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

Quasinormal modes in transformation media

Open Access Open Access

Abstract

The quasinormal mode (QNM) is generally a mathematical tool in analysing open systems. As the concept of Parity-Time (PT) symmetry was transplanted from quantum mechanics to classical wave systems, massive intriguing phenomena have been found, especially in the vicinity of the so-called exceptional points (EPs). In this paper, with the help of transformation optics (TO), we focus on finding EPs of three-layer PT symmetric models, especially the Metal-Insulator-amplifying Metal (MIaM) ones. Theoretical and numerical methods both indicate that the EPs only exist in the complex frequency regime. Searching within the real frequency spectrum would never find EPs across which complex conjugate wavenumbers occur. Equivalently, only QNMs exist in the symmetry broken phase.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Stemming from quantum mechanics [17], PT symmetry is an important feature of a group of non-Hermitian systems which render exotic physical phenomena beyond the conventional Hermitian ones in both quantum and classical physics [116]. With the balanced gain and loss sufficiently strongly coupled, the system renders real eigenvalue spectrum as a Hermitian system does. However, the fascinating physics emerges just as the equilibrium is broken. The critical points across which the system enters the symmetry-broken phase are the so-called EPs [12]. Therefore, it is a fundamental task to find the EPs before the exotic physics can be revealed. EPs can be retrieved by tuning physical parameters of the system in the phase space. Either increasing the gain and loss or decreasing the coupling strength can lead to the emergence of EPs and the symmetry-broken phase. It is also seen in some papers that EPs exist at some points of the dispersion curves. Such EPs are always accompanied by complex conjugate wavenumbers in the symmetry-broken region while the frequency is treated as the tuning parameter and is kept real through the whole spectrum [1719]. However, it is noteworthy that the gain and loss in this kind of systems are generally constants and added to dielectric layers. This is very different from the frequency-dependent loss or gain of metallic layers. Up to our knowledge, few papers have utilized loss or gain in metallic layers to form the PT symmetric system. It is still interesting to investigate the PT symmetry formed by the balanced gain and loss in disperive metallic layers. Meanwhile, transformation optics (TO) has provided an easy way of conceiving functional structures from the geometric point of view [2022]. It is also a useful method in describing physical phenomena in various length scales [23], from celestial mechanics [24,25] down to nano-optics and plasmonics [2628]. With the so-called complex coordinate transformations [29], it is now rather convenient to conceive non-Hermitian or even PT symmetric structures by geometrical mappings [30]. Similar concepts have also occurred in designing the so-called “Pseudo-Hermitian System” with TO theory [31].

Quasinormal-mode (QNM) theory emerges as a profound method in describing open systems [32]. QNMs per se are modes with complex frequencies which is able to depict the time dependent attenuation or amplification of energy of the system. As an analogy of the normal mode expansion in conservative systems, QNM expansion has been well established and applied to various linear classical wave systems and gravitational systems [33]. Recently, QNM theory has been with the rapid development of the research on nanophotonics and non-Hermitian loss-gain systems [34], so that the conventional theoretical framework would be kept consistent with the cutting-edge of the latest research [35]. It is generally considered that QNMs are more of a theoretical tool in analysing open systems. Nevertheless, a recent work which introduced the concept of “virtual PT symmetry” ingeniously utilized a decaying QNM as the input signal and generated the so-called “virtual gain” within the transmission line model [36]. This pioneering work indicates that QNMs are also revealing its potential application in advanced classical wave systems besides its conventional use as a theoretical tool for open system analysing. This has become part of the motivation of conceiving this paper.

In this paper, we focus on the study of the QNMs that exists in a PT symmetric structure obtained via the complex-coordinate transformation. Not only will the QNM be depicted in terms of its dependence of the geometry and the constitutive parameters of the structure, but the physical mechanism of its existence will also be briefly discussed, aiming at a rather full understanding of QNMs in the model.

