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

Localized modes in one-dimensional symmetric Thue-Morse quasicrystals

Open Access Open Access

Abstract

We present occurrence of the strongly localized modes with high transmission in one dimensional symmetric Thue-Morse quasicrystals. This quasicrystal has some interesting properties, including (i) there are strongly localized modes in separated regions which are around odd semi-quarter-wave thickness of the system, (ii) both the frequency of localized mode and the thicknesses of the space layer to appear localized modes are variant for different generation orders of the system, and (iii) the sharpness of the resonant peaks in the transmission spectra increases as the generation order of the system increases.

© 2014 Optical Society of America

Localized mode in dielectric photonic crystals has received intensive interest in both fundamental physics and for device applications [14]. This property is widely used in telecommunications, fibers and lasers. A dielectric periodic bilayer with defect (PBD) is one of the fundamental structures, which consists of two parallel structures with bilayers that reflect light back and forth successively in the cavity to generate a cavity localized field [5, 6]. Moreover, the localized modes in the PBD appear in the frequency region of the band gap of the traditional periodic bilayer in the transmission spectra. As the incident field conforms to the resonant conditions in the PBD, the electric field is highly localized in the defect layer [7, 8]. A variety of optical characteristics result when the arrangements of the structure and the optical thickness of the defect layer are changed. For the PBD structure, there are three important properties that allow diverse designs for optical devices [9]. Firstly, localized mode appears in defect layers of the same thickness as the periodic number increases. Secondly, the electric fields are symmetrically distributed within the structure. Finally, the positions of localized modes in the PBD, which appear around the odd quarter-wave thicknesses, are the same as those in the distributed Bragg reflector (DBR) with different periodic numbers [10, 11].

Recently, considerable effort has been placed on quasperiodic system since the discovery of quasicrystals [1214]. Due to the particular arrangement of quasicrystals, the properties of photon propagation in the quasicrystal are different from those in periodic and disordered structures [1517]. Thus, the quasicrystals have received a lot of interest in the designing of optical materials and devices. Since devices with a quasicrystal system in one dimension are easily manufactured, this study examines a 1-D quasicrystal structure [18, 19]. For quasiperiodic systems, one of the well-known structures is the Thue-Morse sequence [20, 21]. The existence of strongly localized modes has been proposed by inserting a defect layer into the supperlattice structures [22]. This paper proposes a new structure: a one-dimensional symmetrical Thue-Morse supperlattice (STMS). The STMS is formed by inserting a space layer between two parallel structures that are constructed with the symmetry of a Thue-Morse sequence. Moreover, we propose the novel method to find the localized modes in the STMS. Therefore, it is necessary to determine whether there is strongly localized mode in a STMS. If there is localized mode in a STMS, whether the optical properties of STMS’s and PBD’s are different? Firstly, do the localized modes also exist around the quarter-wave thickness in a STMS that are similar to a PBD? Secondly, does the frequency of the resonant mode for a fixed thickness of the space layer in the transmission spectra of the STMS depend on the generation orders? Finally, if the optical thickness of the space layer is changed for a fixed frequency, do resonant modes appear periodically, which are similar to that in a PBD?

For the sake of simplicity, a one-dimensional Thue-Morse structure, which consists of two different materials, A and B, according to the iteration rule: A→AB and B→BA, is considered. The structure of the system with a lower order is given by ν≥2 with S2 = AB, S3 = ABBA, S4 = ABBABAAB, etc. The STMS is constructed by inserting a space layer (ds) into two parallel and symmetrical Thue-Morse dielectric supperlattice; for instance, S´2dsS2, 3dsS3, 4dsS4, etc. The arrangement of S´ is opposite to the arrangement of S. However, a PBD can be denoted as mdDPm, where Pm = (AB)m and P´m = (BA)m and m is the number of periods.

