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

Nonreciprocal parity-time phase in magnetized waveguides

Open Access Open Access

Abstract

In a single magneto-optical (MO) waveguide, the dispersion of guided bulk wave is reciprocal in the Voigt configuration. Here we show that the parity-time (PT) phase in two coupled MO waveguides can be nonreciprocal if the waveguides are properly biased. The nonreciprocal PT phase is closely related to the asymmetric field profile induced by the MO effect that modifies the coupling strength between adjacent waveguides. We show that it is feasible to switch between broken and conserved PT phases by simply reversing the magnetic bias or the propagating direction of wave. Theoretical analysis and numerical calculation prove our theory. This investigation highlights a flexible method in manipulating the field dynamics of waveguide arrays by using the novel properties of PT phase especially the exceptional points.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Recent years we have witnessed the transfer of concepts between different subjects of physics if they are governed by similar mathematical operations. Especially, with the similarities between the Schrödinger equation for electrons and the Maxwell’s equations for photons, the optical classic analog of quantum physics [1] has become a well-known area of optics that is still undergoing extensive investigation. Nowadays, by noting that loss is unavoidable in optical systems [2] and gain can be easily accessed, scientists switch their attention to optical non-Hermitian systems [3, 4] that can fully take advantages of the bright side of loss. Among all the non-Hermitian systems the parity-time (PT) symmetry [5–8] is maybe the most well-known.

The simplest PT symmetric system is two parallel coupled waveguides [9–15]. People showed that if the waveguides are identical except for equal amounts of gain and loss, the coupled bulk modes could still possess real wavevectors under a conserved PT phase and propagate inside the structure with fixed amplitudes. When the loss/gain rate is greater than the mutual coupling rate κ, the PT phase is broken and the wavevectors are complex. Between these two scenarios is a phase transition point termed the exceptional point (EP) [13–16], where the eigenvalues and eigenvectors coalesce. With many attractive properties of PT symmetry and EPs, e.g. unidirectional reflectionless [17], polarization filtering [18], unidirectional mode excitation [19], anti-lasing [20], perfect optical absorption [21, 22], enhanced sensitivity [23, 24], and even slow light [25], it is not a surprise that people then discussed how to manipulate the PT phase, for example, by tuning the resonant frequency, gain/loss rate, coupling strength, and even by introducing nonlinear optical effects [26–31]. Direction-dependent PT phase transition has also been studied by using dynamic modulation scheme [32].

Here we study the influence of magneto-optical (MO) effect on the PT phase of two adjacent slab waveguides. Usually people pay their attention to the consequence of modified dispersion from the MO effect, or the topological waveguiding mechanism toward defect-immune and/or backscattering-immune propagation [33–35]. In this article we emphasize another important but easily overlooked phenomenon that could render a nonreciprocal PT phase. Although the dispersion of guided bulk mode in a single MO slab waveguide is reciprocal (ω(k)=ω(k)), the profile of field becomes asymmetric with respect to the middle line of the waveguide due to the MO-induced non-reciprocality of the two boundaries [36]. The coupling strength between adjacent waveguides consequently depends on the direction of the bias field B applied to the waveguides and the propagating direction of wave. The dispersion of guided bulk mode and the associated PT phase can be nonreciprocal if the two MO waveguides are properly biased. Such an effect provides us with an interesting way in manipulating the PT phase (especially EPs) by simply switching the applied bias fields B or by reversing the propagating direction of wave. We perform theoretical analysis and numerical simulation to prove our prediction. Potential applications and experimental suggestions are discussed.

 figure: Fig. 1

Fig. 1 A schematic of the structure under investigation and the corresponding simple mechanism responsible to the nonreciprocal PT phase. (a) Type-I configuration in which the forward propagating fields in the two MO waveguides shift toward each other, and the effective coupling rate increases to κ+, greater than κ at B = 0. (b) Type-II configuration, that when incident from the backward direction the opposite effect in the field profile takes place, and the coupling rate κ is weaker than κ. Note that the Type-II configuration can be also achieved from the Type-I configuration by simply reversing the bias fields.