2. Transformation optics (TO) framework

Comparing with Ref. [30], we consider a more general coordinate transformation:

$$x' = xp(y,z),\,y' = yq(x,z),\,\,z' = zs(x,y), \quad \left| x \right| \le d,$$
$$x' = x,\,y' = y,\,z' = z, \quad \left| x \right| < d. $$

Here, $(x^{\prime},y^{\prime},z^{\prime})$ and (x, y, z) are the Cartesian coordinates of the auxiliary vacuum space and the physical space. The corresponding Jacobian matrix is formulated as:

$$\mathbf{\Lambda } = \frac{{\partial ({x^{\prime},y^{\prime},z^{\prime}} )}}{{\partial ({x,y,z} )}} = \left( {\begin{array}{ccc} p&{x{p_y}}&{x{p_z}}\\ {y{q_x}}&q&{y{q_z}}\\ {z{s_x}}&{z{s_y}}&s \end{array}} \right)$$

Here, the subscript denotes the partial derivatives. The deduced constitutive parameters in the physical space are written as:

$$\boldsymbol{\mathbf{\epsilon}}(\mathbf{r}) = \mathbf{\mu }(\mathbf{r}) = \left| \mathbf{\Lambda } \right|{\mathbf{\Lambda }^{ - 1}}{\mathbf{\Lambda }^{ - \textrm{T}}}, \quad \left| x \right| \ge d,$$
$${\varepsilon }(\mathbf{r}) = {\mu }(\mathbf{r}) = 1, \quad \left| x \right| < d.$$

Without loss of generality, we aim to focus on two semi-infinite homogeneous slabs with PT symmetry. In this case, the assumed three functions p, q, and s in the coordinate transformation (CT), should firstly be set to

$$p = \left\{ \begin{array}{l} {p_0}\;\;x < - d\\ p_0^\ast \;\;x > \;\;\;d \end{array} \right.,\,\,q = \left\{ \begin{array}{l} {q_0}\;\;x < - d\\ q_0^\ast \;\;x > \;\;\;d \end{array} \right.,\,\,s = \left\{ \begin{array}{l} {s_0}\;\;x < - d\\ s_0^\ast \;\;x > \,\;\;d \end{array} \right. \cdot $$
Here, p0, q0 and s0 are complex constants, the asterisks in the superscript denote the complex conjugate. Thus Eq. (2) reduces to the following form:
$$\mathbf{\Lambda } = \textrm{diag}\,\left[ {p,q,s} \right].$$

Equation (3a) reduces to

$$\boldsymbol{\mathbf{\epsilon}}(\mathbf{r}) = \mathbf{\mu }(\mathbf{r}) = \textrm{diag}\,[{{qs} / p},\,{{ps} / q},\,{{ps} / q}].$$

As we focus on TM modes which is the intrinsic feature of the surface plasmon polaritons (SPPs) at the metal-insulator interface, only the following three parameters are involved

$${\varepsilon _{xx}} = {{qs} / p},\,{\varepsilon _{zz}} = {{pq} / s},\,{\mu _{yy}} = {{ps} / q}.$$

In the optical regime, it is reasonable to set ${\mu _{yy}} = 1$. Meanwhile, as mentioned above, we aimed at homogeneous media, i.e. ${\varepsilon _{xx}} = {\varepsilon _{zz}}$. The two conditions give rise to the following constitutive parameter in the physical space:

$$\textrm{For} \quad \left| x \right| \ge d, \quad {\mu _{yy}} = 1,\,{\varepsilon _{xx}} = {\varepsilon _{zz}} = \left\{ \begin{array}{l} \eta \;\;\;\;\;x \le - d\\ {\eta ^*}\;\;\;\;x \ge \;\;\;d \end{array} \right. \cdot$$
$$\textrm{For} \quad \left| x \right| \le d, \quad \varepsilon = \mu = 1.$$

Here the constants p, q and s are substituted by $\eta = {{qs} / p}$, with p, q, s satisfying $ps = q$ and ${p^2} = {s^2}$.

Up to this stage, based on the complex coordinate transformations, the PT symmetric semi-infinite slabs have been geometrically established (Fig. 1).

 figure: Fig. 1.

Fig. 1. Schematic illustration of the TO framework. Vacuum is the auxiliary space before coordinate transformation. After performing ${r^{\prime}} = {F}({r})$, the system is transformed to the M-I-aM structure, the transformed physical space exhibits PT symmetry.

