave and electric-dipole approximations, we can write down the interaction Hamiltonian and then obtain the density matrix equations

tρ11=+Γ31ρ33+Γ41ρ44+iΩc*ρ31iΩcρ13+iΩp*ρ41iΩpρ14tρ22=+Γ32ρ33+Γ42ρ44+iΩd*ρ42iΩdρ24tρ33=Γ31ρ33Γ32ρ33+iΩcρ13iΩc*ρ31tρ12=γ12ρ12+iΩc*ρ32+iΩp*ρ42iΩdρ14tρ13=γ13ρ13+iΩp*ρ43+iΩc*(ρ33ρ11)tρ14=γ14ρ14+iΩc*ρ34+iΩp*(ρ44ρ11)iΩd*ρ12tρ23=γ23ρ23+iΩd*ρ43iΩc*ρ21tρ24=γ24ρ24iΩp*ρ21+iΩd*(ρ44ρ22)tρ34=γ34ρ34+iΩcρ14iΩp*ρ31iΩd*ρ32,
where γ′ 12 = γ 12 + iΔ12, γ′ 13 = γ 13 + iΔc, γ′ 14 = γ 14 + iΔp, γ′ 23 = γ 23 + iΔ23, γ′ 24 = γ 24 + iΔd, γ′ 34 = γ 34 + iΔ34; Δ12 = Δp − Δd, Δ23 = Δc + Δd − Δp, Δ34 = Δp − Δc. We also assume Γ31 = Γ32 = Γ41 = Γ42 = γ and γ 12 << γ for simplicity, which then yields γ 13 = γ 14 = γ 23 = γ 24 = γ 34/2 = γ. Closure of this atomic system further requires ρij=ρji* and ρ 11 + ρ 22 + ρ 33 + ρ 44 = 1.

In the weak probe limit (Ωpγ), it is viable to obtain first the zero-order solutions and then the first-order solutions of Eqs. (1) in the steady state in terms of Ωp. We are only interested in the optical properties experienced by the probe field, so we obtain here

ρ41(1)=Ωpγ23(Ωc2+γ21γ23)[(Ωc2Ωd2+γ21γ23)Ωcρ13(0)+(Ωd2Ωc2+γ43γ23)Ωdρ24(0)](γ23Ωd2+γ41Ωc2+γ21γ23γ41)(γ21γ23Ωd2+γ43γ23Ωc2+γ21γ43γ232)+(Ωc2Ωd2+γ21γ23)2γ23Ωc2+Ωpi(γ21γ23Ωd2+γ43γ23Ωc2+γ21γ43γ232)(Ωc2+γ21γ23)(ρ44(0)ρ11(0))(γ23Ωd2+γ41Ωc2+γ21γ23γ41)(γ21γ23Ωd2+γ43γ23Ωc2+γ21γ43γ232)+(Ωc2Ωd2+γ21γ23)2γ23Ωc2
with
ρ13(0)=Ωc*(Δc+iγ)Ωd2Ωd2(2Ωc2+Δc2+γ2)+Ωc2(2Ωd2+Δd2+γ2)ρ24(0)=Ωd*(Δd+iγ)Ωc2Ωd2(2Ωc2+Δc2+γ2)+Ωc2(2Ωd2+Δd2+γ2)ρ11(0)=(Ωc2+Δc2+γ2)Ωd2Ωd2(2Ωc2+Δc2+γ2)+Ωc2(2Ωd2+Δd2+γ2)ρ44(0)=Ωc2Ωd2Ωd2(2Ωc2+Δc2+γ2)+Ωc2(2Ωd2+Δd2+γ2)
from which we can examine the probe susceptibility
χp=χp+iχp=N0d1422ε0h¯ρ41(1)
with N 0 denoting the volume density of a homogeneous atomic sample. It is clear that no population inversion ( ρ44(0)<ρ11(0)) can occur on the probe transition, whereas atomic coherence ρ13(0) and ρ24(0) may result in a Raman gain (χ″ p > 0). In the case of 1D atomic lattices under consideration, a constant density N 0 should be replaced, e.g., by a Gaussian distribution
Ni(z)=N0e(zzi)2/σ2,z(zia/2,zi+a/2)
in the ith lattice. Here zi is the ith center of 1D atomic lattices of period a and σ is the standard deviation of the Gaussian distribution from its peak density.