The relationship between the resonant peak in traditional major gaps in the transmission spectrum for the STMS with different generation orders and a PBD is considered firstly. The traditional gap regions are found in gap map of fundamental Thue-Morse structure [23]. The transmission spectra for the generation orders 5 and 6 in the STMS are shown in Fig. 1(a) and 1(b). The gray areas are the regions of the major gap in the band structure of the fundamental Thue-Morse supperlattice. The squared electric field of the light propagation corresponding to the signs A and B in Fig. 1(a) and 1(b) are shown in Fig. 1(c) and 1(d). We find that there is strong localization near the space layer. From Fig. 1, it is seen that there are two properties that are similar to the PBD. The first property is that the resonant modes appear on a conformed thickness of space layer and the corresponding frequencies appear in the traditional major gaps. The other property is that the minimum value for the transmittance decreases as the generation order increases.

 figure: Fig. 1

Fig. 1 The transmission spectra in the STMS’s with generation orders (a) ν = 5, and (b) ν = 6, for normal incidence from air. The gray areas correspond to the region of the major gap in the band structure of the Thue-Morse superlattice. The |E|2 distribution in the ν5 STMS for (c) Ω = 0.2308 and (d) Ω = 0.5192, respectively, which correspond to the signs, A and B, in Fig. 1(a). The parameters of the system are nA = 1, nB = 2, nD = 2, dA = 2/3µm, dB = 1/3µm, and dS = 4/3µm. The normalized frequency, Ω, is defined by Ω = ωD/2πc, where D = dA + dB.

Download Full Size | PDF

However, there are two differences between these two systems. Firstly, by comparing to the fundamental region of the fixed gap in the PBD, it can be found that there are oscillations for the fundamental region of the major gap in the STMS. The fundamental regions of the major gaps approach to a fixed region as the generation order increases. Secondly, the positions of resonant peaks maintain a fixed value in the gap of a PBD with different periodic numbers. However, the positions of resonant peaks appear arbitrarily in the major gap for the STMS with different generation orders.

The gap map diagram for a ν = 6 STMS versus the thickness of the space layer, dS, is shown in Fig. 2(a).The gray areas are the regions of the major gaps in the band structure of the fundamental Thue-Morse supperlattice. There is an important physical relationship between the resonant wavelength and the thickness of the space layer. For shorter resonant wavelength, the appearance of the resonance is in a shorter period as the space layer increases its thickness while a longer period for longer resonant wavelength. In Fig. 2(a), the resonant lines corresponding to different modes shift from a small value of dS to a large value of dS periodically in major gaps. There is periodical variation between adjacent resonant lines in the major gaps. The period function of resonant lines for different values of frequency is written as

PS=(dA+dB)/2nSΩ,
where Ω = ωD/2πc. In Eq. (1), the frequency Ω is inversely related to the period PS. Therefore, the period corresponding to the adjacent resonant lines decreases as the value of frequency increases, as shown in Fig. 2(a). Compared to the PBD, the STMS has different properties that the position of the resonant lines in the gap map changes for different generation orders. However, the resonant lines have a fixed value for arbitrary periodic numbers in the PBD. Moreover, it is seen that more resonant peaks in a gap appear when dS increases.

 figure: Fig. 2

Fig. 2 (a)The plot of the change of the normalized frequency for resonant lines versus the thickness of the space layer for normal incidence in a ν = 6 STMS, for nA = 1, nB = 2. The gray areas correspond to the major gaps of the Thue-Morse superlattice. The signs, A, B, C and D, show the intersection points of the red midline in the major gaps and the black resonant lines. (b) ΦS,U/π in the up major gap and ΦS,D/π in the down major gap, versus the generation orders. The blue solid line and the green dashed line correspond to ΦS,U/π and ΦS,D/π, respectively

Download Full Size | PDF

In order to understand further the relationship between the upper and lower resonant lines in higher generation orders, the intersection of the resonant lines and the middle lines in the major gap are used. The A and B signs are the first and second intersections, respectively, of the resonant lines and the middle lines in the down major gap. Similarly, the C and D signs are the first and second intersections, respectively, of the resonant lines and the middle lines in the up major gap. According to the relationship between dS and PS, the phase ΦS is written as

ΦS=(dS/PS)π.

