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

Investigation of Ge20Sb15Se65 photonic crystal slab waveguides with slow light at infrared wavelength

Open Access Open Access

Abstract

The slow light propagation in Ge20Sb15Se65 chalcogenide photonic crystal slab waveguides of air holes have been investigated. The obtained slow-light waveguides can be divided into two categories by perturbing the holes adjacent to the waveguide core: symmetric and asymmetric waveguides. With a bandwidth of 3~30 nm at the center wavelength of 3 μm, it is possible to achieve the group index of 16~43 within 20% in symmetric waveguides, and the group index can be increased up to 130 in asymmetric ones. The result shows perfect slow-light properties in chalcogenide PCSWs and can be used as affordable reference for further research.

©2013 Optical Society of America

1. Introduction

Chalcogenide glasses (ChGs) are infrared transmitting materials containing chalcogenide elements S, Se or Te, compounded with elements as As, Sb and Ge [13]. They are transparent at infrared wavelengths (0.8~20 μm), exhibiting high optical Kerr nonlinearity (the nonlinear index n2 = 2~20 × 10−18 m2/W, 100~1000 times larger than that of silica), subpicosecond response time (tr < 200 fs) as well as negligible two-photon absorption [46]. These characteristics make ChGs a promising platform for integrated all-optical nonlinear devices at infrared wavelengths.

Photonic crystal slab waveguides (PCSWs), which utilize the band gap of a periodic structure to confine light in the horizontal direction while guiding light vertically by total internal reflection, can enhance nonlinear effects by exploiting its remarkable slow light (large group index) effect. As the practical utilization of ultra-slow light in PCSWs is limited due to an inherent small bandwidth, the balance of group index and bandwidth should be considered. It has been demonstrated that the slow-light properties of SOI-based PCWs at λ = 1.55 μm can be altered via a structural tuning of the waveguide geometry, typically by changing the waveguide width or by introducing bi-periodicity [712]. However, the slow-light properties had rarely been investigated in chalcogenide PCSWs, though the enhancement of self-phase modulation and third harmonic generation were demonstrated at λ = 1.55 μm [1315].

In this paper, we theoretically investigate the slow-light properties of Ge20Sb15Se65 chalcogenide PCSWs at λ = 3.0 μm, and both of symmetric and asymmetric structure are studied via a simple method of tuning the hole radius in the first two rows next to the waveguide. Our results show perfect slow-light properties in asymmetric PCSWs and can be used as affordable reference for further research.

2. Theoretical method

The proposed chalcogenide PCSW is shown in Fig. 1(a), which consists of a two-dimensional triangular lattice of air holes and a line defect in a Ge20Sb15Se65 layer, with air as the top and bottom cladding layers. The parameters of the chalcogenide PCSW are set as below: a - the lattice period, r – the radius of air holes, ru – radius of one row air holes adjacent to the line defect waveguide, rd - radius of another row air holes adjacent to the line defect waveguide, h – the thickness of slab, n - refractive index of Ge20Sb15Se65 slab, and λ - operating wavelength. This kind waveguide structures can avoid large errors caused by shifting holes compared to PCSWs based on the lateral or longitudinal shift of the first two rows of holes adjacent to the waveguides, in which the distance of holes shift is very small, only dozens of nanometer, and therefore errors can be easily introduced just with not serious roughness [9,12,16]. In addition, with only the radius of holes changed, this kind slow-light waveguide structure can be very simple and easily calculated by plane wave expansion method. Table 1 shows the refractive indices of Ge20Sb15Se65 glass fabricated in our laboratory, and n = 2.612 at λ = 3.0 μm is selected in this work. The three-dimensional plane wave expansion (PWE) method is used to calculate the dispersion relations ω(k) of quasi-TE mode(even mode) in the chalcogenide PCSW. As shown in Fig. 1(b), even modes exhibit a slab band gap where there is a guided mode under the light cone. Numerical result shows that PCSW with r = 0.3a and h = 0.5a can support a slab band gap large enough to contain guided modes.

 figure: Fig. 1

Fig. 1 (a) Schematic diagram of the proposed photonic crystal slab waveguide with index of 2.612 fully embedded in air claddings and (b) calculated dispersion curve of the PCSW with r = 0.3a,h = 0.5a, ru = 0.4a and rd = 0.2a.