Then we consider the following two cases where either Ωc (i) or Δd (ii) is periodically modulated in a standing-wave (SW) fashion along the z direction. In case (i), the coupling field is modulated in amplitude as

Ωc(z)=Ωc0+δΩcsin[2π(zz0)/a]
with Δc = Δd = 0 just for simplicity. In case (ii), the driving field is modulated in frequency as
Δd(z)=Δd0sin[2π(zz0)/a]
with Ωc = Ωd just for simplicity. The amplitude modulation in Eq. (6) is straightforward by simply applying an imperfect SW driving field with unequal forward and backward components. The frequency modulation in Eq. (7) is somewhat intractable and needs, e.g., an additional SW field to induce a spatially periodic Stark shift of level |2〉 [26]. With such sinusoidal modulations, optical 𝒫𝒯-symmetry (𝒫𝒯-antisymmetry) will be attained if χ′ p(z) and χ″ p(z) are found to be an odd (even) and an even (odd) function of lattice position z. This will be examined via numerical calculations in the next section with suitable parameters.

3. Results and discussion

Based on Eq. (2)(7), we can implement numerical calculations to examine whether χ′ p(z) and χ″ p(z) can exhibit a fixed π/2 phase shift in spatial modulation for the two specific cases mentioned above.

We start by plotting in Figs. 2(a) and 2(b) the probe absorption and dispersion properties for a homogeneous atomic sample without suffering any spatial modulations (Ωc(z) ≡ Ωc0 and Δd(z) Δd0). It is obvious from the black curve in Fig. 2(a) that we observe a typical EIT spectrum with χ″ p being negative (indicating loss) around the EIT window center when both Ωc and Ωd are not too large. With the increasing of Ωc, the blue and red curves show that χ″ p becomes positive (indicating gain) at both sides of the EIT window center to yield the spectra of coherent Raman gain without population inversion. Accordingly, we find from Fig. 2(b) that χ′ p changes from positive dispersion (indicating slow light) to negative dispersion (indicating fast light) near the EIT window center as Ωc is increased. Then we expect optical 𝒫𝒯-symmetry (requiring both gain and loss) and 𝒫𝒯-antisymmetry (requiring only absorption) may be attained by choosing relevant parameters with suitable modulations, e.g., as in Eq. (6) or Eq. (7).

 figure: Fig. 2

Fig. 2 Imaginary (a) and real (b) parts of probe susceptibility χp vs. probe detuning Δp for a homogeneous sample with atomic density N 0 = 5.0 × 1012 cm−3 and dipole moment d 14 = 2.0 × 10−29 C·m. Other parameters are Γ31 = Γ32 = Γ41 = Γ42 = 3.0 MHz, γ 12 = 2.0 kHz, Δc = Δd = 0, Ωd = 2.0 MHz, and Ωc = 2.0 (black); 2.5 (blue); 3.0 (red).

Download Full Size | PDF

3.1. 𝒫𝒯-symmetric susceptiblities

In Figs. 3(a)–3(d) we plot imaginary (χ″ p) and real (χ′ p) parts of the probe susceptibility in two periods of optical lattices with the trapped atoms exhibiting distributions as in Eq. (5) and the coupling field periodically modulated as in Eq. (6). The two upper panels of 2D plots clearly show that χ″ p and χ′ p are modulated along the z direction in different ways and χ″ p seems more sensitive than χ′ p to the change of probe detuning Δp. We further find from the two lower panels of 1D plots that χ″ p(z) is an odd function while χ′ p(z) is an even function, for Δp = 3.284 MHz, in reference to the ith lattice center z = zi. This is a direct evidence for the realization of optical 𝒫𝒯-symmetry in terms of the probe susceptibility. The alternation of χ″ p > 0 in one half period (e.g., 0.0 < (zzi)/a < +0.5) and χ″ p < 0 in another half period (e.g., −0.5 < (zzi)/a < 0.0) indicates a balanced gain and loss on the whole. To understand why Δp = 3.284 MHz is chosen to demonstrate optical 𝒫𝒯-symmetry, we need to run back over Fig. 2(a) where χ″ p → 0.0 at this point for Ωc = 3.0 MHz and Ωd = 2.0 MHz. It is thus easy to imagine that we can realize χ″ p > 0.0 for Ωc > 3.0 MHz and χ″ p < 0.0 for Ωc < 3.0 MHz at Δp = 3.284 MHz. Accordingly, optical 𝒫𝒯-symmetry is attained at this point with a balanced gain and loss when the coupling field is modulated as in Eq. (6) with Ωc0 = 3.0 MHz and δΩc = 0.3 MHz.