Download Full Size | PDF

3. QNMs: finding exceptional points (EPs) with causality

As seen from Eq. (8), the two semi-infinite slabs form the PT symmetric model. In this section we adopt realistic dispersive permittivities and search for the EPs. $\eta $ is set to the Drude model permittivity of metals:

$$\eta = 1 - \frac{{\omega _\textrm{p}^2}}{{{\omega ^2} + \textrm{i}\gamma \omega }}. $$
The supported SPPs of the three-layer Metal-Insulator-Metal (MIM) configuration has a well-known dispersion relation [37]:
$${\textrm{e}^{ - 4{\kappa _1}d}} = \frac{{{{{\kappa _1}} / {{\varepsilon _1}}} + {{{\kappa _2}} / {{\varepsilon _2}}}}}{{{{{\kappa _1}} / {{\varepsilon _1}}} - {{{\kappa _2}} / {{\varepsilon _2}}}}}\frac{{{{{\kappa _1}} / {{\varepsilon _1}}} + {{{\kappa _3}} / {{\varepsilon _3}}}}}{{{{{\kappa _1}} / {{\varepsilon _1}}} - {{{\kappa _3}} / {{\varepsilon _3}}}}}. $$
Here, ${\kappa _i}$= $\sqrt {{\beta ^2} - k_0^2{\varepsilon _i}} $ is the decaying rate along x direction, ${\varepsilon _i}$ (i = 1, 2, 3) is the relative permittivity of the ith layer. Assume the system works in the quasi-static regime as the gap will be set to nanoscale. Therefore, we have ${\beta ^2} > > k_0^2{\varepsilon _i}$ (i = 1, 2, 3) and thus ${\kappa _i} \approx \beta $ which leads to the simplified dispersion:
$${\textrm{e}^{ - 4\beta d}} = \frac{{{\varepsilon _2} + {\varepsilon _1}}}{{{\varepsilon _2} - {\varepsilon _1}}}\frac{{{\varepsilon _3} + {\varepsilon _1}}}{{{\varepsilon _3} - {\varepsilon _1}}}. $$
Here $\beta $ denotes the propagation constant along the x-direction. Note that in our model, ${\varepsilon _1} = 1$ and ${\varepsilon _2}$ = $\eta $, ${\varepsilon _3}$ = ${\eta ^\ast }$. This configuration forms a typical PT symmetric system. It is natural that the next step is to find the EPs in the dispersion curve so as to obtain the PT symmetric phase and the PT-broken phase. However, an important attention should be paid before numerically solving Eq. (11). Can we follow Refs. [18,19] and sweep real angular frequency $\omega $ to get a dispersion curve with an EP across which the wavenumber k switches from two real values to complex conjugate pairs? The answer should be no. Previous investigations have also shown that for the frequency dispersive media, one can’t observe the symmetry-breaking transition by simply sweeping the real frequency spectrum [38,39]. As a numerical demonstration, we calculated the dispersion relation of the MIaM model by sweeping real frequency. Silver and gold are set as the metals for the calculation. The plasma frequency ${\omega _\textrm{p}}$ and the damping factor $\gamma $ (collision frequency) of the Drude model of silver (gold) are set to $\hbar {\omega _\textrm{p}}$= 9 eV (8.9 eV), $\hbar \gamma $ = 0.1 eV (0.08 eV) [40]. As denoted by Fig. 2, there exist two dispersion curves: the upper (lower) curve corresponds to the (anti-)symmetric modes. The main (inset) figure is for silver (gold) MIaM. But the EP is found in neither case.

 figure: Fig. 2.

Fig. 2. The dispersion curves of the MIaM systems calculated by sweeping real frequency. The main figure is for the silver MIaM. The inset is for gold MIaM. The horizontal axis stands for the propagation constant $\beta $, which is normalized by ${k_\textrm{p}} = {{{\omega _\textrm{p}}} / c}$. Here the plasma frequency for silver (gold) is set to ${\omega _\textrm{p}}$ = 9 eV (8.9 eV) and c is the speed of light in vacuum. The vertical axis is the frequency denoted by its corresponding photonic energy $\hbar \omega $. The small discrepancy of the two cases could be observed from the starting point of the upper branch of the dispersions, where the starting point for gold MIaM is a bit lower than the silver case. The coalescence point of the two branches is shifted left a bit for the gold case. Neither case exhibits EPs in the real frequency spectrum.

