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

Disorder induced rotational-symmetry breaking to control directionality of whispering gallery modes in circularly symmetric nanowire assembly

Open Access Open Access

Abstract

Effect of disorder on the emission directionality of a whispering gallery mode resonator comprising of circularly symmetric nanowire array is investigated using finite-difference time-domain analysis technique. In spite of rotational symmetry breaking, whispering gallery mode-like isotropic emission characteristics are retained by the nanowire array up to a certain degree of spatial disorder. For higher degrees of randomness, Anderson localized resonant modes are obtained with unidirectional emission characteristics, though the beam direction remains unpredictable because of the underlying random-scattering process. This shortcoming is overcome by designing a system of correlated disorder where nanowire spacing is varied gradually along the preferred axis of unidirectionality. This system, in spite of its high degree of disorder, can effectively support high-Q whispering gallery modes with tunable unidirectional emission characteristics.

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

1. Introduction

Whispering gallery mode resonators offer in themselves a unique test bed for exploring fundamental physics, as well for designing exotic applications related to the field of photonics and cavity quantum electrodynamics. Total internal reflections resulting from rotational symmetry of whispering gallery mode (WGM) resonators can result in ultrahigh quality factors in the order of 106108, placing them in equal footing with high-quality photonic crystals as far as optical confinement is concerned [1, 2]. The suitability of WGM resonators have already been studied extensively for sensing [3, 4], ultra-low threshold lasing [2, 5, 6], quantum information processing, probing, imaging and filtering related applications [7–9]. However, unfortunately, the characteristic isotropic emission pattern of WGMs poses a long-standing challenge in designing applications where substantial directional emission is required [2, 5, 6, 10–12].

Previous studies related to the directionality of WGMs have primarily focused on breaking the rotational symmetry of the resonator, employing schemes such as deformation of the cavity shape [10–12], incorporation of scatterer inside the cavity [13], or embedding of scatterers at the periphery of the cavity [14]. These studies have been based on conventional micro-ring, micro-disk, or micro-sphere like WGM resonators. Though there have been a report of directionality enhancement in triangular-shaped deformed cavity comprising of micro-posts [15], the case of circularly symmetric nanowire assembly has remained unexplored in spite of the significant potential of such systems for realizing tunable, high-density, defect-free optoelectronic and photonic systems [16, 17].

Disordered systems designed with non-Hermitian potentials offer the prospect of defining specific behaviors of light, which cannot be otherwise realized in a homogenous environment [18, 19]. Based on this principle, disordered photonic systems have been extensively studied for numerous exotic applications, such as for attaining unusual localization of light [20], and also for enhancing the collimation bandwidth of light [21], to name a few. In the present study, we study the role of disorder on directionality enhancement in a circular nanowire array based on finite-difference time-domain analysis technique. Quality factor in the order of 105 is obtained along with quasi-isotropic emission characteristics for the radially symmetric case. However with increasing disorder, in-plane directionality tends to increase and eventually unidirectional emission characteristics are obtained in the Anderson localized regime. By identifying cross-over between the rotational symmetric and multiple scattering mediated regimes, an optimized design of correlated disorder has been proposed such that directionality can be enhanced without substantial degradation of the Q-factor.

 figure: Fig. 1

Fig. 1 (a) Schematic representation of the rotationally symmetric GaN nanowire array (inset shows the hexagonal unit cell); (b) calculated Q-factors of the resonant modes with corresponding mode numbers (inset shows enlarged view of Q-factor vs. resonant wavelength values of the weakly confined modes existing between 400 nm – 700 nm); (c) near-field distribution of E-field of the WGM; (d) far-field emission characteristics exhibiting directionality characteristics of the corresponding WGM.