Download Full Size | PDF

This article is organized as following. In Section 2 we develop rigorous formulas for the guided bulk mode, and briefly discuss the main mechanism of the nonreciprocal PT phase. In Section 3 we perform numerical calculation to prove our theory and to provide more information about the influence of MO effect over the PT phase. Discussion about the importance of our investigation is made in Section 4.

2. Theory

The configuration under investigation is shown in Fig. 1. It is a simple system made of two geometrically-identical slab waveguides (with a width of b) placed adjacent with each other (with a distance of 2a) in the xz plane. Without loss of generality we assume that the upper (lower) waveguide WG1 (WG2) provides gain (loss). Medium surrounding these waveguides is air.

Now, let us assume that the dielectric constants ε1,2 of WG1 and WG2 can be changed via the MO effect. In the Voigt configuration where the magnetic bias B is perpendicular to the slab waveguide, i.e. in the y direction, the permittivity tensor becomes anisotropic [36]

ε¯¯1,2=ε0(ε1,20jγ1,20ε||0jγ1,20ε1,2),
where the off-diagonal element γ1,2 is determined not only by the magnitude of B but also its direction [36]. Since here we are interested in the scenarios that B in WG1 and WG2 are equal in amplitude but opposite with each other, we set
γ1=γ2.

Furthermore, to make the analysis simple we assume that the gain and loss are induced by proper imaginary parts of ±δ in the diagonal elements, that

ε1,2=ε0(εm±jδ).

Paying attention to the guided bulk transverse-magnetic (TM) mode (with a wavevector of k) inside the structure, the field component Hy can be expressed as

Hy=ejkz+jωt{H1eα(xab),x>a+bH2ejβ1(xa)+H3e+jβ1(xa),a+b>x>aH4eαx+H5e+αx,+a>x>aH6ejβ2(x+a)+H7e+jβ2(x+a),a>x>abH8e+α(x+a+b),x<ab

From ×H=jωD we can find the expressions of the other two field components Ex and Ez. Since ×E=jωμ0H should return to Eq. (4), we can prove that

α2=k2k02,
β1,22=ε1,22γ1,22ε1,2k02k2,
where k0=ω/c=2π/λ0, λ0 is the free-space wavelength.

From the boundary continuum conditions on Hy and Ez we can find that the dispersion (ω,k) of the guided bulk mode is governed by an equation of MH=0, as

(1AA100000011BB1000000B1B11000000C1C11M1aAM1bA1000000M1aM1bBB1000000B1BM2aM2b000000M2aC1M2bC1)(H1H2H3H4H5H6H7H8)=0,
where A=exp (jβ1b), B=exp (αa), C=exp (jβ2b),
M1,2a=γ1,2k+jβ1,2ε1,2α(ε1,22γ1,22),
M1,2b=γ1,2kjβ1,2ε1,2α(ε1,22γ1,22).

Before resorting to numerical calculation we would like to briefly discuss the mechanism of the non-reciprocal PT phase within our interest. It is evident that the introduced MO effect renders a nonzero γ1,2 value, which reverses its sign (γ1,2γ1,2) when B changes its direction. However, Eq. (6) has a quadratic dependence on γ1,2, and in a single slab MO waveguide the two air-MO interfaces are time-reversed partners and switch their roles when kk. Consequently, the mode dispersion inside a single MO waveguide is reciprocal [36], that ω(k)=ω(k). However, the broken symmetry over the field profile cannot be ignored. As explained in [36], a nonzero γ value would break the degeneracy of the two air-MO interfaces. As a result, the field profile shifts away from the middle of the waveguide and is no longer symmetric or anti-symmetric (see Fig. 1). When the MO waveguide is oppositely biased (BB), a mirror inversion is imposed over the field profile and the field is switched to other side of the MO waveguide.