Download Full Size | PDF

Therefore, based on Eq. (11), we alternatively swept real wavenumber (propagation constant) $\beta$ to search for EPs. For simplicity, the horizontal and vertical coordinates of EPs are written as (${\beta _{\textrm{EP}}}$, ${\omega _{\textrm{EP}}}$). In Fig. 3, $\beta $ is swept from 0 to 15kp for silver MIaM systems. Here, ${k_\textrm{p}} = {{{\omega _\textrm{p}}} / c}$, ${\omega _\textrm{p}}$ is the plasma frequency and c is the speed of light in vacuum. Take the vacuum gap 2d = 10 nm in Fig. 3(a) for example. When the two blue branches coalesce at $\beta = 9.12{k_\textrm{p}}$, $\omega $ switches from real values into complex conjugates, which indicate that the point ${\beta _{\textrm{EP}}} = 9.12$ ${k_\textrm{p}}$ is the EP of the MIaM system. When $\beta > \,{\beta _{\textrm{EP}}}$, the complex $\tilde{\omega }$ (from now on, the tilde represents complex values) indicates that the supported SPP modes are with time-dependent attenuation or amplification. They are the QNMs of the PT symmetry broken phase. Varying the gap size 2d (10 nm, 20 nm, 100 nm) reveals that ${\beta _{\textrm{EP}}}$ is inversely proportional to the gap size. This can be qualitatively explained as: the PT symmetry is easier to be broken down as the coupling strength is weakened when increasing the gap size. At the EP, ${\omega _{\textrm{EP}}} \approx {{{\omega _\textrm{p}}} / {\sqrt 2 }}$ hold for all the three different cases, which is the characteristic surface plasmon frequency ${\omega _{\textrm{SP}}}$ of the planar interface and is thus independent of the gap size. For all the three cases with different gap sizes in Fig. 3(a), the imaginary parts of $\tilde{\omega }$ ‘s of the QNMs approach the same asymptotic value $\gamma /2$, ($\gamma $ is the damping factor of the Drude model of silver). Therefore, for all the QNMs in the symmetry broken phase, $\tilde{\omega } \to {\omega _{\textrm{SP}}} \pm \textrm{i}{\gamma / 2}$ when $\beta \to \infty $. In Fig. 3(b), we investigate the dependence of the QNMs on the damping (amplifying) factor when the gap size is fixed. It can be observed that larger damping (amplifying) induces smaller ${\beta _{\textrm{EP}}}$. This is also an intuitive conclusion because the fixed coupling strength would not be able to balance too large attenuation and amplification. When the electric field of the surface modes becomes sufficiently large, the nonlinearity of metals would emerge and lead to second harmonic generations (SHG) [41,42]. However, it should also be importantly noted that the above description is based on the ideal Drude model. Further considerations should be taken because real metallic materials would not offer a constant amplification. Saturation effect [43] of the amplification factor would emerge as the amplitude of the modes increases. This would confine the whole system within finite amount of energy so that no singularity exists.

 figure: Fig. 3.

Fig. 3. (a) Dispersion curves of the silver MIaM of different gap sizes. Blue (black, green) solid line and red (pink, orange) dash dot line represent the real and imaginary part of complex frequency of the QNMs with the gap size set to 2d = 10 (20, 100) nm. The horizontal axis denotes the propagation constant $\beta $ normalized to ${k_\textrm{P}}$ = ${{{\omega _\textrm{P}}} / c}$. The vertical axis denotes the frequency with its corresponding photon energy in unit eV. (b) Dispersion curves of the silver MIaM of different Drude damping factors.

