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

Wide band dispersionless slow light in hetero-MIM plasmonic waveguide

Open Access Open Access

Abstract

A flat slow-light band over a wide frequency range is obtained in the hetero-MIM (metal-insulator-metal) waveguide with zero group velocity dispersion (GVD). The zero GVD originates from dispersion compensation by the photonic mode and the plasmonic mode, the mechanism of which does not exist in the homo-MIM structure. By changing dielectric permittivity of the insulator or the difference of two different metallic plasma frequencies, the group index and the bandwidth can be tuned. The dispersionless slow light characteristic in the hetero-MIM waveguide may be useful in the new design of plasmonic devices.

© 2016 Optical Society of America

1. Introduction

Surface plasmon polaritons (SPPs) are electromagnetic waves propagating along the interface between metal and insulator [1], which have applications in nonlinear optics, sensors, nanoimaging and slow light [2–6]. The simplest structure supporting SPPs is a single metal-insulator interface, whereas several derivatives have been investigated in the past [7–13]. Among these structures, metal-insulator-metal (MIM) waveguides have been intensively studied due to stronger field confinement, and most researches have focused on homo-MIM structures with the same metal claddings. However, hetero-MIM structures with different metal claddings were found interesting in many applications in recent years. Shin et al. demonstrated the imaging process for surface plasmons with a hetero-MIM lens [14]. Lezec et al. experimentally realized negative refraction at visible frequency by an asymmetric MIM waveguide [15]. Tai et al. achieved giant angular dispersion due to modal competition in heterogeneously coupled MIM waveguides [16]. Kim studied the guided dispersion relations of Ag-Air-Al for lossless case when varying the gap width, providing an initial picture of the guided dispersion curves of hetero-MIM waveguide in a wide frequency range [17]. We thus intend to investigate the dispersion properties of the hetero-MIM waveguide and explore new phenomena in this work.

In this letter, we reported a flat slow-light band over a wide frequency range in the hetero-MIM waveguide, supporting nearly constant group index and the zero group velocity dispersion (GVD). The dispersion relations of the hetero-MIM and the homo-MIM waveguides were studied and compared. It was found that the zero GVD originates from the dispersion compensation with opposite signs of GVDs of the photonic mode and the plasmonic mode. By changing the permittivity of the middle insulator layer or the difference of two different metallic plasma frequencies, the group index and the bandwidth could be tuned. We used the indium tin oxide (ITO) for the calculations in this study, but results can be similar by applying other plasmonic materials in the hetero-MIM structure. To our knowledge, the dispersionless slow light characteristic and its mechanism in the hetero-MIM planar waveguide, which may find its applications in plasmonic devices, had not been reported previously.

2. Structure model

Figure 1 illustrates the hetero-MIM plasmonic waveguide composed of an insulator layer sandwiched between the semi-infinite substrate and superstrate of different metals. The dielectric permittivity and the thickness of the insulator layer is εd and d, respectively. Silicon dioxide (εd = 2.25) is chosen as the insulator with thickness d = 150nm. The two different metals are assumed to be indium tin oxide (ITO) with different carrier concentrations. The permittivity of ITO is characterized by the Drude model εm(ω)=εωp2/(ω2+iωγ), where ε is the dielectric constant at infinite frequency, ωp is the plasma frequency, γ is the damping constant, and ω is the angular frequency of the incident light. For εm2, we set ωp2 = 2.03eV, ε = 4, γ = 0.06eV [18]. The parameters for εm1 are assumed to be ωp1 = 1.2eV, ε = 4, γ = 0.06eV for the calculations and the plasma frequency can be changed by varying the molar fraction of tin (Sn) [19]. The propagation is in the x direction and the magnetic field is along y-axis. The dispersion relations and field distributions are calculated using the transfer matrix method for the TM eigenmode solution [20], in which the dispersion equation is:

γ˜sn˜xxs2m11+γ˜cn˜xxc2m22m21γ˜sγ˜cn˜xxs2n˜xxc2m12=0,
where γ˜s=±γ˜2k02n˜yys2, γ˜c=±γ˜2k02n˜yyc2, and n˜yys and n˜yyc are the substrate and cover complex refractive indices along the y direction, respectively. The zeros of Eq. (1) which can be solved via Cauchy integration method (CIM) correspond to the complex propagation constans γ˜=k0(βiα)where β and α are the normalized phase and attenuation constants, respectively, and k0 = 2π/λ0 is the free space wavenumber.

 figure: Fig. 1

Fig. 1 Schematic of the hetero-MIM plasmonic waveguide.