Now, by referring to the two configurations shown in Fig. 1 we can see how the PT phase in coupled waveguides is modified by the MO effect. When no magnetic field is applied (γ1,2=0), the system can be described by a 2 × 2 non-Hermitian matrix [6, 7]

(ω0+jgκκω0jg)(ψ1ψ2)=ωPT(ψ1ψ2).
where κ is the coupling rate determined by the overlapping of fields in adjacent waveguides, and g is the gain/loss rate for a given resonance ω0 and is determined by the geometry of the waveguide and the permittivity tensor of Eq. (1). As a standard PT-symmetric system [6, 7], the eigenvalues are characterized by
ωPT±=ω0±κ2g2,
where an EP is obtained when κ=g [16]. When κ>g (κ<g) the solutions are real (complex) and the PT phase is conserved (broken).

When the waveguides are oppositely biased, we can firstly consider the Type-I configuration shown in Fig. 1(a). Now the two otherwise independent bulk modes inside the MO waveguides would be pushed together (see right side of Fig. 1(a)), and the effective coupling rate is increased to κ+>κ. As a result, to reach the EP a larger value of g=κ+ is required.

A totally different scenario is obtained when the wave propagates in the opposite direction (see the Type-II configuration shown in Fig. 1(b)), or when the applied magnetic bias B are reversed. Now the field in each waveguide is switched to the other boundary. In sharp contrast with that of Fig. 1(a), the overlapping of the two bulk modes decreases since the fields are stretched apart, and a smaller effective coupling rate of κ is obtained, i.e. κ<κ. The EP is achieved at another frequency where g=κ.

Since these two configurations are time-reversed partners, once the PT phases of them are not equal, e.g. EP appears at two different frequencies for forward and backward modes, we can make the conclusion that the system is nonreciprocal. Above analysis also proves that the PT phase is reciprocal if the two waveguides are equally biased (γ1=γ2), which is not discussed here.

 figure: Fig. 2

Fig. 2 Dispersions (k0 = ω/c,k) of the guided bulk mode when γ1 = −γ2 = −1, ε1,2 = 4±j0.5, a = 0.3 mm, and b = 3 mm. EPs are labeled out by red stars.

Download Full Size | PDF

3. Simulation

To prove the existence of nonreciprocal PT phase based on the mechanism discussed in the former section, we calculate the dispersion relation (k0=ω/c, k) of guided bulk modes by using Eq. (7). In this section we assume that all the parameters (including εm and δ) in the permittivity tensor are constants and do not vary with the angular frequency ω. It violates the causality principle [37], but makes it easier to observe and analysis the nonreciprocal features in our proof-of-concept numerical investigation. Parameters utilized in our calculation are b = 3

mm, a=0.3 mm, εm=4, δ=0.5, and γ1=γ2=1. The dispersion (k0, k) is found by firstly calculating the determinant of the 8 × 8 matrix M in Eq. (7) versus k0 and k, and then projecting the colored parametric surface of

Ω=log |det {M}|
into the (k0,k) plane. Dispersion curves are represented by shadows in the plots. Broken PT phase cannot be imagined from our calculation because only real k is utilized in this treatment.

 figure: Fig. 3

Fig. 3 Variation of Ω versus k0 when k = 0.785 mm−1 (blue line) and k = −0.785 mm−1 (red line), respectively. In each case two eigenmodes characterized by dips can be observed.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Distributions of Hy (dark) and phase (red, in the unit of π) in the structure at the four eigenmodes of Fig. 3.

Download Full Size | PDF

Figure 2 shows the mode dispersions of the two configurations. We can see in both cases a loop of dispersion similar to that in a PT symmetric system [25] is obtained in the low frequency regime (before EPs). It implies that the system is PT symmetric there. When the magnitude of the wavevector k increases, each dispersion curve coalesces at an EP (red star), across which the PT phase becomes broken. The most important character of Fig. 2 is that the dispersion loop of the Type-II configuration is smaller and narrower than that of the Type-I configuration, and the EP has a smaller k value. Figure 2 verifies that the PT phase of this system is nonreciprocal.

To provide more information about the nonreciprocal PT phase, we plot the curves of Ω versus k0 at k=±0.785 mm 1 in Fig. 3. In each curve we can observe two dips between k0=0.50 mm 1 and k0=0.61 mm 1. Each dip represents a guided bulk mode. It is very clear that the guided resonances are sensitive to the propagating direction of wave, or the magnetic bias.