It is shown that difference of ΦS between points A and B is the multiple of π in Eq. (2). This is also true for points C and D. Similarly, difference of ΦD for two adjacent resonant lines with a fixed value of Ω is also π in the PBD. The phase ΦD in the PBD can be written as ΦD = (dD/PS). Moreover, the values of ΦS oscillate as the generation order increases in the STMS, as shown in Fig. 2(b). They approach a fixed value as the generation order increases. The sum of ΦS,U/π around the three semi-quarter-wave thickness and ΦS,D/π around the semi-quarter-wave thickness is unity for any generation order. The semi-quarter-wave is defined as an intersection of kAdA = π/4 and kBdB = π/4. kA and kB are the wavevectors in layers A and B defined by kA = nAωcos(θA)/c and kB = nBωcos(θB)/c, respectively. θA and θB are the angles of the wavevectors in layers A and B, respectively. However, the positions of the resonant lines are constant for all periodic numbers in gap map in the PBD. Thus, value of ΦD /π does not oscillate, and the sum of ΦD/π around the three quarter-wave thickness and ΦD /π around the quarter-wave thickness is not unity.

Next, we consider the transmission for the STMS in terms of ΦS, as shown in Fig. 3.It is seen that the positions of the resonant peak are arbitrary for the up and down major gaps, in Figs. 3(a) and 3(b), respectively. Although the ΦS of the resonant peak are arbitrary, the period of the ΦS to appear the resonant peak is π for all generation orders. For the PBD, the ΦD of the resonant peak appear on , where m is an integer, in Fig. 3(c). Moreover, the peak becomes sharper for ν in these two systems.

 figure: Fig. 3

Fig. 3 The transmission spectra in the STMS versus ΦS/π at the middle positions are in (a) the up major gap and (b) the down major gap, respectively. (c) The transmission spectra versus ΦD/π at the middle position in the fundamental major gap for the PBD. The dashed red, the dotted green and the solid blue lines correspond to the generation orders ν = 3 to 5 in the STMS and to the numbers of period, m = 2, 4 and 8, in the PBD, respectively. (d) The FWHM and the group delay, S/, on the first whole resonant line in major gap versus the generation orders of the STMS. The black solid line and the blue dotted line correspond to the up and down major gaps for the FWHM. The red solid line and the green dotted line correspond to the up and down major gaps for the group delay

Download Full Size | PDF

The effect on the full width at half maximum (FWHM) and the group delay, S/, for different generation orders in the STMS is of interest. The minimum value for the FWHM and the maximum value for the group delay on the first whole resonant line in major gap are chosen as a reference. The relationship between the FWHM and the group delay for different generation orders is shown in Fig. 3(d). It is seen that the value of the FWHM decreases as the generation order increases. In other words, the sharpness of the resonant peaks increases as the generation order increases. However, the value for the group delay increases as the generation order increases. It is seen that the values of the FWHM and values for the group delay around the semi-quarter-wave thickness and the three semi-quarter-wave thickness are very close. Similarly, there are the same situations for the FWHM and group delay in the PBD. The value of the FWHM also decreases as the periodic number increases. Moreover, the value of group delay also increases as the periodic number increases.