Download Full Size | PDF

2. Radially symmetric nanowire array

The nanowire assembly considered in this work comprises of homogenous GaN nanowires having air medium in between. Experimental growth and fabrication of high quality GaN-nanowire arrays in both patterned and self-organized manner have been reported extensively employing electron-beam lithography, molecular beam epitaxy, focused ion beam milling and chemical vapor deposition techniques [16, 17, 22, 23]. For investigating whispering gallery modes, a uniform distribution of GaN nanowires having experimentally reported diameter of 70 nm is considered [16, 17]. These nanowires are considered to be vertically oriented along z-axis such that they form a radially symmetric pattern along the x-y plane, which is schematically illustrated in Fig. 1(a). Aclose examination of this pattern indicates the presence of a hexagonal unit cell, which is shown as an inset to Fig. 1(a). In this unit cell, the distance between adjacent nanowires along x- and y- directions are defined as hX and hY respectively. For the radial symmetric case, hX and hY are taken to be 85 nm, and the nanowires are oriented such that the overall diameter of the ensemble forming the cavity becomes 3 μm. This overall diameter has been kept constant throughout the study.

Resonant characteristics of the nanowire array have been studied using open-source finite-difference time-domain (FDTD) software package MEEP [24]. A Gaussian pulse-source is placed at the center of the array and the resonant modes are identified by decomposing the electric (E) and magnetic (H) field components of the electromagnetic wave using filter diagonalization method [25]. Because transverse electric (TE) modes are expected to be rather weakly confined in GaN-nanowire based micro-cavities [6], only the cases of transverse magnetic (TM) modes have been considered in this work. Q-factors corresponding to these modes are plotted as a function of wavelength in Fig. 1(b). Near-field distribution of the mode having the highest Q-factor of2.2× 105 is shown in Fig. 1(c), which clearly indicates the presence of whispering gallery like characteristics with a standing wave formed along the cavity boundary and a null of energy at the center. The radial and azimuthal mode numbers obtained for this mode are r =1 and m =15 respectively. Similar WGM characteristics are observed for the strongly confined modes located at 914.7 nm and 793.2 nm, and also for the relatively weakly confined modes residing within the wavelength range of 400nm – 700 nm. Radial and azimuthal mode numbers obtained from corresponding near-field distributions of the modes are indicated as (r, m) in Fig. 1(b). Directionality characteristics of these resonant modes are obtained from their corresponding far-field distributions. Far-field radiation pattern obtained for the most strongly confined mode of (1, 15) is shown in Fig. 1(d), which clearly indicates the presence of quasi-isotropic emission characteristics. Similar quasi-isotropic directionality profiles are also obtained for the remaining whispering gallery modes of this array, thereby confirming the presence of perfect rotational symmetry in the structure.

 figure: Fig. 2

Fig. 2 (a) Nanowire arrays generated with percentage randomness values of 15%, 50%, and 100% respectively; (b) Q-factors of the most strongly confined modes plotted as a function of disorder strength and percentage randomness; (c) corresponding resonant wavelengths plotted as a function of disorder strength and percentage randomness; inset (i) shows disorder strength plotted as a function of percentage randomness and inset (ii) shows the variation of l/lc with percentage randomness in regards to the satisfaction of Ioffe-Regel criterion.

Download Full Size | PDF

3. Directionality and resonant characteristics in disordered array

In order to disturb rotational symmetry of the array, spatial locations of the nanowires shown in Fig. 1(a) are systematically altered by a percentage randomness parameter, P [0, 100]. For a given value of P, arbitrary co-ordinates (x+δx, y+δy) are generated such that δx = (P/100)×(hX/2)×rx and δy = (P/100)×(hY/2)×ry, where {rx,ry} [-1, 1] are random numbers, and (x, y) represent center positions of the nanowires shown in the symmetric array of Fig. 1(a). Three disordered systems derived from this symmetric structure are schematically shown in Fig. 2(a), where percentage randomness values of 15%, 50% and 100% have been considered. To compare disorder between arrays having different diameters and fill factors of the nanowires, a disorder strength parameter ζ, defined as the correlation between refractive index profiles of the symmetric and disordered arrays, is also utilized. In order to calculate ζ, the two-dimensional refractive index profiles having N elements of the periodic and disordered structures are first transformed into one-dimensional vectors A and B respectively. In this transformation, consecutive rows of a two-dimensional refractive index profile of an array have been cascaded to form the corresponding one-dimensional vector. The mean and standard deviation of the vector A (B) being μA (μB) and σA (σB) respectively, the disorder strength is calculated as [26]