 figure: Fig. 3

Fig. 3 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi)/a and probe detuning Δp; imaginary (c) and real (d) parts of probe susceptibility χp vs. lattice position (zzi)/a with probe detuning Δp = 3.284 MHz. The coupling field is modulated as in Eq. (5) with Ωc0 = 3.0 MHz and δΩc = 0.3 MHz. Other parameters used in the calculations are the same as in Fig. 2 except N 0 = 5.0 × 1012 cm−3 and σ = 0.2a for the Gaussian density distribution in Eq. (5).

Download Full Size | PDF

In Figs. 4(a) and 4(b) we further check what will happen when the coupling field shows a larger modulation amplitude δΩc. It is clear that χ′ p is not sensitive to δΩc while χ″ p changes remarkably with the increasing of δΩc. The black curve in Fig. 4(a) shows, in particular, that the maximal gain in one half period is obviously lower than the maximal loss in another half period for a sufficiently large δΩc so that optical 𝒫𝒯-symmetry is destroyed somewhat in the presence of an unbalanced gain and loss. This fact indicates that the response of χ″ p to δΩc enters an unexpected nonlinear regime, e.g., with δΩc > 0.6 MHz, for the realization of optical 𝒫𝒯-symmetry.

 figure: Fig. 4

Fig. 4 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi)/a for Δp = 3.284 MHz and δΩc = 0.3 MHz (red); δΩc = 0.6 MHz (blue); δΩc = 0.9 MHz (black) with other parameters being the same as in Fig. 3.

Download Full Size | PDF

3.2. 𝒫𝒯-antisymmetric susceptibilities

In Figs. 5(a)–5(d) we make similar plots as in Figs. 3(a)–3(d) with the Gaussian distribution of atomic density as in Eq. (5) and the driving field periodically modulated instead as in Eq. (7). The two upper panels of 2D plots once again show that χ″ p and χ′ p are modulated along the z direction in different ways and χ″ p seems more sensitive than χ′ p to the change of probe detuning Δp. The two lower panels of 1D plots further show that χ″ p(z) is an even function while χ′ p(z) is an odd function, for Δp = 0.0 MHz, in reference to the ith lattice center z = zi. This is a direct evidence for the realization of optical 𝒫𝒯-antisymmetry in terms of the probe susceptibility. It is clear that optical 𝒫𝒯-antisymmetry requires the alternation of χ′ p > 0 in one half period and χ′ p < 0 in another half period with varying absorptive loss in each period of the considered atomic lattices. To understand why Δp = 0.0 MHz is chosen to demonstrate optical 𝒫𝒯-antisymmetry, we once again run back over Fig. 2(a) where χ″ p → 0.0 at this point for Ωc = 2.0 MHz and Ωd = 2.0 MHz. An introduction of nonzero detuning Δd0 thus will result in the space-dependent probe absorption characterized with χ″ p(z) < 0 at Δp = 0.0 MHz everywhere in each period. Accordingly, optical 𝒫𝒯-antisymmetry is attained at this point in the absence of gain when the driving field is modulated as in Eq. (7) with Δd0 = 1.0 MHz.

 figure: Fig. 5

Fig. 5 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi)/a and probe detuning Δp; imaginary (c) and real (d) parts of probe susceptibility χp vs. lattice position (zzi)/a with probe detuning Δp = 0.0 MHz. The driving field is modulated as in Eq. (7) with Δd0 = 1.0 MHz. Other parameters used in the calculations are the same as in Fig. 2 except Ωc = 2.0 MHz as well as N 0 = 5.0 × 1012 cm−3 and σ = 0.2a for the Gaussian density distribution in Eq. (5).