Figure 4 shows the distributions of field component Hy and the phase of these guided bulk modes. We can see the magnitude of Hy is symmetric with respect to the middle of the structure. However, the phase is not a constant and varies smoothly in the two waveguides. It is in sharp contrast with the scenarios of Hermitian system (δ = 0), in which the phase difference between the two waveguides is either zero or π (not shown here). Furthermore, the distribution of field is evidently sensitive to the sign of k, especially the ones shown in Figs. 4(a) and 4(c). This result proves the mechanism we proposed above.

 figure: Fig. 5

Fig. 5 Variation of Ω versus the distance a and wavevector k at a fixed angular frequency of ω = 2πc/λ0 with λ0 = 8 mm. EPs are labeled out by red stars.

Download Full Size | PDF

Above calculation is based on the assumption that all the parameters are constants and dispersionless. However, such an assumption violates the causality principle [37], which permits the satisfaction of ε1=ε2* only at isolate frequencies. The dispersion in εm and γ due to the cyclotronic motion of effective charges [36] should also be taken into account in experiments. Now we propose to fix the angular frequency ω=ck0 and the waveguide thickness b, and then change the distance 2a to check the nonreciprocal PT phase. This approach obeys the limitations set by the causality principle [37], and can be utilized as a possible guideline in verifying our proposed mechanism in future experiments.

Assuming that εm=4, δ=0.1, and |γ1,2|=1 can be realized at k0=0.785 mm 1 (λ0=8 mm), we analyze the different PT phases in the two configurations. The variation of Ω versus a is shown in Fig. 5. We can see the EPs of the two configurations have different a values, which split the whole figure into three different regions. When a<0.655mm (Region-1) the PT phases are conserved. When a>0.824 mm (Region-3) the phases are broken. Between them (Region-2) the PT phase is broken (conserved) in the Type-II (Type-I) configuration. Region-2 might possess novel applications because here the system can be switched between broken and conserved PT phases by simply reversing the propagating direction of wave, or by reversing the bias fields.

 figure: Fig. 6

Fig. 6 Variation of Ω versus k when a = 0.655 mm. Inset shows the distributions of field and phase for the EP in the Type-II configuration. Here δ = 0.1 is assumed.

Download Full Size | PDF

We check the distributions of field Hy and its phase at EPs of Fig. 5. A standard result is shown in Fig. 6 for a=0.655 mm. Now an EP is obtained in the Type-II configuration, while in the Type-I configuration the PT phase is conserved because two dips can be observed. From the inset we can see the magnitude of Hy is still symmetric distributed. However, in sharp contrast with these shown in Fig. 4, the phase in the two waveguides is jumped by exactly π/2. This is a standard property of EP that reveals its chirality feature [16].

4. Discussion

Above analysis and numerical simulation prove that nonreciprocal PT phase can be obtained in two MO waveguides with properly bias fields. As for potential applications, the nonreciprocal PT phase demonstrated here enables us to extend the state-of-the-art advances of PT symmetry [6, 7]. For example, it might enable us to control the switch of field between two adjacent waveguides.

Using the parameters of Fig. 5 we perform full-wave numerical simulation by using COMSOL Multiphysics 5.4. A y-directional magnetic current is placed in WG1 to excite the field with a wavelength of λ0=8 mm, and the simulation is performed in the frequency domain. Figure 7 shows the results at a=0.655 mm where the Type-II configuration reaches EP while the Type-I configuration holds a conserved PT phase. We can see in the Type-I configuration the field switches periodically between the two MO waveguides due to the beating of the two PT symmetric eigenmodes (see Fig. 7(a)). However, in the Type-II configuration the field is very strong (note that the colormap in Fig. 7(b) is 10 times greater than that in Fig. 7(a)) and is distributed uniformly in the two waveguides. Figure 7(b) demonstrates an standard evidence of EP where only a single chiral solution exists [16], and highlights the presence of slow light [25] because a zero group velocity permits the extreme concentration of field power. Similar effect of efficiently excitation of field at EPs can be found from [38], and the enhanced spontaneous emission by utilizing the high density of states at EPs has been discussed in [39, 40]. Figure 7 indeed proves that thefield dynamics in coupled MO waveguides can be manipulated, for example, by dynamically tuning the bias fields.