In order to further understand the properties of localization in the one and three semi-quarter-wave thicknesses for the STMS for different generation orders, the squared electric field of the light propagation is considered, as shown in Fig. 4.It is shown that the squared electric fields are almost invariably symmetrical in the one and three semi-quarter-wave thickness for the ν5 STMS shown in Figs. 4(a) and 4(b), and the ν6 STMS shown in Figs. 4(c) and 4(d). In particular, there is stronger light localization simultaneously in the space layer. There is greater localization enhancement as the generation order increases. Similarly, this is also true for the quarter-wave thickness in the PBD with the same thicknesses dA and dB. The space mode can also be changed by changing the thickness of the space layer dS. In other words, one resonant line represents only one mode style. For example, there is one period of wavelength for the squared electric field in the space layer on the sign, C, in Fig. 2(a), but there are two periods of wavelength for the squared electric field in the space layer on the sign, D, in Fig. 2(a). However, the first resonant line is an incomplete line at the down major gap that has a space layer of insufficient thickness, as shown in Fig. 2(a). Thus, the squared electric field in the space layer has an imperfect period of wavelength. Furthermore, the squared electric field adds one period when dS increases from one resonant line to the next. This is also true for the PBD. Moreover, these two structures differ in that the peak values of squared electric field are the same in the one and three quarter-wave thickness in the PBD. However, the peak values of squared electric field are different in the one and three semi-quarter-wave thickness in the STMS.

 figure: Fig. 4

Fig. 4 The |E|2 distribution in the ν5 STMS, for (a) dS = 0.37µm and Ω = 0.519 in the three semi-quarter-wave thickness and (b) dS = 0.245µm and Ω = 0.2315 in the semi-quarter-wave thickness, respectively, and in the ν6 STMS, for (c) dS = 0.43µm and Ω = 0.5 in the three semi-quarter-wave thickness and (d) dS = 0.135µm and Ω = 0.25 in the semi-quarter-wave thickness, respectively, which correspond to the signs, C and A, in Fig. 2. The white areas symbolize the bonding media. The yellow, green and light blue areas symbolize materials A and B and the space layer, respectively.

Download Full Size | PDF

In this paper, we proposed a novel method to determine the occurrence of the localized mode in one-dimensional symmetric Thue-Morse quasicrystals based on the plot of the change of the normalized frequency for resonant lines. We find that there are three properties of the localized mode in the STMS. Firstly, the localized modes exist around odd semi-quarter-wave thickness in the STMS rather than the ones around quarter-wave thickness in a traditional PBD. Secondly, the frequency of localized mode in the transmission spectra of the STMS changes for different generation orders. Lastly, even if the thickness of the space layer in the first localized mode appears randomly for different generation orders, the localized modes appear periodically. The last situation is similar to that in the PBD when the thickness of the space layer is increased for a fixed frequency. We also find that the value of FWHM decreases and the value of group delay increases as the generation order increases. The sharpness of the resonant peaks becomes sharp as the generation order increases. These results show that the optical properties of STMS’s are similar to those for traditional PBD. Thus, they are suitable for use in advanced optical devices. Moreover, the propose method may be extended to investigate the localization problems for other quasiperiodic systems.

Acknowledgments

The authors acknowledge the support in part by the National Science Council of Taiwan under grant numbers NSC 102-2221-E-002-105 and MOST 103-2221-E-002-118.

References and links

1. A. Yariv and P. Yeh, Optical Waves in Crystals: Propagation and Control of Laser Radiation (Wiley, 1988).

2. S. Kalusniak, H. J. Wünsche, and F. Henneberger, “Random semiconductor lasers: scattered versus Fabry-Perot feedback,” Phys. Rev. Lett. 106(1), 013901 (2011). [CrossRef]   [PubMed]  

3. A. P. Hibbins, M. J. Lockyear, and J. R. Sambles, “The resonant electromagnetic fields of an array of metallic slits acting as Fabry-Perot cavities,” J. Appl. Phys. 99(12), 124903 (2006). [CrossRef]  

4. X. R. Huang, D. P. Siddons, A. T. Macrander, R. W. Peng, and X. S. Wu, “Multicavity X-ray Fabry-Perot resonance with ultrahigh resolution and contrast,” Phys. Rev. Lett. 108(22), 224801 (2012). [CrossRef]   [PubMed]  