Download Full Size | PDF

Tables Icon

Table 1. Refractive indices of Ge20Sb15Se65 glass at different wavelength

Slow light in the periodic waveguide is explained as an interaction between the forward and backward propagating modes [17]. The strongest interaction is at the edge of the Brillouin zone where the group velocity vanishes. When the guided mode is moved away from the zone boundary, the interaction becomes weaker, due to phase mismatch between the forward and backward propagating modes. The group velocity vg of propagation light with frequency ω in an optical waveguide can easily be obtained from the dispersion relation:

vg=cng=dωdk
where k is the wavevector along the waveguide and ng is the group index. In order to get small vg, a large ng should be achieved from the equation. It is possible to increase the group index ng or the useful normalized bandwidth Δω/ω0 by perturbing the holes adjacent to the line defect waveguide [7], where Δω is the absolute frequency bandwidth of the slow light region and ω0 is the middle normalized frequency. Usually, PCSW is difficult to achieve with high group index and large normalized bandwidth simultaneously and so there is a balance between the group index and normalized bandwidth. The normalized delay-bandwidth product (NDBP) is introduced to represent this slow-light property [10,18]:
NDBP=n¯g×Δωω0
where n¯gis the average group index. Generally, a large NDBP is necessary for flat band slow-light PCSW.

3. Numerical simulation and discussion

The band diagram for even modes in PCSW with r = 0.3a and h = 0.5a is shown in Fig. 2(a). There are three guided modes in the slab band gap under the light cone, named band 1, band 2, and band 3 respectively. Clearly, the waveguide remains single mode because these guided modes do not overlap in the normalized frequency. From the figure, band 1 tends to leak into slab modes and enlarges the transmission loss eventually as it is closed to slab band gap edge. Figure 2(b) shows the corresponding group indices ng calculated by relation (1) of these guided modes. Compared to band 1 and band 3, band 2 has a much larger group index with the same propagation constant, though its bandwidth is slightly narrower.

 figure: Fig. 2

Fig. 2 (a) Dispersion relation and (b) group index of different modes in PCSW with r = 0.3a and h = 0.5a.

Download Full Size | PDF

It is known through calculation that by tuning the size of the air holes adjacent to the waveguide core, it is difficult to increase the group indices of band 1 or band 3(Our calculation shows that their ng cannot be increased above ~12 within 20% with a certain bandwidth), whereas it is easy to change the bandwidth of band 2 to obtain relative large value both in bandwidth and group index. In order to obtain relative high value in both of bandwidth and group index, only the band 2 will be concerned in our following study. Next, desired PCSWs with large bandwidth and high group index will be demonstrated, which can be achieved by perturbing the holes adjacent to the waveguide core. To simplify matters, the work can be processed in two steps: first, investigating slow light properties of symmetric PCSWs by only changing rs from 0 to 0.5a, where rs = ru = rd; second, changing ru and rd respectively to study slow light properties of asymmetric PCSWs.