 figure: Fig. 7

Fig. 7 COMSOL simulation of the field inside the structure with λ0 = 8 mm and a = 0.655 mm. (a) For a forward propagation the PT phase is conserved and the field switches periodically between the two adjacent MO waveguides. (b) For a backward propagation the field is distributed uniformly in the two waveguides because an EP is reached.

Download Full Size | PDF

Our investigation presented here is about the coupled modes in two parallel waveguides. It is also of great interest to discuss how similar idea performs in multi-waveguide systems, especially the nonreciprocal feature and the emergence of high-order EPs [16, 23, 41, 42].

Experimental realization of the proposed scheme can refer to literatures of PT symmetry and MO effect [33–35, 43]. We believe that the experiments prefer the microwave region where the metal has a low loss and the cyclotronic frequency of electrons is comparable to that of electromagnetic field [36, 43]. The low frequency nature of microwave also enables us to easily control the geometric parameters of the structure especially the distance a. Note that since the ratio of off-diagonal element γ to the diagonal element ε is critical in determining the sensitivity of MO performance [36], we can also utilize the concept of zero-index medium [44–46] to enhance the ratio and make the experimental observation more pronounced. Before finishing this article we also notice that Thomas etc. al. [47] have proposed a user-friendly framework of lumped circuits to study the giant nonreciprocity near EPs in PT symmetric gyrotropic structures in the microwave domain. The feasibility in utilizing this approach to verify our theory is an interesting question to be answered.

5. Conclusion

In summary, we show that the PT phase in coupled MO waveguides is sensitive to the directions of external applied biased fields. Although the bulk mode of a single MO slab waveguide is reciprocal, the coupling between two adjacent MO waveguides leads to two pairs of distinct modes propagating differently along opposite directions. When PT symmetry is introduced, the guided bulk modes can be modified to achieve EPs that exists asymmetrically for forward and backward propagations, which is a by-product of non-reciprocality. To prove that the nonreciprocal PT phase is closely related to the asymmetric field profile induced by the MO effect, we develop proper formula and provide numerical simulations. Potential applications and experimental suggestions are discussed.

Funding

National Natural Science Foundation of China (NSFC) (11574162, 11674244, 11874228).

References

1. S. Longhi, “Quantum-optical analogies using photonic structures,” Laser Photon. Rev. 3, 243–261 (2009). [CrossRef]  

2. J. B. Khurgin, “How to deal with the loss in plasmonics and metamaterials,” Nat. Nanotech. 10, 2–6 (2015). [CrossRef]  

3. C. M. Bender, “Making sense of non-Hermitian Hamiltonians,” Rep. Prog. Phys. 70, 947–1018 (2007). [CrossRef]  

4. H. Cao and J. Wiersig, “Dielectric microcavities: model systems for wave chaos and non-Hermitian physics,” Rev. Mod. Phys. 87, 61–111 (2015). [CrossRef]  

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

6. L. Feng, R. El-Ganainy, and L. Ge, “Non-Hermitian photonics based on parity-time symmetry,” Nat. Photo. 11, 752–762 (2017). [CrossRef]  

7. R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, “Non-Hermitian physics and PT symmetry,” Nat. Phys. 14, 11–19 (2018). [CrossRef]  

8. J. Gear, F. Liu, S. T. Chu, S. Rotter, and J. Li, “Parity-time symmetry from stacking purely dielectric and magnetic slabs,” Phys. Rev. A 91, 033825 (2015). [CrossRef]  

9. R. El-Ganainy, K. G. Makris, D. N. Christodoulides, and Z. H. Musslimani, “Theory of coupled optical PT-symmetric structures,” Opt. Lett. 32, 2632–2634 (2007). [CrossRef]   [PubMed]  