Download Full Size | PDF

4. Conclusion

In conclusion, starting from the complex coordinate transformation we have investigated the existence of EPs and QNMs in a dispersive MIaM system with PT symmetry. It has been demonstrated that the EPs can’t be found by sweeping the real frequency spectrum. By extending the frequency to the complex regime, the EP and the QNMs in the symmetry broken phase emerge. The MIaM is a very typical example which indicates the importance of considering the causality principle before searching for EPs in a realistic dispersive PT symmetric system. On the other side, owing to the rapid development of the research on PT symmetry in photonic systems, QNMs have been regarded as the input signal to realize the so-called virtual PT symmetry [36]. This is a rather new application of the modes in the PT symmetry broken phase. When the transmission line circuit with uniform decaying factor ${\gamma _0}$ is input by a time-decaying signal with a complex frequency $\tilde{\omega } = \omega - \textrm{i}{\gamma / 2}$, viz the QNM in this paper, the circuit (transmission line TL) renders the spatially amplifying time-average power flow when ${\gamma / 2} > {\gamma _0}$. It can be expected that the time-decaying (amplifying) characteristics of QNMs would be utilized to achieve further exotic phenomena in future.

Funding

Natural Science Foundation of Jiangsu Province (BK20190325); National Natural Science Foundation of China (61971134).

Acknowledgments

The author thanks the National Science Foundation of Jiangsu Province and National Natural Science Foundation of China for funding support.

Disclosures

The authors declare no competing interests.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. C. M. Bender and S. Boettcher, “Real spectra in non-hermitian hamiltonians having PT symmetry,” Phys. Rev. Lett. 80(24), 5243–5246 (1998). [CrossRef]  

2. C. M. Bender, S. Böttcher, and P. N. Meisinger, “PT-symmetric quantum mechanics,” J. Math. Phys. 40(5), 2201–2229 (1999). [CrossRef]  

3. G. Levai and M. Znojil, “Systematic search for PT-symmetric potentials with real spectra,” J. Phys. A 33(40), 7165–7180 (2000). [CrossRef]  

4. M. V. Berry, “Physics of non-Hermitian degeneracies,” Czech. J. Phys. 54(10), 1039–1047 (2004). [CrossRef]  

5. A. Mostafazadeh, “PT-symmetric Quantum Mechanics: A Precise and Consistent Formulation,” Czech. J. Phys. 54(10), 1125–1132 (2004). [CrossRef]  

6. W. Heiss, “Exceptional Points – Their Universal Occurrence and Their Physical Significance,” Czech. J. Phys. 54(10), 1091–1099 (2004). [CrossRef]  

7. C. M. Bender, “Making sense of non-Hermitian Hamiltonians,” Rep. Prog. Phys. 70(6), 947–1018 (2007). [CrossRef]  

8. R. El-Ganainy, K. G. Makris, D. N. Christodoulides, and Z. H. Musslimani, “Theory of coupled optical PT-symmetric structures,” Opt. Lett. 32(17), 2632–2634 (2007). [CrossRef]  

9. K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and Z. H. Musslimani, “Beam dynamics in PT symmetric optical lattices,” Phys. Rev. Lett. 100(10), 103904 (2008). [CrossRef]  

10. S. Klaiman, U. Günther, and N. Moiseyev, “Visualization of branch points in PT-symmetric waveguides,” Phys. Rev. Lett. 101(8), 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(9), 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(3), 192–195 (2010). [CrossRef]  

13. M. Brandstetter, M. Liertzer, C. Deutsch, P. Klang, J. Schöberl, H. E. Türeci, G. Strasser, K. Unterrainer, and S. Rotter, “Reversing the pump dependence of a laser at an exceptional point,” Nat. Commun. 5(1), 4034 (2014). [CrossRef]  

14. B. Peng, ŞK Özdemir, 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(5), 394–398 (2014). [CrossRef]  

15. H. Hodaei, M.-A. Miri, M. Heinrich, D. N. Christodoulides, and M. Khajavikhan, “Parity-time-symmetric microring lasers,” Science 346(6212), 975–978 (2014). [CrossRef]  

16. W. Li, C. Li, and H. Song, “Theoretical realization and application of parity-time-symmetric oscillators in a quantum regime,” Phys. Rev. A 95(2), 023827 (2017). [CrossRef]  

17. H. Alaeian and J. A. Dionne, “Controlling electric, magnetic, and chiral dipolar emission with PT-symmetric potentials,” Phys. Rev. B 91(24), 245108 (2015). [CrossRef]  

18. H. Alaeian and J. A. Dionne, “Parity-time-symmetric plasmonic metamaterials,” Phys. Rev. A 89(3), 033829 (2014). [CrossRef]  