The dispersion relations of the guided modes in symmetric PCSWs with rs ranging from 0 to 0.5a can be seen from Fig. 3(a). In the figure, the guided modes with different rs represented by the colorized curves are almost in the slab band gap except the mode with rs = 0.5a, which demonstrates that the chosen value of rs is acceptable to achieve a large bandwidth and a high group index for symmetric PCSWs, with r = 0.3a and h = 0.5a. Obviously, the frequency of the guided mode decreases when rs decreases, as the effective index increase.

 figure: Fig. 3

Fig. 3 (a) Curves of the guided modes for different rs and (b) curves of the group indices for different rs

Download Full Size | PDF

As can be seen from Fig. 3(b), PCSWs have large bandwidth Δω/ω0 with small rs (0.1a ~0.25a) and high refractive index with large rs (0.3a ~0.4a). As a result, we are only concerned about modes with 0.2ars ≤ 0.4a because they have a relative large group index above 15, and some of them have a wide band. There is a phenomenon in Fig. 3(b) that guided mode will has a larger bandwidth when rs is away from the radius of photonic crystal slab air holes(0.3a), as the disorder degree increase. When rs < 0.3a, the guided mode will have a relatively smaller group index, and vice versa. From the analysis above, we can tune dispersion properties properly to achieve the desired group velocity with a relative large bandwidth, by increasing ru and decreasing rd simultaneously, which brings out an asymmetric structure.

The results of asymmetric PCSWs are depicted in Fig. 4, which shows the guided mode group index as a function of the normalized frequency ω with ru and rd varied from 0.2a to 0.4a. For comparison, the group indices of symmetric PCSWs with ru = rd = 0.2a, 0.3a and 0.4a are also included in this figure.

 figure: Fig. 4

Fig. 4 the curves of the group index for different ru and rd.

Download Full Size | PDF

As can be seen in Fig. 4, most curves have inflection points where a relative large bandwidth will be achieved, and thus the perturbed PCSW can support the desired group index with a certain bandwidth. It is very clear in Fig. 4 that the group index and bandwidth can hardly be increased simultaneously, for increasing group index will result in the decrease of bandwidth, and vice versa. Keeping ru = 0.2a and increasing rd from 0.2a to 0.4a, the group index at inflection point is increased from ~15 to 35 within 20% over bandwidth of ~30 nm to 10 nm at λ = 3.0 μm, while keeping ru = 0.4a and decreasing rd from 0.4a to 0.2a, the group index at inflection point is increased rapidly from ~40 to 120 within 20% with a relatively narrow bandwidth. The detailed data are shown in Table 2. Comparing to symmetric ones, the asymmetric PCSWs can increase guided mode group index in a larger range, especially for the structure with ru = 0.4a. We also notice that the NDBPs are not changed remarkably by breaking the symmetry of PCSWs in this situation, and so we can selectively tailor the slow-light properties of a PCSW with specific NDBP ranging from 0.1 to 0.2.

Tables Icon

Table 2. Slow light properties of symmetric and asymmetric PCSWs at λ = 3 μm

In the structure designation, disorders of period (a = λ*ω) and radius (ru and rs) are very critical and have great impacts on middle normalized frequency or the position of guided mode, resulting in the change of the bandwidth and group index of slow light, according to Fig. 4. With laser of a certain linewidth (Δλ), the tolerance to disorder of the period (Δa) can be simply achieved by:

Δaa=Δλλ
In Fig. 4, the middle normalized frequency ω is ranging from 0.35 to 0.39. If , then Δa is ranging from 10.5 nm to 11.7 nm. In this paper, the tolerance of the radius disorder can be obtained by the range of group index according to Fig. 4 and Table 2. For instance, if small group index of 10~30 is needed, then we can selected the radius rd in 0.2a~0.3a and ru in 0.2a~0.25a.

In practical applications, the slow light loss should be considered. Compared to the common 1550nm band, mid-infrared wavelengths (e.g. λ = 3.0 μm) can generally reduce slow light loss according to equations below [19, 20]:

α=α1ng+α2ng2
in the equations, α is loss coefficient, α1 and α2 are radiation and backscattering loss factors, respectively, where α11/λ2 and α21/λ4. The backscattering loss generally dominates the radiation loss when ng >10 [19]. Therefore, for given ng which is bigger than 10 in our designed PCSWs, wavelength of 3.0 μm generates one sixteenth the loss of the common 1550nm, which suggests slow-light waveguides with mid-infrared wavelength have a huge advantage in terms of the propagation loss over ones with 1550nm.