ζ=11N1j=1N(AjμAσA)(BjμBσB)

Using Eq. (1), ζ values of 0.105, 0.523 and 0.95 are obtained respectively for the patterns shown in Fig. 2(a). Further calculations for patterns obtained for different values of P indicate a positive correlation between P and ζ such that ζ increases from 0 to 0.95 as P is increased from 0 to 100%.

 figure: Fig. 3

Fig. 3 (a) Electric-field distributions of WGM, quasi-WGM and Anderson localized modes obtained for percentage randomness values of 5%, 20% and 50% respectively; (b) corresponding far-field distribution profiles illustrating directionality characteristics of these modes; (c) beam divergence angle plotted as a function of both disorder strength and percentage randomness; the error bars indicate the variation of FWHM obtained for random configurations having identical values of percentage randomness and nanowire diameter.

Download Full Size | PDF

To investigate the effect of disorder on resonant conditions of the array, Q-factor of the most strongly confined mode and its corresponding wavelength are next calculated and plotted as a function of both P and ζ in Figs. 2(b) and 2(c). In order to avoid the artifact of a random sample, configurational averaging on several realizations of a disordered array having same parameters has been performed. The relation between disorder strength and percentage randomness for these configurational averaged arrays is shown as an inset to Fig. 2(c), where ζ is plotted as a function of P. The range of variation of ζ for a fixed value of P, nanowire spacing and diameter, is indicated as an error bar in this plot. Higher values of P results in greater variation of ζ, thereby indicating the presence of enhanced spatial disorder. As can be observed in Fig. 2(b), the loss of rotational symmetry with increasing disorder results in degradation of the Q-factor. However, it is noteworthy that for a wide range of the disorder strength, the Q-factor remains above 104. Moreover, resonant wavelength of the most tightly confined mode shifts from near-infrared to ultra-violet regime of the EM spectra as disorder is increased. To explain this behavior, near-field pattern of the E-field for P = 5%, 20% and 50% have been examined. As can be seen in Fig. 3(a), though WGM persists for the case of 5% randomness, Q-factor decreases by about one order of magnitude because of the reduced symmetry of the structure. For P = 20%, the peak resonant characteristics changes from being WGM to quasi-WGM, an operation regime also known as the triangular WGM [5, 6]. This regime in fact represents the co-existence of regular and chaotic wave dynamics- an attribute which has been previously described in the framework of Kolmogorov-Arnold-Moser (KAM) theorem [27] and dynamical tunneling [28]. Such regular-to-chaotic wave dynamics have also been exploited to experimentally realize momentum transformation in optical resonators [29], and also to suppress spatiotemporal instabilities in microcavity lasers [30]. The qausi-WGM mode in our case is rather weakly confined compared to the case of WGM and the corresponding Q-factors monotonically decrease to about 7× 103. As disorder is further increased, wave dynamics inside the cavity shifts from multiple total internal reflection mediated whispering-gallery regime to multiple random scattering mediated strong-localization regime. Such strong-localization of light resulting from interference of multiply scattered wave in a disordered medium is commonly referred to as Anderson localization (AL) effect [17, 31–33]. Transition to Anderson localized regime has been further verified based on the Ioffe-Regel criterion. According to this criterion, Anderson localization of wave in the disordered medium is governed by the condition l/lc1, where l and lc denote the scattering length and critical scattering length respectively. In our study, the values of l/lc for each isotropic disordered structure have been calculated using the theoretical formulation reported in [33]. As has been shown as an inset to Fig. 2(c), the value of l/lc decreases with the increase of percentage randomness, and eventually becomes smaller than 1 when P > 30%. This confirms that Ioffe-Regel criteria is indeed satisfied for such values of percentage randomness and the system consequently operates in the Anderson localized regime. Quality factor of the Anderson localized mode is observed to increase with percentage randomness because of enhanced multiple scattering events in the array. For the considered nanowire diameter of 70 nm, localization wavelength resides within the range of 300nm − 330 nm, which is consistent with our previous findings [33]. Three operation regimes, namely WGM, quasi-WGM and AL regime, have been identified and indicated in Figs. 2(b) and 2(c) based on the observed transitions of wave dynamics as the degree of randomness is gradually changed.