Download Full Size | PDF

In Figs. 6(a) and 6(b) we make similar plots as in Figs. 4(a) and 4(b) for different values of modulation detuning Δd0 of the driving field. It is clear that both χ′ p and χ″ p change remarkably, but to different extent, with the increasing of Δd0. It is more important that the maximal amplitude of χ′ p > 0 in one half period is always the same as the maximal amplitude of χ′ p < 0 in another half period. This means that optical 𝒫𝒯-antisymmetry is well reserved no matter Δd0 is small or large. The reason is simply that the increasing of Δd0 will weaken the optical pumping and therefore not result in a nonlinear response of χ′ p to Δd0 for the realization of optical 𝒫𝒯-antisymmetry. Finally we note that the approach shown here for achieving optical 𝒫𝒯-antisymmetry is similar to that in Ref. [29] with also regard to four-level N-type atomic lattices. The main difference rests with two facts: (i) the probe, coupling, and driving fields are applied on three transitions in a way different from that in Ref. [29] where no population pumping occur and probe gain is always absent; (ii) the driving field is near resonant and modulated in frequency here but far detuned and modulated in amplitude in Ref. [29].

 figure: Fig. 6

Fig. 6 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi)/a for Δp = 0.0 MHz and Δd0 = 1.0 MHz (red); Δd0 = 1.5 MHz (blue); Δd0 = 2.0 MHz (black) with other parameters being the same as in Fig. 5.

Download Full Size | PDF

4. Conclusions

In summary, we have studied the realization of optical 𝒫𝒯-symmetry and 𝒫𝒯-antisymmetry in 1D atomic lattices as far as the probe susceptibility χp is concerned. Given that cold atoms in each dipole trap are driven into the N level configuration and have a Gaussian density distribution, we find real (χ′ p) and imaginary (χ″ p) parts of the probe susceptibility can be, respectively, an even function and an odd function or vice versa of the lattice position when the coupling or driving field is modulated along the z direction in a suitable way. This indicates optical 𝒫𝒯-symmetry requires as usual the alternation of balanced gain and loss in each period with positive χ′ p everywhere, whereas optical 𝒫𝒯-antisymmetry requires the alternation of balanced χ′ p > 0 and χ′ p < 0 in each period with absorptive loss everywhere. Such complex photonic crystals of driven atomic lattices have the obvious advantages of real-time, all-optical tunable and reconfigurable features and may be used to develop functional devices inaccessible for real photonic crystals.

Acknowledgments

This work is supported by National Natural Science Foundation (Nos. 61378094 and 11534002) and National Basic Research Program (No. 2011CB921603) of China.

References and links

1. C. M. Bender and S. Boettcher, “Real spectra in non-Hermitian Hamiltonians having 𝒫𝒯 symmetry,” Phys. Rev. Lett. 80, 5243–5246 (1998). [CrossRef]  

2. C. M. Bender, S. Boettcher, and P. N. Meisinger, “𝒫𝒯-symmetric quantum mechanics,” J. Math. Phys. 40, 2201–2229 (1999). [CrossRef]  

3. A. Ruschhaupt, F. Delgado, and J. G. Muga, “Physical realization of 𝒫𝒯-symmetric potential scattering in a planar slab waveguide,” J. Phys. A 38, L171–L176 (2005). [CrossRef]  

4. K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and Z. H. Musslimani, “Beam dynamics in 𝒫𝒯 symmetric optical lattices,” Phys. Rev. Lett. 100, 103904 (2008). [CrossRef]  

5. A. Regensburger, C. Bersch, M.-A. Miri, G. Onishchukov, D. N. Christodoulides, and U. Peschel, “Parity-time synthetic photonic lattices,” Nature 488, 167–171 (2012). [CrossRef]   [PubMed]  

6. S. Klaiman, U. Günther, and N. Moiseyev, “Visualization of branch points in 𝒫𝒯-symmetric waveguides,” Phys. Rev. Lett. 101, 080402 (2008). [CrossRef]  

7. S. Longhi, “Bloch oscillations in complex crystals with 𝒫𝒯 symmetry,” Phys. Rev. Lett. 103, 123601 (2009). [CrossRef]  

8. C. E. Rüter, K. G. Makris, R. El-Ganainy, D. N. Christodoulides, M. Segev, and D. Kip, “Observation of parity-time symmetry in optics,” Nat. Phys. 6, 192–195 (2010). [CrossRef]  

9. L. Feng, M. Ayache, J. Q. Huang, Y. L. Xu, M. H. Lu, Y. F. Chen, Y. Fainman, and A. Scherer, “Nonreciprocal light propagation in a silicon photonic circuit,” Science 333, 729–733 (2011). [CrossRef]   [PubMed]  