On the other hand, for given operating wavelength, slow-light loss is proportional to ng when ng is small, and is proportional to ng2 when ng is large. This can guide us to select suitable geometry parameters of PCSWs in practical applications. If relative small group index can introduce nonlinearity large enough, then it is not necessary to select lager group index, for slow-light loss would increases rapidly when group index become larger. In this situation, symmetric PCSW with ru = rd = 0.2a or asymmetric PCSW with ru = 0.2a and rd = 0.25a ~0.30a can be suitable choices. If we want to achieve super large nonlinearity enhanced by slow-light, then asymmetric PCSW with ru = 0.4a and rd = 0.2a ~0.3a would be a good choice.

4. Conclusion

In this work, we present a photonic crystal slab slow-light waveguide based on Ge20Sb15Se65 chalcogenide glass at a mid-infrared wavelength of λ = 3.0 μm. By changing the size of the first two rows of air holes adjacent to the line defect waveguide, two kinds of structures, symmetric and asymmetric waveguide, can be achieved. These waveguides have very similar NDBP between 0.1 and 0.2. The symmetric waveguides has a range of group indices, typically between 16 and 43 within 20% over bandwidth of 30 nm to 10 nm, and the group index can be increased up to 130 in asymmetric waveguides, which is more attractive in structure design. Our work shows the possibilities of using chalcogenide PCSWs in the slow-light regime for practical applications, especially in the enhancement of nonlinear effects.

Acknowledgments

This work has been supported by the National Natural Science Foundation of China (No. 61107047), Natural Science Foundation of Zhejiang Provincial (No. LQ12F05004), Ningbo Optoelectronic Materials and Devices Creative Team (No.2009B21007), and the Magna Fund sponsored by K.C. Wong in Ningbo University of China.

References and links

1. D. Freeman, S. Madden, and B. Luther-Davies, “Fabrication of planar photonic crystals in a chalcogenide glass using a focused ion beam,” Opt. Express 13(8), 3079–3086 (2005). [CrossRef]   [PubMed]  

2. J. T. Gopinath, M. Soljačić, E. P. Ippen, V. N. Fuflyigin, W. A. King, and M. Shurgalin, “Third order nonlinearities in Ge-As-Se-based glasses for telecommunications applications,” J. Appl. Phys. 96(11), 6931 (2004). [CrossRef]  

3. Y. Chen, X. Shen, R. Wang, G. Wang, S. Dai, T. Xu, and Q. Nie, “Optical and structural properties of Ge-Sb-Se thin films fabricated by sputtering and thermal evaporation,” J. Alloys Comp. (2012).

4. A. Seddon, “Chalcogenide glasses: a review of their preparation, properties and applications,” J. Non-Cryst. Solids 184, 44–50 (1995). [CrossRef]  

5. F. Smektala, C. Quemard, V. Couderc, and A. Barthélémy, “Non-linear optical properties of chalcogenide glasses measured by Z-scan,” J. Non-Cryst. Solids 274(1-3), 232–237 (2000). [CrossRef]  

6. A. Zakery and S. Elliott, “Optical properties and applications of chalcogenide glasses: a review,” J. Non-Cryst. Solids 330(1-3), 1–12 (2003). [CrossRef]  

7. L. H. Frandsen, A. V. Lavrinenko, J. Fage-Pedersen, and P. I. Borel, “Photonic crystal waveguides with semi-slow light and tailored dispersion properties,” Opt. Express 14(20), 9444–9450 (2006). [CrossRef]   [PubMed]  

8. A. Säynätjoki, M. Mulot, J. Ahopelto, and H. Lipsanen, “Dispersion engineering of photonic crystal waveguides with ring-shaped holes,” Opt. Express 15(13), 8323–8328 (2007). [CrossRef]   [PubMed]  