 figure: Fig. 4

Fig. 4 (a) Schematic of the correlated disordered nanowire array having spatially varying density of nanowires; (b) Q-factors of the resonant modes plotted as a function of wavelength; (c) normalized E-field distribution of the highest Q-factor WGM; (d) far-field directionality profiles of the WGM for the nanowire array and also for the equivalent effective medium structure; (e) equivalent effective medium structure for the array shown in Fig. 4(a).

Download Full Size | PDF

In order to investigate the emission directionality in these three regimes, far-field distribution of the E-fields are next calculated for arrays of different disorder strengths. Representative far-field patterns for these regimes are shown in Fig. 3(b). As can be observed, in spite of the loss of rotational symmetry, the emission profile is nearly isotropic for 5% randomness, and for the quasi-WGM case, the emission collimates as several narrow beams in different directions. In the AL regime however, complete breakdown of rotational symmetry results in a highly directional emission with a narrow beam angle. To further distinguish directionality characteristics of the three operation regimes, beam divergence angles are calculated as full-width at half maxima (FWHM) of the corresponding far-field distributions. In Fig. 3(c), beam divergence angles plotted as a function of percentage randomness indicate gradual improvement of beam directionality with increasing disorder. A relatively sharp decrease in FWHM, accompanied by a concurrent blue-shift of the resonant wavelength of the highest Q-factor mode, is observed in the quasi-WGM regime. For higher degrees of randomness, in the AL regime, the FWHM value resides within 15   50  and nearly unidirectional emission characteristics are obtained. The underlying random scattering events in this regime however dictate that the direction of emission is predominantly governed by the local randomness of the specific disordered pattern, and therefore the beam-direction remains random for all values of the disorder strength. The error bars shown in Fig. 3(c) indicate the range of variation of FWHM obtained for random configurations of nanowire arrays having identical values of percentage randomness and nanowire diameter. As can be observed, the FWHM varies over a wider range in the quasi-WGM regime and tends to decrease as the degree of randomness increases. To have an estimate of the fraction of power radiated along a specific direction, output power density at each angle has also been calculated for each resonances and is normalized by incident power density. For the different arrays considered in this study, the maximum fraction of power radiated along the direction of maximum intensity is observed to range from 0.65 − 0.82.

Tables Icon

Table 1. Comparison of FDTD Analysis Results with the Effective Medium Theory Based TIR Relation