19. H. Alaeian and J. A. Dionne, “Non-Hermitian nanophotonic and plasmonic waveguides,” Phys. Rev. B 89(7), 075136 (2014). [CrossRef]  

20. J. B. Pendry, D. Schurig, and D. R. Smith, “Controlling electromagnetic fields,” Science 312(5781), 1780–1782 (2006). [CrossRef]  

21. U. Leonhardt, “Optical conformal mapping,” Science 312(5781), 1777–1780 (2006). [CrossRef]  

22. H. Chen, C. Chan, and P. Sheng, “Transformation optics and metamaterials,” Nat. Mater. 9(5), 387–396 (2010). [CrossRef]  

23. J. B. Pendry, A. I. Fernández-Domínguez, Y. Luo, and R. Zhao, “Capturing photons with transformation optics,” Nat. Phys. 9(8), 518–522 (2013). [CrossRef]  

24. D. Genov, S. Zhang, and X. Zhang, “Mimicking celestial mechanics in metamaterials,” Nat. Phys. 5(9), 687–692 (2009). [CrossRef]  

25. H. Chen, S. Tao, J. Bělín, J. Courtial, and R.-X. Miao, “Transformation cosmology,” Phys. Rev. A 102(2), 023528 (2020). [CrossRef]  

26. J. Zhang, M. Wubs, P. Ginzburg, G. Wurtz, and A. V. Zayats, “Transformation quantum optics: designing spontaneous emission using coordinate transformations,” J. Opt. 18(4), 044029 (2016). [CrossRef]  

27. J. Pendry, P. A. Huidobro, Y. Luo, and E. Galiffi, “Compacted dimensions and singular plasmonic surfaces,” Science 358(6365), 915–917 (2017). [CrossRef]  

28. M. Kraft, A. Braun, Y. Luo, S. A. Maier, and J. B. Pendry, “Transformation Optics: A Time- and Frequency-Domain Analysis of Electron-Energy Loss Spectroscopy,” Nano Lett. 16(8), 5156–5162 (2016). [CrossRef]  

29. S. Savoia, G. Castaldi, and V. Galdi, “Complex-coordinate non-Hermitian transformation optics,” J. Opt. 18(4), 044027 (2016). [CrossRef]  

30. G. Castaldi, S. Savoia, V. Galdi, A. Alù, and N. Engheta, “PT Metamaterials via Complex-Coordinate Transformation Optics,” Phys. Rev. Lett. 110(17), 173901 (2013). [CrossRef]  

31. L. Luo, J. Luo, H. Chu, and Y. Lai, “Pseudo-Hermitian Systems Constructed by Transformation Optics with Robustly Balanced Loss and Gain,” Adv Photo Res 2, 2000081 (2021). [CrossRef]  

32. E. S. C. Ching, P. T. Leung, A. Maassen van den Brink, W. M. Suen, S. S. Tong, and K. Young, “Quasinormal-mode expansion for waves in open systems,” Rev. Mod. Phys. 70(4), 1545–1554 (1998). [CrossRef]  

33. J. Luis Jaramillo, R. Panosso Macedo, and L. Al Sheikh, “Pseudospectrum and Black Hole Quasinormal Mode Instability,” Phys. Rev. X 11(3), 031003 (2021). [CrossRef]  

34. S. Franke, J. Ren, M. Richter, A. Knorr, and S. Hughes, “Fermi’s Golden Rule for Spontaneous Emission in Absorptive and Amplifying Media,” Phys. Rev. Lett. 127(1), 013602 (2021). [CrossRef]  

35. J. Ren, S. Franke, and S. Hughes, “Quasinormal Modes, Local Density of States, and Classical Purcell Factors for Coupled Loss-Gain Resonators,” Phys. Rev. X 11(4), 041020 (2021). [CrossRef]  

36. H. Li, A. Mekawy, A. Krasnok, and A. Alù, “Virtual Parity-Time Symmetry,” Phys. Rev. Lett. 124(19), 193901 (2020). [CrossRef]  

37. E. N. Economou, “Surface plasmons in thin films,” Phys. Rev. 182(2), 539–554 (1969). [CrossRef]  

38. A. A. Zyablovsky, A. P. Vinogradov, A. A. Pukhov, A. V. Dorofeenko, and A. A. Lisyansky, “PT-symmetry in optics,” Phys.-Usp. 57(11), 1063–1082 (2014). [CrossRef]  