9. J. Li, T. P. White, L. O’Faolain, A. Gomez-Iglesias, and T. F. Krauss, “Systematic design of flat band slow light in photonic crystal waveguides,” Opt. Express 16(9), 6227–6232 (2008). [CrossRef]   [PubMed]  

10. B. Meng, L. L. Wang, W. Q. Huang, X. F. Li, X. Zhai, and H. Zhang, “Wideband and low dispersion slow-light waveguide based on a photonic crystal with crescent-shaped air holes,” Appl. Opt. 51(23), 5735–5742 (2012). [CrossRef]   [PubMed]  

11. S. Rahimi, A. Hosseini, X. Xu, H. Subbaraman, and R. T. Chen, “Group-index independent coupling to band engineered SOI photonic crystal waveguide with large slow-down factor,” Opt. Express 19(22), 21832–21841 (2011). [CrossRef]   [PubMed]  

12. Y. Hamachi, S. Kubo, and T. Baba, “Slow light with low dispersion and nonlinear enhancement in a lattice-shifted photonic crystal waveguide,” Opt. Lett. 34(7), 1072–1074 (2009). [CrossRef]   [PubMed]  

13. C. Monat, M. Spurny, C. Grillet, L. O’Faolain, T. F. Krauss, B. J. Eggleton, D. Bulla, S. Madden, and B. Luther-Davies, “Third-harmonic generation in slow-light chalcogenide glass photonic crystal waveguides,” Opt. Lett. 36(15), 2818–2820 (2011). [CrossRef]   [PubMed]  

14. K. Suzuki, Y. Hamachi, and T. Baba, “Fabrication and characterization of chalcogenide glass photonic crystal waveguides,” Opt. Express 17(25), 22393–22400 (2009). [CrossRef]   [PubMed]  

15. K. Suzuki and T. Baba, “Nonlinear light propagation in chalcogenide photonic crystal slow light waveguides,” Opt. Express 18(25), 26675–26685 (2010). [CrossRef]   [PubMed]  

16. F. Wang, J. S. Jensen, J. Mørk, and O. Sigmund, “Systematic design of loss-engineered slow-light waveguides,” J. Opt. Soc. Am. A 29(12), 2657–2666 (2012). [CrossRef]   [PubMed]  

17. T. F. Krauss, “Slow light in photonic crystal waveguides,” J. Phys. D Appl. Phys. 40(9), 2666–2670 (2007). [CrossRef]  

18. T. Baba, “Slow light in photonic crystals,” Nat. Photonics 2(8), 465–473 (2008). [CrossRef]  

19. W. Song, R. A. Integlia, and W. Jiang, “Slow light loss due to roughness in photonic crystal waveguides: An analytic approach,” Phys. Rev. B 82(23), 235306 (2010). [CrossRef]  

20. J. Tan, R. A. Soref, and W. Jiang, “Interband scattering in a slow light photonic crystal waveguide under electro-optic tuning,” Opt. Express 21(6), 6756–6763 (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 (a) Schematic diagram of the proposed photonic crystal slab waveguide with index of 2.612 fully embedded in air claddings and (b) calculated dispersion curve of the PCSW with r = 0.3a,h = 0.5a, ru = 0.4a and rd = 0.2a.
Fig. 2
Fig. 2 (a) Dispersion relation and (b) group index of different modes in PCSW with r = 0.3a and h = 0.5a.
Fig. 3
Fig. 3 (a) Curves of the guided modes for different rs and (b) curves of the group indices for different rs
Fig. 4
Fig. 4 the curves of the group index for different ru and rd.

Tables (2)

Tables Icon

Table 1 Refractive indices of Ge20Sb15Se65 glass at different wavelength

Tables Icon

Table 2 Slow light properties of symmetric and asymmetric PCSWs at λ = 3 μm

Equations (4)

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

v g = c n g = dω dk
NDBP= n ¯ g × Δω ω 0
Δ a a = Δ λ λ
α= α 1 n g + α 2 n g 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.