In order to achieve improved directionality while retaining control over the beam-direction, a correlated disordered structure is next proposed and analyzed such that it can be conveniently fabricated employing experimental techniques such as focused ion-beam milling, electron beam lithography, or patterned epitaxy. This structure, which is illustrated in Fig. 4(a), is designed by gradually reducing hX from 135 nm to 85 nm, while keeping hY constant. The geometry, albeit rotationally asymmetric, is symmetric with respect to x-axis. Resonant characteristics obtained for this array is shown in Fig. 4(b) along with corresponding radial and azimuthal mode numbers. A maximum Q-factor of 3853 is obtained at 719.5 nm for the (1, 20) WGM, the near-field distribution of which is shown in Fig. 4(c). Unidirectional emission characteristics of this mode is confirmed by the normalized far-field pattern (Fig. 4(d)), which exhibits a directional emission along the axis of symmetry (x-axis). A beam divergence angle of 140  is obtained in this case. It may be noted that in order to have an estimate of the minimum FWHM attainable with the graded-index-like structure, directionality characteristics of several such structures were also explored by gradually varying the value of hX. However the lower limit of FWHM for such structures appeared to be 140 , which is the case already illustrated in Fig. 4(d). It is envisaged that there is scope of further tuning the value of FWHM for graded-index-like structures by adopting other means, such as by non-linearly reducing the value of hX. However such a detailed study was beyond the scope of the present work. An important aspect of the graded-index-like structure is that the emission direction for this structure remains controllable by all means as it is governed by the axis along which disorder is introduced by gradually changing the nanowire spacing. The maximum fraction of radiated power along this controlled direction is calculated to be 0.75. This value is well within the range of previously obtained fraction of power radiated along maximum directionalities for arrays having different disorder strengths.

Furthermore, the correlated disordered array offers in itself a system for which resonant wavelengths can be predicted in spite of the inherent disorder of the structure. To justify this claim, ray-optics based multiple total internal reflection (TIR) relation, κL = 2πm is considered, where κ is the wavenumber (κ = ηEM×(2π/λ)), L is the optical path length, m is the azimuthal mode number, ηEM is the effective refractive index, and λ is the resonant wavelength corresponding to the WGM. Here effective refractive index of the composite medium is calculated employing the following Maxwell-Garnett model [34]

ηEM2=ηGaN22(1fair)ηGaN2+(1+2fair)ηair2(2+fair)ηGaN2+(1fair)ηair2
where ηGaN and ηair are refractive indices of GaN and air respectively, fair is the air filling factor and fGaN = 1fair is the GaN nanowire filling factor. Effective radius of optical path (Rr) for different radial modes is calculated from the corresponding near-field distributions of the electric field. Finally resonant wavelength for a given whispering gallery mode is calculated using the relation λ(r,m)=(2πηEMRr)/m. As shown in Table 1, a good agreement between theoretical calculations and results obtained from FDTD analysis is obtained for both periodic and disordered arrays.

To compare directionality profile of the effective medium with that of the actual structure shown in Fig. 4(a), its equivalent composite medium is also estimated based on Eq. (2). As indicated by the grayscale bar in Fig. 4(e), refractive index of the equivalent medium gradually varies from 1.623 to 1.839 along the x-axis. The far-field distribution of this medium is shown in Fig. 4(d), which is in good agreement with the directionality profile obtained for the actual nanowire array. These results indicate that in spite of rotational symmetry breaking, resonant modes, as well as directionality characteristics of a disordered structure having gradually varying refractive index can be estimated from its underlying whispering gallery mode-like wave dynamics. Such a design holds promise for realizing nanowire-based photonic and optoelectronic systems where spatial orientation of the nanowires can be tailored to attain specific near- and far-field characteristics of the whispering gallery modes, therefore making them suitable for exotic applications related to lasing, sensing, probing and quantum information processing.

4. Conclusion

In summary, whispering gallery modes in GaN nanowire based circular cavity have been investigated with particular focus on the role of disorder on its directionality properties. FDTD based analysis in the near- and far-field regime shows that notwithstanding the presence of disorder, a circular array of nanowires retains its WGM-like isotropic emission behavior up to a certain degree of spatial randomness. As the disorder is increased, high-Q Anderson localized resonant modes are obtained with unidirectional light output, though the beam direction remains random because of the underlying random scattering process. This limitation can be overcome by designing a system of correlated disorder, where the gradual change of refractive index predefines the beam direction, as well as manifests in itself predictable WGM-like characteristics.