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

11. 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]  

12. 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]  

13. G. W. Hanson and A. B. Yakovlev, “Investigation of mode interaction on planar dielectric waveguides with loss and gain,” Radio Sci. 34, 1349–1359 (1999). [CrossRef]  

14. G. W. Hanson and A. B. Yakovlev, “An analysis of leaky wave dispersion phenomena in the vicinity of cutoff using complex frequency plane singularities,” Radio Sci. 33, 803–819 (1998). [CrossRef]  

15. A. B. Yakovlev and G. W. Hanson, “Fundamental modal phenomena on isotropic and anisotropic planar slab dielectric waveguides,” IEEE Trans. Antennas Propag. 51, 888–897 (2003). [CrossRef]  

16. M.-A. Miri and A. Alù, “Exceptional points in optics and photonics,” Science 363, eaar7709 (2019). [CrossRef]   [PubMed]  

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. M. Lawrence, N. Xu, X. Zhang, L. Cong, J. Han, W. Zhang, and S. Zhang, “Manifestation of PT symmetry breaking in polarization space with terahertz metasurfaces,” Phys. Rev. Lett. 113, 093901 (2014). [CrossRef]  

19. W. Wang, L. Q. Wang, R. D. Xue, H. L. Chen, R. P. Guo, Y. Liu, and J. Chen, “Unidirectional excitation of radiative-loss-free surface plasmon polaritons in PT-symmetric systems,” Phys. Rev. Lett. 119, 077401 (2017). [CrossRef]  

20. W. Wan, Y. Chong, L. Ge, H. Noh, A. D. Stone, and H. Cao, “Time-reversed lasing and interferometric control of absorption,” Science 331, 889–892 (2011). [CrossRef]   [PubMed]  

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

22. D. G. Baranov, A. Krasnok, T. Shegai, A. Alù, and Y. D. Chong, “Coherent perfect absorbers: linear control of light with light,” Nat. Rev. Mater. 2, 17064 (2017). [CrossRef]  

23. H. Hodaei, A. U. Hassan, S. Wittek, H. Garcia-Gracia, R. El-Ganainy, D. N. Christodoulides, and M. Khajavikhan, “Enhanced sensitivity at higher-order exceptional points,” Nature 548, 187–191 (2017). [CrossRef]   [PubMed]  

24. W. Chen, S. K. Ozdemir, G. Zhao, J. Wiersig, and L. Yang, “Exceptional points enhanced sensing in an optical microcavity,” Nature 548, 192–196 (2017). [CrossRef]   [PubMed]  

25. T. Goldzak, A. A. Mailybaev, and N. Moiseyev, “Light stops at exceptional points,” Phys. Rev. Lett. 120, 013901 (2018). [CrossRef]   [PubMed]  

26. Y. Lumer, Y. Plotnik, M. C. Rechtsman, and M. Segev, “Nonlinearly induced PT transition in photonic systems,” Phys. Rev. Lett. 111, 263901 (2013). [CrossRef]  

27. F. Nazari, N. Bender, H. Ramezani, M. K. Moravvej-Farshi, D. N. Christodoulides, and T. Kottos, “Optical isolation via PT-symmetric nonlinear Fano resonances,” Opt. Express 22, 9574 (2014). [CrossRef]   [PubMed]  

28. L. Ge and R. El-Ganainy, “Nonlinear modal interactions in parity-time (PT) symmetric lasers,” Sci. Rep. 6, 24889 (2016). [CrossRef]  

29. X. Zhou and Y. D. Chong, “PT symmetry breaking and nonlinear optical isolation in coupled microcavities,” Opt. Express 24, 6916–6930 (2016). [CrossRef]   [PubMed]  

30. V. V. Konotop, J. Yang, and D. A. Zezyulin, “Nonlinear waves in PT-symmetric systems,” Rev. Mod. Phys. 88, 035002 (2016). [CrossRef]  