Download Full Size | PDF

3. Results and discussions

In Fig. 2(a), we illustrate the real part of the dispersion relations for the hetero-MIM and homo-MIM structures with parameters of ωp1 = 1.2eV, ωp2 = 2.03eV, ε = 4 and γ = 0.06eV. The dispersion relation of a homo-MIM waveguide [7, 21] is well known to have the lower and the upper branches connected through a regime known as the forbidden band where the curve turns over in accompany with extremely large absorption. Figure 2(b) shows the forbidden bands above their surface plasmon resonance frequencies ωsp1=ωp1/ε+εd and ωsp2=ωp2/ε+εd for the εm1-εd-εm1 and εm2-εd-εm2 homo-MIM waveguides, respectively, with large imaginary parts [22]. For the εm1-εd-εm2 hetero-MIM waveguide, the dispersion curve in Fig. 2(a) approximately follows that of the εm1-εd-εm1 homo-MIM waveguide first and then above the forbidden band tends to follow that of the εm2-εd-εm2 waveguide. Interestingly between the forbidden bands of the two homo-MIM waveguides, the curve of the hetero-MIM waveguide becomes flattened, presenting the slow-light feature. A slow-light passband is thus found between the two forbidden bands in addition to conclusions given in [17] for the hetero-MIM waveguide.

 figure: Fig. 2

Fig. 2 (a) Real part and (b) Imaginary part of the dispersion relations for εm1-εd-εm2 (solid line) waveguide with ωp1 = 1.2eV, ωp2 = 2.03eV, ε = 4, and γ = 0.06eV. Results for εm1-εd-εm1 (dashed line) and εm2-εd-εm2 (dashed-dotted line) waveguides are presented for comparison. The dotted line represents the light line in silicon dioxide.

Download Full Size | PDF

Figure 3 shows the dispersion relations for the εm1-εd-εm2 waveguide in the lossless and lossy cases with above parameters. As shown in Fig. 3(a), finite damping results in modification of the curve shape at ωsp1 and ωsp2. However, in both cases, the lineshape keeps the same in regime ω1~ωsp2 with ω1=ωp1/ε, i.e. with opposite curvatures toward both ends, and thus is not affected by the damping constant. Corresponding to the imaginary part in Fig. 3(b), in the lossy case, the FOM = ∣Real(kx)/Imag(kx)∣in the flat slow-light band can reach 5. The small FOM is limited by the large damping constant γ of ITO. In contrast, the FOM becomes infinity due to the zero imaginary part in the lossless case, which means low loss can be achieved with low-loss plasmonic materials.

 figure: Fig. 3

Fig. 3 (a) Real part and (b) Imaginary part of the dispersion relations for the εm1-εd-εm2 waveguide in the lossless (dash line) and lossy cases (solid line) with ωp1 = 1.2eV, ωp2 = 2.03eV, and ε = 4 and FOMs (dashed-dotted line) in the lossy case.

Download Full Size | PDF

The slow light characteristics of the εm1-εd-εm2 hetero-MIM waveguide were further investigated in Fig. 4, which shows the group index ng = c/νg = cdk/ as the slow down factor and the GVD β2 = d2k/2 as a measure of pulse distortion. At ω = 0.68eV, there are ng ≈8.37 and β2 = 0. For greater or less frequency, ng increases while β2 changes its sign. That means a pulse with center frequency ω = 0.68eV can propagate in the waveguide with almost no distortion. For 10% variation with respect to the mean value of ng and the normalized bandwidth Δf/f [23], the bandwidth here is around Δf = 16.9 THz and Δf/f = 0.102, in contrast to Δf = 8.6THz and Δf/f = 0.024 in MIM waveguide with two stubs based on analogue of electromagnetically induced transparency (EIT) [24].

 figure: Fig. 4

Fig. 4 Dispersion relation (solid line), group index (dashed line), and GVD (dashed-dotted line) in the εm1-εd-εm2 waveguide. The parameters are the same as that in Fig. 2(a).

Download Full Size | PDF

The evolution of the magnetic field distribution in the hetero-MIM waveguide for varying frequency is illustrated in Fig. 5. At ω = 0.62eV, according to the Drude model, the real part of εm1 becomes positive, so no SPPs exist at the εm1-εd interface. The εm1-εd-εm1 homo-MIM waveguide shows a field of photonic mode at this frequency, which has negative GVD as seen in Fig. 2(a). In the hetero-MIM waveguide, the field is modified from the photonic mode due to the existence of SPPs at the εm2-εd interface with positive GVD. With opposite signs of GVDs for the photonic mode and the plasmonic mode, dispersion compensation occurs and reaches balance at ω = 0.68eV, at which the hetero-MIM waveguide presents dispersionless slow light characteristic. For ω > 0.68eV, the plasmonic mode effect becomes dominant and the field distribution, i.e. at ω = 0.78eV as shown in Fig. 5, is merely of SPPs at the εm2-εd interface, so the GVD becomes positive. In contrast, the dispersionless characteristic due to dispersion compensation does not exist in the homo-MIM structure. Such dispersionless slow light mechanism is also different from that in the plasmonic waveguide formed by single metal and several insulator layers [9, 12], where the mode moves asymptotically to different SPPs supported between the metal and different insulator layers.