5. W. J. Hsueh and S. J. Wun, “Simple expressions for the maximum omnidirectional bandgap of bilayer photonic crystals,” Opt. Lett. 36(9), 1581–1583 (2011). [CrossRef]   [PubMed]  

6. Y. Fink, J. N. Winn, S. Fan, C. Chen, J. Michel, J. D. Joannopoulos, and E. L. Thomas, “A dielectric omnidirectional reflector,” Science 282(5394), 1679–1682 (1998). [CrossRef]   [PubMed]  

7. X. Wang, J. Young, Z. Chen, D. Weinstein, and J. Yang, “Observation of lower to higher bandgap transition of one-dimensional defect modes,” Opt. Express 14(16), 7362–7367 (2006). [CrossRef]   [PubMed]  

8. K. K. Law, R. H. Yan, J. L. Merz, and L. A. Coldren, “Normally-off high-contrast asymmetric Fabry–Perot reflection modulator using Wannier–Stark localization in a superlattice,” Appl. Phys. Lett. 56(19), 1886–1888 (1990). [CrossRef]  

9. K. Zhou, Z. Guo, J. Wang, and S. Liu, “Defect modes in defective parity-time symmetric periodic complex potentials,” Opt. Lett. 35(17), 2928–2930 (2010). [CrossRef]   [PubMed]  

10. A. Fainstein, N. D. Lanzillotti-Kimura, B. Jusserand, and B. Perrin, “Strong optical-mechanical coupling in a vertical GaAs/AlAs microcavity for subterahertz phonons and near-infrared light,” Phys. Rev. Lett. 110(3), 037403 (2013). [CrossRef]   [PubMed]  

11. W. Liang, Y. Xu, J. M. Choi, and A. Yariv, “Engineering transverse Bragg resonance waveguides for large modal volume lasers,” Opt. Lett. 28(21), 2079–2081 (2003). [CrossRef]   [PubMed]  

12. C. H. Chang, C. H. Chen, and W. J. Hsueh, “Strong photoluminescence emission from resonant Fibonacci quantum wells,” Opt. Express 21(12), 14656–14661 (2013). [CrossRef]   [PubMed]  

13. J. Mikhael, J. Roth, L. Helden, and C. Bechinger, “Archimedean-like tiling on decagonal quasicrystalline surfaces,” Nature 454(7203), 501–504 (2008). [CrossRef]   [PubMed]  

14. D. Shechtman, I. Blech, D. Gratias, and J. W. Cahn, “Metallic phase with long-range orientational order and no translational symmetry,” Phys. Rev. Lett. 53(20), 1951–1953 (1984). [CrossRef]  

15. L. Mahler, A. Tredicucci, F. Beltram, C. Walther, J. Faist, H. E. Beere, D. A. Ritchie, and D. S. Wiersma, “Quasi-periodic distributed feedback laser,” Nat. Photonics 4(3), 165–169 (2010). [CrossRef]  

16. W. J. Hsueh, C. T. Chen, and C. H. Chen, “Omnidirectional band gap in Fibonacci photonic crystals with metamaterials using a band-edge formalism,” Phys. Rev. A 78(1), 013836 (2008). [CrossRef]  

17. D. Levine and P. J. Steinhardt, “Quasicrystals: a new class of ordered structures,” Phys. Rev. Lett. 53(26), 2477–2480 (1984). [CrossRef]  

18. W. J. Hsueh, C. H. Chang, Y. H. Cheng, and S. J. Wun, “Effective Bragg conditions in a one-dimensional quasicrystal,” Opt. Express 20(24), 26618–26623 (2012). [CrossRef]   [PubMed]  

19. L. Dal Negro, M. Stolfi, Y. Yi, J. Michel, X. Duan, L. C. Kimerling, J. LeBlanc, and J. Haavisto, “Photon band gap properties and omnidirectional reflectance in Si/SiO2 Thue–Morse quasicrystals,” Appl. Phys. Lett. 84(25), 5186–5188 (2004). [CrossRef]  