Acknowledgments

The authors thankfully acknowledges the support and facilities obtained from the Institute of Information and Communication Technology (IICT) and the Department of Electrical and Electronic Engineering (EEE), Bangladesh University of Engineering and Technology (BUET) during the course of this research work.

References

1. S. Yang, Y. Wang, and H. Sun, “Advances and prospects for whispering gallery mode microcavities,” Adv. Opt. Mater. 3, 1136–1162 (2015). [CrossRef]  

2. X. F. Jiang, C. L. Zou, L. Wang, Q. Gong, and Y. F. Xiao, “Whispering-gallery microcavities with unidirectional laser emission,” Laser Photonics Rev. 10, 40–61 (2016). [CrossRef]  

3. X. Lopez-Yglesias, J. M. Gamba, and R. C. Flagan, “The physics of extreme sensitivity in whispering gallery mode optical biosensors,” J. Appl. Phys. 111, 84701 (2012). [CrossRef]  

4. H. Nadgaran and M. A. Garaei, “Enhancement of a whispering gallery mode microtoroid resonator by plasmonic triangular gold nanoprism for label-free biosensor applications,” J. Appl. Phys. 118, 43101 (2015). [CrossRef]  

5. C. Tessarek, R. Röder, T. Michalsky, S. Geburt, H. Franke, R. Schmidt-Grund, M. Heilmann, B. Hoffmann, C. Ronning, M. Grundmann, and S. Christiansen, “Improving the optical properties of self-catalyzed GaN microrods toward whispering gallerymode lasing,” ACS Photonics 1, 990–997 (2014). [CrossRef]  

6. H. Baek, J. K. Hyun, K. Chung, H. Oh, and G. C. Yi, “Selective excitation of Fabry-Perot or whispering-gallery mode-type lasing in GaN microrods,” Appl. Phys. Lett. 105, 201108 (2014).

7. D. V. Strekalov, C. Marquardt, A. B. Matsko, H. G. Schwefel, and G. Leuchs, “Nonlinear and quantum optics with whispering gallery resonators,” J. Opt. 18, 123002 (2016). [CrossRef]  

8. S. H. Huang, S. Sheth, E. Jain, X. Jiang, S. P. Zustiak, and L. Yang, “Whispering gallery mode resonator sensor for in situ measurements of hydrogel gelation,” Opt. Express 26, 51–62 (2018). [CrossRef]   [PubMed]  

9. V. S. Ilchenko and A. B. Matsko, “Optical resonators with whispering-gallery modes-part II: applications,” IEEE J. Sel. Top. Quantum Electron. 12, 15–32 (2006). [CrossRef]  

10. S. V. Boriskina, T. M. Benson, P. D. Sewell, and A. I. Nosich, “Directional emission, increased free spectral range, and mode Q-factors in 2-D wavelength-scale optical microcavity structures,” IEEE J. Sel. Top. Quantum Electron. 12, 1175–1182 (2006). [CrossRef]  

11. M. W. Kim, K. W. Park, C. H. Yi, and C. M. Kim, “Directional and low-divergence emission in a rounded half-moon shaped microcavity,” Appl. Phys. Lett. 98, 241110 (2011). [CrossRef]  

12. G. Y. Zhu, F. F. Qin, J. Y. Guo, C. X. Xu, and Y. J. Wang, “Unidirectional ultraviolet whispering gallery mode lasing from floating asymmetric circle GaN microdisk,” Appl. Phys. Lett. 111, 202103 (2017). [CrossRef]  

13. S. Preu, S. I. Schmid, F. Sedlmeir, J. Evers, and H. G. Schwefel, “Directional emission of dielectric disks with a finite scatterer in the THz regime,” Opt. Express 21, 16370–16380 (2013). [CrossRef]   [PubMed]  

14. Q. J. Wang, C. Yan, N. Yu, J. Unterhinninghofen, J. Wiersig, C. Pflügl, L. Diehl, T. Edamura, M. Yamanishi, H. Kan, and F. Capasso, “Whispering-gallery mode resonators for highly unidirectional laser action,” Proc. Natl. Acad. Sci. U.S.A. 107, 22407–22412 (2010). [CrossRef]   [PubMed]  

