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

Modeling and design methodology for metal-insulator-metal plasmonic Bragg reflectors

Open Access Open Access

Abstract

In this paper, we present a modeling and design methodology based on characteristic impedance for plasmonic waveguides with Metal-Insulator-Metal (MIM) configuration. Finite-Difference Time-Domain (FDTD) simulations indicate that the impedance matching results in negligible reflection at discontinuities in MIM heterostructures. Leveraging the MIM impedance model, we present a general Transfer Matrix Method model for MIM Bragg reflectors and validate our model against FDTD simulations. We show that both periodically stacked dielectric layers of different thickness or different material can achieve the same performance in terms of propagation loss and minimum transmission at the central bandgap frequency in the case of a finite number of periods.

©2008 Optical Society of America

1. Introduction

Photonic band gap (PBG) structures are composed of periodically modulated regions of different refractive indices. They modify propagation properties of electromagnetic waves and forbid their propagation in some frequency intervals in the same way semiconductors affect electrons. Bragg reflectors are in fact 1D PBG structures, where stacks of alternating high- and low-index layers reflect the incident radiation at certain frequencies. Bragg reflectors play a central role in several optoelectronic devices such as filters, modulators, vertical-cavity surface emitting lasers, microcavities, light-emitting diodes and resonant-cavity photodetectors [1, 2, 3].

With the ultimate goal of realization of highly integrated optical components and circuits, the surface plasmon polaritons (SPPs) can be considered as a solution to overcome diffraction limit that restricts down-scaling of photonic devices to a lower-size limit of half the effective wavelength [4]. The SPP field profiles decay exponentially into the neighboring media from their maxima at the interface. In contrast to conventional dielectric components that confine electromagnetic waves to an optically dense core, SPPs are localized at the interface resulting in a subwavelength light-confinement. Low-loss transmission through sharp bends and power dividers has been demonstrated in plasmonic waveguides [5].

Plasmonic PBG components have been used in low threshold SPP-based light-emitting diodes and single-mode surface-plasmon lasers [6, 7, 8]. However, most of the experimentally characterized Bragg reflector structures consist of corrugated metallic strips based on the Insulator-Metal-Insulator (IMI) geometry, which cannot provide true subwavelength light confinement, and therefore, is not suitable for high integration levels [9, 10]. In fact, the Metal-Insulator-Metal (MIM) geometry has been proposed for future on-chip optical communication, which can provide both low-loss propagation and light confinement to lateral dimension of 10% the wavelength [4]. MIMBragg reflectors have been proposed in few numerical studies [11, 12].

In this paper, we present an accurate characteristic impedance forMIMwaveguide structures. Using Finite-Difference Time-Domain (FDTD) simulations, we show that impedance matching can diminish the reflection at discontinuities in MIM based heterostructures. Then, we present a general formulation based on the Transfer Matrix Method (TMM) for MIM Bragg reflectors with subwavelength dielectric layer thickness. Based on the TMM analysis, we investigate the impact of geometric and material parameters on propagation loss and PBG strength.

2. MIM waveguide modeling

In MIM structures [Fig. 1(a) inset], assuming the functional form of exp[i(βz-ωt)] for the field components propagating in the z-direction, the dispersion equation for the fundamental TM mode, where Ey,Bx and Bz are identically zero, is given by [4]

εdkx2+εmkx1coth(ikx1d2)=0,
kx1,22=εd,m(ωc)2k2,
 figure: Fig. 1.

Fig. 1. (a) Real part and (b) imaginary part of effective index (neff) variation as a function of wavelength for d=25 nm, 50 nm, 100 nm and 200 nm, for Ag/SiO 2/Ag (solid line) and Ag/Si/Ag (circled line) waveguides. The refractive indices (n) of SiO 2 and Si are assumed to be 1.46 and 3.4, respectively. A schematic of MIM waveguide is shown in (a) inset.