10. N. I. Landy, S. Sajuyigbe, J. J. Mock, D. R. Smith, and W. J. Padilla, “Perfect metamaterial absorber,” Phys. Rev. Lett. 100, 207402 (2008). [CrossRef]   [PubMed]  

11. Y.-D. Chong, L. Ge, H. Cao, and A. D. Stone, “Coherent perfect absorbers: Time-reversed lasers,” Phys. Rev. Lett. 105, 053901 (2010). [CrossRef]   [PubMed]  

12. S. Longhi, “𝒫𝒯-symmetric laser absorber,” Phys. Rev. A 82, 031801 (2010). [CrossRef]  

13. Y. Sun, W. Tan, H. Q. Li, J. Li, and H. Chen, “Experimental demonstration of a coherent perfect absorber with 𝒫𝒯 phase transition,” Phys. Rev. Lett. 112, 143903 (2014). [CrossRef]  

14. M. Kulishov, J. M. Laniel, N. Bélanger, J. Azaña, and D. V. Plant, “Nonreciprocal waveguide Bragg gratings,” Opt. Express 13, 3068–3078 (2005). [CrossRef]   [PubMed]  

15. S. Longhi, “Invisibility in 𝒫𝒯-symmetric complex crystals,” J. Phys. A 44, 485302 (2011). [CrossRef]  

16. Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, “Unidirectional invisibility induced by 𝒫𝒯-symmetric periodic structures,” Phys. Rev. Lett. 106, 213901 (2011). [CrossRef]  

17. L. Feng, Y.-L. Xu, W. S. Fegadolli, M.-H. Lu, J. E. B. Oliveira, V. R. Almeida, Y.-F. Chen, and A. Scherer, “Experimental demonstration of a unidirectional reflectionless parity-time metamaterial at optical frequencies,” Nat. Mater. 12, 108–113 (2013). [CrossRef]  

18. B. He, S.-B Yan, J. Wang, and M. Xiao, “Quantum noise effects with Kerr-nonlinearity enhancement in coupled gain-loss waveguides,” Phys. Rev. A 91, 053832 (2015). [CrossRef]  

19. H. Benisty, A. Lupu, and A. Degiron, “Transverse periodic 𝒫𝒯 symmetry for modal demultiplexing in optical waveguides,” Phys. Rev. A 91, 053825 (2015). [CrossRef]  

20. A. Szameit, M. C. Rechtsman, O. Bahat-Treidel, and M. Segev, “𝒫𝒯-symmetry in honeycomb photonic lattices,” Phys. Rev. A 84, 021806 (2011). [CrossRef]  

21. M. Turduev, M. Botey, I. Giden, R. Herrero, H. Kurt, E. Ozbay, and K. Staliunas, “Two-dimensional complex parity-time-symmetric photonic structures,” Phys. Rev. A 91, 023825 (2015). [CrossRef]  

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

23. H. Jing, S. K. Ozdemir, X.-Y. Lü, J. Zhang, L. Yang, and F. Nori, “𝒫𝒯-symmetric phonon laser,” Phys. Rev. Lett. 113, 053604 (2014). [CrossRef]  

24. X.-Y Lü, H. Jing, J.-Y. Ma, and Y. Wu, “𝒫𝒯-symmetry-breaking chaos in optomechanics,” Phys. Rev. Lett. 114, 253601 (2015). [CrossRef]  

25. X.-W. Xu, Y.-X. Liu, C.-P. Sun, and Y. Li, “Mechanical 𝒫𝒯 symmetry in coupled optomechanical systems,” Phys. Rev. A 92, 013852 (2015). [CrossRef]  

26. C. Hang, G.-X. Huang, and V. V. Konotop, “𝒫𝒯 symmetry with a system of three-level atoms,” Phys. Rev. Lett. 110, 083604 (2013). [CrossRef]  

27. J.-T. Sheng, M.-A. Miri, D. N. Christodoulides, and M. Xiao, “𝒫𝒯-symmetric optical potentials in a coherent atomic medium,” Phys. Rev. A 88, 041803 (2013). [CrossRef]  

28. H.-J. Li, J.-P. Dou, and G.-X. Huang, “𝒫𝒯 symmetry via electromagnetically induced transparency,” Opt. Express 21, 32053–32062 (2013). [CrossRef]  