15. Y. Kim, S. Y. Lee, J. W. Ryu, I. Kim, J. H. Han, H. S. Tae, M. Choi, and B. Min, “Designing whispering gallery modes via transformation optics,” Nat. Photonics 10, 647 (2016). [CrossRef]  

16. W. Guo, M. Zhang, A. Banerjee, and P. Bhattacharya, “Catalyst-free InGaN/GaN nanowire light emitting diodes grown on (001) silicon by molecular beam epitaxy,” Nano Lett. 10, 3355–3359 (2010). [CrossRef]   [PubMed]  

17. K. H. Li, X. Liu, Q. Wang, S. Zhao, and Z. Mi, “Ultralow-threshold electrically injected AlGaN nanowire ultraviolet lasers on Si operating at low temperature,” Nat. Nanotechnol. 10, 140 (2015). [CrossRef]   [PubMed]  

18. S. Yu, X. Piao, and N. Park, “Bohmian photonics for independent control of the phase and amplitude of waves,” Phys. Rev. Lett. 120, 193902 (2018).

19. K. G. Makris, Z. H. Musslimani, D. N. Christodoulides, and S. Rotter, “Constant-intensity waves and their modulation instability in non-hermitian potentials,” Nat. Commun. 6, 7257 (2015). [CrossRef]   [PubMed]  

20. S. Yu, X. Piao, J. Hong, and N. Park, “Metadisorder for designer light in random systems,” Sci. Adv. 2, e1501851 (2016). [CrossRef]   [PubMed]  

21. P. Hsieh, C. Chung, J. F. McMillan, M. Tsai, M. Lu, N. C. Panoiu, and C. W. Wong, “Photon transport enhanced by transverse anderson localization in disordered superlattices,” Nat. Phys. 11, 268 (2015). [CrossRef]  

22. E. A. Serban, J. Palisaitis, P. O. A. Persson, L. Hultman, J. Birch, and C. L. Hsiao, “Site-controlled growth of GaN nanorod arrays by magnetron sputter epitaxy,” Thin Solid Films 660, 950–955 (2018). [CrossRef]  

23. S. D. S. D. Hersee, X. Sun, and X. Wang, “The controlled growth of GaN nanowires,” Nano Lett. 6, 1808–1811 (2016). [CrossRef]  

24. A. F. Oskooi, D. Roundy, M. Ibanescu, P. Bermel, J. D. Joannopoulos, and S. G. Johnson, “MEEP: a flexible free-software package for electromagnetic simulations by the FDTD method,” Comput. Phys. Commun. 181, 687–702 (2010). [CrossRef]  

25. V. A. Mandelshtam and H. S. Taylor, “Harmonic inversion of time signals and its applications,” J. Chem. Phys. 107, 6756–6769 (1997). [CrossRef]  

26. A. K. Boddeti, R. Kumar, and S. Mujumdar, “Figures-of-merit of Anderson localization cavities in membrane-based periodic-on-average random templates,” Opt. Commun. 397, 39–43 (2017). [CrossRef]  

27. H. Cao and J. Wiersig, “Dielectric microcavities: Model systems for wave chaos and non-hermitian physics,” Rev. Mod. Phys. 87, 61 (2015). [CrossRef]  

28. A. Bäcker, R. Ketzmerick, S. Löck, M. Robnik, G. Vidmar, R. Höhmann, U. Kuhl, and H.-J. Stöckmann, “Dynamical tunneling in mushroom billiards,” Phys. Rev. Lett. 100, 174103 (2008). [CrossRef]   [PubMed]  