39. A. A. Zyablovsky, A. P. Vinogradov, A. V. Dorofeenko, and A. A. Pukhov, “Causality and phase transitions in PT -symmetric optical systems,” Phys. Rev. A 89(3), 033808 (2014). [CrossRef]  

40. 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(25), 253601 (2014). [CrossRef]  

41. K. N. Reddy, P. Y. Chen, A. I. Fernández-Domínguez, and Y. Sivan, “Surface second-harmonic generation from metallic-nanoparticle configurations: A transformation-optics approach,” Phys. Rev. B 99(23), 235429 (2019). [CrossRef]  

42. J. Rudnick and E. A. Stern, “Second-harmonic radiation from metal surfaces,” Phys. Rev. B 4(12), 4274–4290 (1971). [CrossRef]  

43. S. Assawaworrarit, X. Yu, and S. Fan, “Robust wireless power transfer using a nonlinear parity–time-symmetric circuit,” Nature 546(7658), 387–390 (2017). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (3)

Fig. 1.
Fig. 1. Schematic illustration of the TO framework. Vacuum is the auxiliary space before coordinate transformation. After performing ${r^{\prime}} = {F}({r})$, the system is transformed to the M-I-aM structure, the transformed physical space exhibits PT symmetry.
Fig. 2.
Fig. 2. The dispersion curves of the MIaM systems calculated by sweeping real frequency. The main figure is for the silver MIaM. The inset is for gold MIaM. The horizontal axis stands for the propagation constant $\beta $, which is normalized by ${k_\textrm{p}} = {{{\omega _\textrm{p}}} / c}$. Here the plasma frequency for silver (gold) is set to ${\omega _\textrm{p}}$ = 9 eV (8.9 eV) and c is the speed of light in vacuum. The vertical axis is the frequency denoted by its corresponding photonic energy $\hbar \omega $. The small discrepancy of the two cases could be observed from the starting point of the upper branch of the dispersions, where the starting point for gold MIaM is a bit lower than the silver case. The coalescence point of the two branches is shifted left a bit for the gold case. Neither case exhibits EPs in the real frequency spectrum.
Fig. 3.
Fig. 3. (a) Dispersion curves of the silver MIaM of different gap sizes. Blue (black, green) solid line and red (pink, orange) dash dot line represent the real and imaginary part of complex frequency of the QNMs with the gap size set to 2d = 10 (20, 100) nm. The horizontal axis denotes the propagation constant $\beta $ normalized to ${k_\textrm{P}}$ = ${{{\omega _\textrm{P}}} / c}$. The vertical axis denotes the frequency with its corresponding photon energy in unit eV. (b) Dispersion curves of the silver MIaM of different Drude damping factors.

Equations (14)

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

x = x p ( y , z ) , y = y q ( x , z ) , z = z s ( x , y ) , | x | d ,
x = x , y = y , z = z , | x | < d .
Λ = ( x , y , z ) ( x , y , z ) = ( p x p y x p z y q x q y q z z s x z s y s )
ϵ ( r ) = μ ( r ) = | Λ | Λ 1 Λ T , | x | d ,
ε ( r ) = μ ( r ) = 1 , | x | < d .
p = { p 0 x < d p 0 x > d , q = { q 0 x < d q 0 x > d , s = { s 0 x < d s 0 x > d
Λ = diag [ p , q , s ] .
ϵ ( r ) = μ ( r ) = diag [ q s / p , p s / q , p s / q ] .
ε x x = q s / p , ε z z = p q / s , μ y y = p s / q .
For | x | d , μ y y = 1 , ε x x = ε z z = { η x d η x d
For | x | d , ε = μ = 1.
η = 1 ω p 2 ω 2 + i γ ω .
e 4 κ 1 d = κ 1 / ε 1 + κ 2 / ε 2 κ 1 / ε 1 κ 2 / ε 2 κ 1 / ε 1 + κ 3 / ε 3 κ 1 / ε 1 κ 3 / ε 3 .
e 4 β d = ε 2 + ε 1 ε 2 ε 1 ε 3 + ε 1 ε 3 ε 1 .
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.