29. J.-H. Wu, M. Artoni, and G. C. La Rocca, “Non-Hermitian degeneracies and unidirectional reflectionless atomic lattices,” Phys. Rev. Lett. 113, 123004 (2014). [CrossRef]   [PubMed]  

30. J.-H. Wu, M. Artoni, and G. C. La Rocca, “Parity-time-antisymmetric atomic lattices without gain,” Phys. Rev. A 91, 033811 (2015). [CrossRef]  

31. K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and Z. H. Musslimani, “𝒫𝒯-symmetric optical lattices,” Phys. Rev. A 81, 063807 (2010). [CrossRef]  

32. Z. H. Musslimani, K. G. Makris, R. El-Ganainy, and D. N. Christodoulides, “Optical solitons in 𝒫𝒯 periodic potentials,” Phys. Rev. Lett. 100, 030402 (2008). [CrossRef]  

33. A. Guo, G. J. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. A. Siviloglou, and D. N. Christodoulides, “Observation of 𝒫𝒯-symmetry breaking in complex optical potentials,” Phys. Rev. Lett. 103, 093902 (2009). [CrossRef]  

34. G. Della Valle and S. Longhi, “Spectral and transport properties of time-periodic 𝒫𝒯-symmetric tight-binding lattices,” Phys. Rev. A 87, 022119 (2013). [CrossRef]  

35. S. Nixon and J. Yang, “Light propagation in periodically modulated complex waveguides,” Phys. Rev. A 91, 033807 (2015). [CrossRef]  

36. S. A. Horsley, M. Artoni, and G. C. La Rocca, “Spatial Kramers-Kronig relations and the reflection of waves,” Nat. Photonics 9, 436–439 (2015). [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 (6)

Fig. 1
Fig. 1 (a) Diagram of a four-level N configuration of ultracold atoms, which are driven by a weak probe field Ω p , a strong coupling field Ω c , and a strong driving field Ω d . (b) An ensemble of N-type ultracold atoms trapped at the bottoms of dipole traps forming 1D optical lattices in the z direction. The atomic density in each dipole trap of spatial period a is assumed to exhibit a Gaussian distribution.
Fig. 2
Fig. 2 Imaginary (a) and real (b) parts of probe susceptibility χp vs. probe detuning Δ p for a homogeneous sample with atomic density N 0 = 5.0 × 1012 cm−3 and dipole moment d 14 = 2.0 × 10−29 C·m. Other parameters are Γ31 = Γ32 = Γ41 = Γ42 = 3.0 MHz, γ 12 = 2.0 kHz, Δ c = Δ d = 0, Ω d = 2.0 MHz, and Ω c = 2.0 (black); 2.5 (blue); 3.0 (red).
Fig. 3
Fig. 3 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi )/a and probe detuning Δ p ; imaginary (c) and real (d) parts of probe susceptibility χp vs. lattice position (zzi )/a with probe detuning Δ p = 3.284 MHz. The coupling field is modulated as in Eq. (5) with Ω c0 = 3.0 MHz and δΩ c = 0.3 MHz. Other parameters used in the calculations are the same as in Fig. 2 except N 0 = 5.0 × 1012 cm−3 and σ = 0.2a for the Gaussian density distribution in Eq. (5).
Fig. 4
Fig. 4 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi )/a for Δ p = 3.284 MHz and δΩ c = 0.3 MHz (red); δΩ c = 0.6 MHz (blue); δΩ c = 0.9 MHz (black) with other parameters being the same as in Fig. 3.
Fig. 5
Fig. 5 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi )/a and probe detuning Δ p ; imaginary (c) and real (d) parts of probe susceptibility χp vs. lattice position (zzi )/a with probe detuning Δ p = 0.0 MHz. The driving field is modulated as in Eq. (7) with Δ d0 = 1.0 MHz. Other parameters used in the calculations are the same as in Fig. 2 except Ω c = 2.0 MHz as well as N 0 = 5.0 × 1012 cm−3 and σ = 0.2a for the Gaussian density distribution in Eq. (5).
Fig. 6
Fig. 6 Imaginary (a) and real (b) parts of probe susceptibility χp vs. lattice position (zzi )/a for Δ p = 0.0 MHz and Δ d0 = 1.0 MHz (red); Δ d0 = 1.5 MHz (blue); Δ d0 = 2.0 MHz (black) with other parameters being the same as in Fig. 5.