 figure: Fig. 5

Fig. 5 Field distribution profiles for various frequencies in the εm1-εd-εm2 waveguide. The field at ω = 0.62eV in the εm1-εd-εm1 waveguide is also presented.

Download Full Size | PDF

The group index ng and the bandwidth Δf versus the permittivity of insulator εd for zero GVD are plotted in Fig. 6(a). As we know, ωsp2 decreases with increasing εd and thus the separation between the two forbidden bands shrinks, leading to more flattened dispersion curve and higher ng. At the same time, the bandwidth Δf decreases, which shows the trade-off relationship between delay and bandwidth [25]. Changing the difference value Δωp = ωp1- ωp2 will produce the similar effect, as shown in Fig. 6(b). Although ng obtained in this study is smaller than the values reported in [24], the bandwidth here is much greater and the structure is quite simple.

 figure: Fig. 6

Fig. 6 (a) Group index ng and bandwidth Δf versus εd with ωp1 = 1.2eV and ωp2 = 2.03 eV. (b) Group index ng and bandwidth Δf versus Δωp with ωp1 = 1.2eV and εd = 2.25. The zero GVD is kept.

Download Full Size | PDF

4. Conclusion

In summary, we investigate the flat slow-light band over a wide frequency range with nearly constant group index and zero GVD in the hetero-MIM waveguide. The zero GVD property originates from dispersion compensation by the photonic mode and the plasmonic mode. Such an effect is unique in the hetero-MIM structure while does not exist in the homo-MIM structure. In addition, by changing dielectric permittivity of the insulator or the difference of plasma frequencies of the two plasmonic materials, the group index and the bandwidth can be tuned. The hetero-MIM structures may raise interest in applications that appreciate the slow light effect.

Funding

National Natural Science Foundation of China (61377055, 11404263); Fundamental Research Funds for the Central Universities (3102015ZY060).

References and links

1. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [CrossRef]   [PubMed]  

2. M. Kauranen and A. V. Zayats, “Nonlinear plasmonics,” Nat. Photonics 6(11), 737–748 (2012). [CrossRef]  

3. D. K. Gramotnev and S. I. Bozhevolnyi, “Plasmonics beyond the diffraction limit,” Nat. Photonics 4(2), 83–91 (2010). [CrossRef]  

4. S. Kawata, Y. Inouye, and P. Verma, “Plasmonics for near-field nano-imaging and superlensing,” Nat. Photonics 3(7), 388–394 (2009). [CrossRef]  

5. M. Sandtke and L. Kuipers, “Slow guided surface plasmons at telecom frequencies,” Nat. Photonics 1(10), 573–576 (2007). [CrossRef]  

6. M. I. Stockman, “Nanofocusing of optical energy in tapered plasmonic waveguides,” Phys. Rev. Lett. 93(13), 137404 (2004). [CrossRef]   [PubMed]  

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

8. B. Prade, J. Y. Vinet, and A. Mysyrowicz, “Guided optical waves in planar heterostructures with negative dielectric constant,” Phys. Rev. B Condens. Matter 44(24), 13556–13572 (1991). [CrossRef]   [PubMed]  

9. A. Karalis, E. Lidorikis, M. Ibanescu, J. D. Joannopoulos, and M. Soljacić, “Surface-plasmon-assisted guiding of broadband slow and subwavelength light in air,” Phys. Rev. Lett. 95(6), 063901 (2005). [CrossRef]   [PubMed]  

10. E. P. Fitrakis, T. Kamalakis, and T. Sphicopoulos, “Slow light in insulator-metal-insulator plasmonic waveguides,” J. Opt. Soc. Am. B 28(9), 2159–2164 (2011). [CrossRef]  

11. M. S. Jang and H. Atwater, “Plasmonic rainbow trapping structures for light localization and spectrum splitting,” Phys. Rev. Lett. 107(20), 207401 (2011). [CrossRef]   [PubMed]  

12. A. Karalis, J. D. Joannopoulos, and M. Soljacić, “Plasmonic-dielectric systems for high-order dispersionless slow or stopped subwavelength light,” Phys. Rev. Lett. 103(4), 043906 (2009). [CrossRef]   [PubMed]  