20. T. Ma, C. Liang, L.-G. Wang, and H.-Q. Lin, “Electronic band gaps and transport in aperiodic graphene superlattices of Thue-Morse sequence,” Appl. Phys. Lett. 100(25), 252402 (2012). [CrossRef]  

21. N.-H. Liu, “Propagation of light waves in Thue-Morse dielectric multilayers,” Phys. Rev. B 55(6), 3543–3547 (1997). [CrossRef]  

22. C. Fort, L. Fallani, V. Guarrera, J. E. Lye, M. Modugno, D. S. Wiersma, and M. Inguscio, “Effect of optical disorder and single defects on the expansion of a Bose-Einstein condensate in a one-dimensional waveguide,” Phys. Rev. Lett. 95(17), 170410 (2005). [CrossRef]   [PubMed]  

23. C. W. Tsao, W. J. Hsueh, C. H. Chang, and Y. H. Cheng, “Quasi-Bragg conditions in Thue-Morse dielectric multilayers,” Opt. Lett. 38(22), 4562–4565 (2013). [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 The transmission spectra in the STMS’s with generation orders (a) ν = 5, and (b) ν = 6, for normal incidence from air. The gray areas correspond to the region of the major gap in the band structure of the Thue-Morse superlattice. The |E|2 distribution in the ν5 STMS for (c) Ω = 0.2308 and (d) Ω = 0.5192, respectively, which correspond to the signs, A and B, in Fig. 1(a). The parameters of the system are nA = 1, nB = 2, nD = 2, dA = 2/3µm, dB = 1/3µm, and dS = 4/3µm. The normalized frequency, Ω, is defined by Ω = ωD/2πc, where D = dA + dB.
Fig. 2
Fig. 2 (a)The plot of the change of the normalized frequency for resonant lines versus the thickness of the space layer for normal incidence in a ν = 6 STMS, for nA = 1, nB = 2. The gray areas correspond to the major gaps of the Thue-Morse superlattice. The signs, A, B, C and D, show the intersection points of the red midline in the major gaps and the black resonant lines. (b) ΦS,U/π in the up major gap and ΦS,D/π in the down major gap, versus the generation orders. The blue solid line and the green dashed line correspond to ΦS,U/π and ΦS,D/π, respectively
Fig. 3
Fig. 3 The transmission spectra in the STMS versus ΦS/π at the middle positions are in (a) the up major gap and (b) the down major gap, respectively. (c) The transmission spectra versus ΦD/π at the middle position in the fundamental major gap for the PBD. The dashed red, the dotted green and the solid blue lines correspond to the generation orders ν = 3 to 5 in the STMS and to the numbers of period, m = 2, 4 and 8, in the PBD, respectively. (d) The FWHM and the group delay, S/, on the first whole resonant line in major gap versus the generation orders of the STMS. The black solid line and the blue dotted line correspond to the up and down major gaps for the FWHM. The red solid line and the green dotted line correspond to the up and down major gaps for the group delay
Fig. 4
Fig. 4 The |E|2 distribution in the ν5 STMS, for (a) dS = 0.37µm and Ω = 0.519 in the three semi-quarter-wave thickness and (b) dS = 0.245µm and Ω = 0.2315 in the semi-quarter-wave thickness, respectively, and in the ν6 STMS, for (c) dS = 0.43µm and Ω = 0.5 in the three semi-quarter-wave thickness and (d) dS = 0.135µm and Ω = 0.25 in the semi-quarter-wave thickness, respectively, which correspond to the signs, C and A, in Fig. 2. The white areas symbolize the bonding media. The yellow, green and light blue areas symbolize materials A and B and the space layer, respectively.

Equations (2)

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

P S = ( d A + d B ) / 2 n S Ω ,
Φ S =( d S / P S )π.
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.