Equations (7)

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

t ρ 11 = + Γ 31 ρ 33 + Γ 41 ρ 44 + i Ω c * ρ 31 i Ω c ρ 13 + i Ω p * ρ 41 i Ω p ρ 14 t ρ 22 = + Γ 32 ρ 33 + Γ 42 ρ 44 + i Ω d * ρ 42 i Ω d ρ 24 t ρ 33 = Γ 31 ρ 33 Γ 32 ρ 33 + i Ω c ρ 13 i Ω c * ρ 31 t ρ 12 = γ 12 ρ 12 + i Ω c * ρ 32 + i Ω p * ρ 42 i Ω d ρ 14 t ρ 13 = γ 13 ρ 13 + i Ω p * ρ 43 + i Ω c * ( ρ 33 ρ 11 ) t ρ 14 = γ 14 ρ 14 + i Ω c * ρ 34 + i Ω p * ( ρ 44 ρ 11 ) i Ω d * ρ 12 t ρ 23 = γ 23 ρ 23 + i Ω d * ρ 43 i Ω c * ρ 21 t ρ 24 = γ 24 ρ 24 i Ω p * ρ 21 + i Ω d * ( ρ 44 ρ 22 ) t ρ 34 = γ 34 ρ 34 + i Ω c ρ 14 i Ω p * ρ 31 i Ω d * ρ 32 ,
ρ 41 ( 1 ) = Ω p γ 23 ( Ω c 2 + γ 21 γ 23 ) [ ( Ω c 2 Ω d 2 + γ 21 γ 23 ) Ω c ρ 13 ( 0 ) + ( Ω d 2 Ω c 2 + γ 43 γ 23 ) Ω d ρ 24 ( 0 ) ] ( γ 23 Ω d 2 + γ 41 Ω c 2 + γ 21 γ 23 γ 41 ) ( γ 21 γ 23 Ω d 2 + γ 43 γ 23 Ω c 2 + γ 21 γ 43 γ 23 2 ) + ( Ω c 2 Ω d 2 + γ 21 γ 23 ) 2 γ 23 Ω c 2 + Ω p i ( γ 21 γ 23 Ω d 2 + γ 43 γ 23 Ω c 2 + γ 21 γ 43 γ 23 2 ) ( Ω c 2 + γ 21 γ 23 ) ( ρ 44 ( 0 ) ρ 11 ( 0 ) ) ( γ 23 Ω d 2 + γ 41 Ω c 2 + γ 21 γ 23 γ 41 ) ( γ 21 γ 23 Ω d 2 + γ 43 γ 23 Ω c 2 + γ 21 γ 43 γ 23 2 ) + ( Ω c 2 Ω d 2 + γ 21 γ 23 ) 2 γ 23 Ω c 2
ρ 13 ( 0 ) = Ω c * ( Δ c + i γ ) Ω d 2 Ω d 2 ( 2 Ω c 2 + Δ c 2 + γ 2 ) + Ω c 2 ( 2 Ω d 2 + Δ d 2 + γ 2 ) ρ 24 ( 0 ) = Ω d * ( Δ d + i γ ) Ω c 2 Ω d 2 ( 2 Ω c 2 + Δ c 2 + γ 2 ) + Ω c 2 ( 2 Ω d 2 + Δ d 2 + γ 2 ) ρ 11 ( 0 ) = ( Ω c 2 + Δ c 2 + γ 2 ) Ω d 2 Ω d 2 ( 2 Ω c 2 + Δ c 2 + γ 2 ) + Ω c 2 ( 2 Ω d 2 + Δ d 2 + γ 2 ) ρ 44 ( 0 ) = Ω c 2 Ω d 2 Ω d 2 ( 2 Ω c 2 + Δ c 2 + γ 2 ) + Ω c 2 ( 2 Ω d 2 + Δ d 2 + γ 2 )
χ p = χ p + i χ p = N 0 d 14 2 2 ε 0 h ¯ ρ 41 ( 1 )
N i ( z ) = N 0 e ( z z i ) 2 / σ 2 , z ( z i a / 2 , z i + a / 2 )
Ω c ( z ) = Ω c 0 + δ Ω c sin [ 2 π ( z z 0 ) / a ]
Δ d ( z ) = Δ d 0 sin [ 2 π ( z z 0 ) / a ]
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.