31. S. Assawaworrarit, X. Yu, and S. Fan, “Robust wireless power transfer using a nonlinear parity-time-symmetric circuit,” Nature 546, 387–390 (2017). [CrossRef]   [PubMed]  

32. A. Y. Song, Y. Shi, Q. Lin, and S. Fan, “Direction-dependent parity-time phase transition and nonreciprocal amplification with dynamic gain-loss modulation,” Phys. Rev. A 99, 013824 (2019). [CrossRef]  

33. S. A. H. Gangaraj and F. Monticone, “Topological waveguiding near an exceptional point: defect-immune, slow-light, and loss-immune propagation,” Phys. Rev. Lett. 121, 093901 (2018). [CrossRef]  

34. S. A. H. Gangaraj and F. Monticone, “Coupled topological surface modes in gyrotropic structures: Green’s function analysis,” IEEE Ant. Wireless Propag. Lett. 17, 1993–1997 (2018). [CrossRef]  

35. L. Shen, Y. You, Z. Wang, and X. Deng, “Backscattering-immune one-way surface magnetoplasmons at terahertz frequencies,” Opt. Express 23, 950–962 (2015). [CrossRef]   [PubMed]  

36. T. F. Li, T. J. Guo, H. X. Cui, M. Yang, M. Kang, Q. H. Guo, and J. Chen, “Guided modes in magneto-optical waveguides and the role in resonant transmission,” Opt. Express 21, 9563–9572 (2013). [CrossRef]   [PubMed]  

37. A. A. Zyablovsky, A. P. Vinogradov, A. V. Dorofeenko, A. A. Pukhov, and A. A. Lisyansky, “Causality and phase transitions in PT-symmetric optical systems,” Phys. Rev. A 89, 033808 (2014). [CrossRef]  

38. Y. R. Zhang, Z. Z. Zhang, J. Q. Yuan, W. Wang, L. Q. Wang, Z. X. Li, R. D. Xue, and J. Chen, “Parity-time symmetry in periodically curved optical waveguides,” Opt. Express 26, 027141 (2018). [CrossRef]  

39. Z. Lin, A. Pick, M. Loncar, and A. W. Rodriguez, “Enhanced spontaneous emission at third-order Dirac exceptional points in inverse-designed photonic crystals,” Phys. Rev. Lett. 117, 107402 (2016). [CrossRef]   [PubMed]  

40. A. Pick, B. Zhen, O. D. Miller, C. W. Hsu, F. Hernandez, A. W. Rodriguez, M. Soljacic, and S. G. Johnson, “General theory of spontaneous emission near exceptional points,” Opt. Express 25, 012325 (2017). [CrossRef]  

41. W. D. Heiss, “Chirality of wavefunctions for three coalescing levels,” J. Phys. A 41, 244010 (2008). [CrossRef]  

42. Y. R. Zhang, Z. Z. Zhang, J. Q. Yuan, M. Kang, and J. Chen, “High-order exceptional points in non-Hermitian Moire lattices,” Front. Physics 14, 53603 (2019). [CrossRef]  

43. R. P. Guo, L. T. Wu, M. Yang, T. J. Guo, H. X. Cui, X. W. Cao, and J. Chen, “Nonreciprocal propagating electromagnetic modes without phase gradients,” Phys. Rev. A 91, 023808 (2015). [CrossRef]  

44. I. Liberal and N. Engheta, "Near-zero refractive index photonics,” Nat. Photo. 11, 149–158 (2017). [CrossRef]  

45. B. Edwards, A. Alù, M. E. Young, M. Silveirinha, and N. Engheta, “Experimental verification of epsilon-near-zero metamaterial coupling and energy squeezing using a microwave waveguide,” Phys. Rev. Lett. 100, 033903 (2008). [CrossRef]   [PubMed]  

46. Y. Li and C. Argyropoulos, “Exceptional points and spectral singularities in active epsilon-near-zero plasmonic waveguides,” Phys. Rev. B 99, 075413 (2019). [CrossRef]  

47. R. Thomas, H. Li, F. M. Ellis, and T. Kottos, “Giant nonreciprocity near exceptional-point degeneracies,” Phys. Rev. A 94, 043829 (2016). [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 (7)