29. X. Jiang, L. Shao, S.-X. Zhang, X. Yi, J. Wiersig, L. Wang, Q. Gong, M. Lončar, L. Yang, and Y.-F. Xiao, “Chaos-assisted broadband momentum transformation in optical microresonators,” Science 358, 344–347 (2017). [CrossRef]   [PubMed]  

30. S. Bittner, S. Guazzotti, Y. Zeng, X. Hu, H. Yılmaz, K. Kim, S. S. Oh, Q. J. Wang, O. Hess, and H. Cao, “Suppressing spatiotemporal lasing instabilities with wave-chaotic microcavities,” Science 361, 1225–1231 (2018). [CrossRef]   [PubMed]  

31. P. W. Anderson, “Absence of diffusion in certain random lattices,” Phys. Rev. 109, 1492 (1958). [CrossRef]  

32. J. Liu, P. D. Garcia, S. Ek, N. Gregersen, T. Suhr, M. Schubert, J. Mørk, S. Stobbe, and P. Lodahl, “Random nanolasing in the Anderson localized regime,” Nat. Nanotechnol. 9, 285 (2014). [CrossRef]   [PubMed]  

33. D. J. Paul, A. A. Mimi, A. Hazari, P. Bhattacharya, and M. Z. Baten, “Finite-difference time-domain analysis of the tunability of Anderson localization of light in self-organized GaN nanowire arrays,” J. Appl. Phys. 125, 43104 (2019). [CrossRef]  

34. R. H. Siddique, G. Gomard, and H. Hölscher, “The role of random nanostructures for the omnidirectional anti-reflection properties of the glasswing butterfly,” Nat. Commun. 6, 6909 (2015). [CrossRef]   [PubMed]  

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 (a) Schematic representation of the rotationally symmetric GaN nanowire array (inset shows the hexagonal unit cell); (b) calculated Q-factors of the resonant modes with corresponding mode numbers (inset shows enlarged view of Q-factor vs. resonant wavelength values of the weakly confined modes existing between 400 nm – 700 nm); (c) near-field distribution of E-field of the WGM; (d) far-field emission characteristics exhibiting directionality characteristics of the corresponding WGM.
Fig. 2
Fig. 2 (a) Nanowire arrays generated with percentage randomness values of 15%, 50%, and 100% respectively; (b) Q-factors of the most strongly confined modes plotted as a function of disorder strength and percentage randomness; (c) corresponding resonant wavelengths plotted as a function of disorder strength and percentage randomness; inset (i) shows disorder strength plotted as a function of percentage randomness and inset (ii) shows the variation of l/lc with percentage randomness in regards to the satisfaction of Ioffe-Regel criterion.
Fig. 3
Fig. 3 (a) Electric-field distributions of WGM, quasi-WGM and Anderson localized modes obtained for percentage randomness values of 5%, 20% and 50% respectively; (b) corresponding far-field distribution profiles illustrating directionality characteristics of these modes; (c) beam divergence angle plotted as a function of both disorder strength and percentage randomness; the error bars indicate the variation of FWHM obtained for random configurations having identical values of percentage randomness and nanowire diameter.
Fig. 4
Fig. 4 (a) Schematic of the correlated disordered nanowire array having spatially varying density of nanowires; (b) Q-factors of the resonant modes plotted as a function of wavelength; (c) normalized E-field distribution of the highest Q-factor WGM; (d) far-field directionality profiles of the WGM for the nanowire array and also for the equivalent effective medium structure; (e) equivalent effective medium structure for the array shown in Fig. 4(a).

Tables (1)

Tables Icon

Table 1 Comparison of FDTD Analysis Results with the Effective Medium Theory Based TIR Relation

Equations (2)

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

ζ = 1 1 N 1 j = 1 N ( A j μ A σ A ) ( B j μ B σ B )
η E M 2 = η G a N 2 2 ( 1 f a i r ) η G a N 2 + ( 1 + 2 f a i r ) η a i r 2 ( 2 + f a i r ) η G a N 2 + ( 1 f a i r ) η a i r 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.