13. M. I. Stockman, “Slow propagation, anomalous absorption, and total external reflection of surface plasmon polaritons in nanolayer systems,” Nano Lett. 6(11), 2604–2608 (2006). [CrossRef]   [PubMed]  

14. H. Shin and S. Fan, “All-angle negative refraction for surface plasmon waves using a metal-dielectric-metal structure,” Phys. Rev. Lett. 96(7), 073907 (2006). [CrossRef]   [PubMed]  

15. H. J. Lezec, J. A. Dionne, and H. A. Atwater, “Negative refraction at visible frequencies,” Science 316(5823), 430–432 (2007). [CrossRef]   [PubMed]  

16. C. Y. Tai, W. H. Yu, and S. H. Chang, “Giant angular dispersion mediated by plasmonic modal competition,” Opt. Express 18(24), 24510–24515 (2010). [CrossRef]   [PubMed]  

17. K. Y. Kim, “Effects of using different plasmonic metals in metal/dielectric/metal subwavelength waveguides on guided dispersion characteristics,” J. Opt. A, Pure Appl. Opt. 11(7), 075003 (2009). [CrossRef]  

18. M. A. Noginov, L. Gu, J. Livenere, G. Zhu, A. K. Pradhan, R. Mundle, M. Bahoura, Y. A. Barnakov, and V. A. Podolskiy, “Transparent conductive oxides: Plasmonic materials for telecom wavelengths,” Appl. Phys. Lett. 99(2), 021101 (2011). [CrossRef]  

19. P. R. West, S. Ishii, G. V. Naik, N. K. Emani, V. M. Shalaev, and A. Boltasseva, “Searching for better plasmonic materials,” Laser Photonics Rev. 4(6), 795–808 (2010). [CrossRef]  

20. C. Chen, P. Berini, D. Feng, S. Tanev, and V. Tzolov, “Efficient and accurate numerical analysis of multilayer planar optical waveguides in lossy anisotropic media,” Opt. Express 7(8), 260–272 (2000). [CrossRef]   [PubMed]  

21. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).

22. R. H. Ritchie, “Surface plasmons in solids,” Surf. Sci. 34(1), 1–19 (1973). [CrossRef]  

23. R. Hao, E. Cassan, H. Kurt, X. Le Roux, D. Marris-Morini, L. Vivien, H. Wu, Z. Zhou, and X. Zhang, “Novel slow light waveguide with controllable delay-bandwidth product and utra-low dispersion,” Opt. Express 18(6), 5942–5950 (2010). [CrossRef]   [PubMed]  

24. G. Wang, H. Lu, and X. Liu, “Dispersionless slow light in MIM waveguide based on a plasmonic analogue of electromagnetically induced transparency,” Opt. Express 20(19), 20902–20907 (2012). [CrossRef]   [PubMed]  

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

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Schematic of the hetero-MIM plasmonic waveguide.
Fig. 2
Fig. 2 (a) Real part and (b) Imaginary part of the dispersion relations for εm1-εd-εm2 (solid line) waveguide with ωp1 = 1.2eV, ωp2 = 2.03eV, ε = 4, and γ = 0.06eV. Results for εm1-εd-εm1 (dashed line) and εm2-εd-εm2 (dashed-dotted line) waveguides are presented for comparison. The dotted line represents the light line in silicon dioxide.
Fig. 3
Fig. 3 (a) Real part and (b) Imaginary part of the dispersion relations for the εm1-εd-εm2 waveguide in the lossless (dash line) and lossy cases (solid line) with ωp1 = 1.2eV, ωp2 = 2.03eV, and ε = 4 and FOMs (dashed-dotted line) in the lossy case.
Fig. 4
Fig. 4 Dispersion relation (solid line), group index (dashed line), and GVD (dashed-dotted line) in the εm1-εd-εm2 waveguide. The parameters are the same as that in Fig. 2(a).
Fig. 5
Fig. 5 Field distribution profiles for various frequencies in the εm1-εd-εm2 waveguide. The field at ω = 0.62eV in the εm1-εd-εm1 waveguide is also presented.
Fig. 6
Fig. 6 (a) Group index ng and bandwidth Δf versus εd with ωp1 = 1.2eV and ωp2 = 2.03 eV. (b) Group index ng and bandwidth Δf versus Δωp with ωp1 = 1.2eV and εd = 2.25. The zero GVD is kept.

Equations (1)

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

γ ˜ s n ˜ x x s 2 m 11 + γ ˜ c n ˜ x x c 2 m 22 m 21 γ ˜ s γ ˜ c n ˜ x x s 2 n ˜ x x c 2 m 12 = 0 ,
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.