Fig. 1
Fig. 1 A schematic of the structure under investigation and the corresponding simple mechanism responsible to the nonreciprocal P T phase. (a) Type-I configuration in which the forward propagating fields in the two MO waveguides shift toward each other, and the effective coupling rate increases to κ+, greater than κ at B = 0. (b) Type-II configuration, that when incident from the backward direction the opposite effect in the field profile takes place, and the coupling rate κ is weaker than κ. Note that the Type-II configuration can be also achieved from the Type-I configuration by simply reversing the bias fields.
Fig. 2
Fig. 2 Dispersions (k0 = ω/c,k) of the guided bulk mode when γ1 = −γ2 = −1, ε1,2 = 4±j0.5, a = 0.3 mm, and b = 3 mm. EPs are labeled out by red stars.
Fig. 3
Fig. 3 Variation of Ω versus k0 when k = 0.785 mm−1 (blue line) and k = −0.785 mm−1 (red line), respectively. In each case two eigenmodes characterized by dips can be observed.
Fig. 4
Fig. 4 Distributions of Hy (dark) and phase (red, in the unit of π) in the structure at the four eigenmodes of Fig. 3.
Fig. 5
Fig. 5 Variation of Ω versus the distance a and wavevector k at a fixed angular frequency of ω = 2πc/λ0 with λ0 = 8 mm. EPs are labeled out by red stars.
Fig. 6
Fig. 6 Variation of Ω versus k when a = 0.655 mm. Inset shows the distributions of field and phase for the EP in the Type-II configuration. Here δ = 0.1 is assumed.
Fig. 7
Fig. 7 COMSOL simulation of the field inside the structure with λ0 = 8 mm and a = 0.655 mm. (a) For a forward propagation the P T phase is conserved and the field switches periodically between the two adjacent MO waveguides. (b) For a backward propagation the field is distributed uniformly in the two waveguides because an EP is reached.

Equations (12)

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

ε ¯ ¯ 1 , 2 = ε 0 ( ε 1 , 2 0 j γ 1 , 2 0 ε | | 0 j γ 1 , 2 0 ε 1 , 2 ) ,
γ 1 = γ 2 .
ε 1 , 2 = ε 0 ( ε m ± j δ ) .
H y = e j k z + j ω t { H 1 e α ( x a b ) , x > a + b H 2 e j β 1 ( x a ) + H 3 e + j β 1 ( x a ) , a + b > x > a H 4 e α x + H 5 e + α x , + a > x > a H 6 e j β 2 ( x + a ) + H 7 e + j β 2 ( x + a ) , a > x > a b H 8 e + α ( x + a + b ) , x < a b
α 2 = k 2 k 0 2 ,
β 1 , 2 2 = ε 1 , 2 2 γ 1 , 2 2 ε 1 , 2 k 0 2 k 2 ,
( 1 A A 1 0 0 0 0 0 0 1 1 B B 1 0 0 0 0 0 0 B 1 B 1 1 0 0 0 0 0 0 C 1 C 1 1 M 1 a A M 1 b A 1 0 0 0 0 0 0 M 1 a M 1 b B B 1 0 0 0 0 0 0 B 1 B M 2 a M 2 b 0 0 0 0 0 0 M 2 a C 1 M 2 b C 1 ) ( H 1 H 2 H 3 H 4 H 5 H 6 H 7 H 8 ) = 0 ,
M 1 , 2 a = γ 1 , 2 k + j β 1 , 2 ε 1 , 2 α ( ε 1 , 2 2 γ 1 , 2 2 ) ,
M 1 , 2 b = γ 1 , 2 k j β 1 , 2 ε 1 , 2 α ( ε 1 , 2 2 γ 1 , 2 2 ) .
( ω 0 + j g κ κ ω 0 j g ) ( ψ 1 ψ 2 ) = ω P T ( ψ 1 ψ 2 ) .
ω P T ± = ω 0 ± κ 2 g 2 ,
Ω = log   | det   { M } |
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.