Download Full Size | PDF

where εd and εm are dielectric functions of dielectric and metallic materials, respectively. We assume silver for the metallic cladding layers, characterized by the Drude model

εm(ω)=ε()ωp2ω(ω+iγ),

where the material-dependent constants ωp=1.38×1016 Hz and γ=2.73×1013 Hz are the bulk plasma and damping frequencies, respectively, and ε(∞)=3.7 [13]. Propagation constant (β) is usually presented by a dimensionless effective index neff=β/k 0 for the guided modes, where k 0 is the vacuum wave-vector. Variation of the effective index values with wavelength is depicted in Fig. 1 for different dielectric materials and thicknesses (t). Figure 1 shows that the frequency dependent effective index associated with an MIM waveguide, highly depends on both t and the dielectric material (n). Therefore, a periodic change in core thickness as well as in dielectric material may result in an effective index modulation needed to realize Bragg scattering around a desired frequency.

For an MIM waveguide with a dielectric layer of refractive index n, Z0 associated with the supported Transverse Magnetic (TM) mode is given as follows

Z=ExdHyw=βdn2ωε0w,

where waveguide width w=1µm is assumed in the y-direction. Note that Eq. (4) is valid since electromagnetic field penetration into the metal exponentially decays from the dielectric-metal interface, and the field profile inside the dielectric region is almost uniform at wavelengths far from the surface plasmon resonance. Equation (4) is the general form of the MIM waveguide characteristic impedance (Z=βdωε0w) derived in [14], which is correct assuming air (n=1).

The structures shown in Fig. 2 are the building blocks for both Thickness-Modulated (ThM) and Index-Modulated (InM) Bragg reflectors, where the thickness and dielectric constant of the middle waveguide is different from those of the left and right sections. Assuming n 1=2, t 1=140 nm and n 2=1, effective index and characteristic impedance at λ=1550 nm become n eff1,1=2.3090+0.0036i and Z1=Z3=30.5 Ω.µm, respectively. Iteratively solving for t 2, so that Z2(t 2)=Z1, we find t 2=60 nm, for which n eff,2=1.3201+0.0033i and Z2(t 2)=29.9 Ω.µm. In order to validate Eq. (4), we simulated the structures shown in Fig. 2(a) using FDTD, where impedance matching condition is satisfied using Eq. (4). Propagation of light through the discontinuities suffers negligibly due to the impedance matching. For the structure shown in Fig. 2(b), β×d is the same for all sections, however, the FDTD simulation demonstrates considerable reflection at the discontinuities. Therefore, n 2 must be considered in the denominator of Eq. (4) for correct calculation of Z.

 figure: Fig. 2.

Fig. 2. FDTD simulation results for a thickness modulated waveguide designed based on (a) our characteristic impedance matching technique provided in Eq. 4 and (b) the characteristic impedance matching technique introduced in [14]. The schematic of both structures with actual physical dimensions, effective indices, and characteristic impedances are shown on top of each FDTD simulation result.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. (a) Schematic of a generalized Bragg reflector with dielectric index and thickness modulations. (b) Bandgap strength variation as a function of thickness for ThM Bragg reflector [t 2=t 1/2, n 1=n 2=1.46] and InM Bragg reflector [n 1=1.46, n 2=3.4, t 1=t 2] from TMM and FDTD simulations. (c) Variation of sin -1{abs[(Z1-Z2)/(Z1+Z2)]} versus t 2.

Download Full Size | PDF

3. MIM Bragg reflector formulation

An MIM Bragg reflector is shown in Fig. 3(a). Note that in ThM and InM Bragg reflectors, n 1=n 2,t 1t 2 and n 1n 2, t 1=t 2, respectively. Field analysis of the discontinuity problems (physical dimensions or material discontinuity) can be simplified by modal analysis technique. Considering the dominant propagating mode inside the waveguide, we can find the electric and magnetic fields, which are correlated together by characteristic impedance defined in Eq. 4. The dominant mode in the MIM waveguide is the fundamental TM mode, which has insignificant Ez component. Therefore, this mode can be considered as a quasi-TEMmode [14]. Electromagnetic wave propagation inside an MIM Bragg reflector can be modeled based on the boundary conditions for transverse fields at the discontinuities resulting in a transfer matrix (A) representing the reflection and transmission at the discontinuities, which is defined as

A11=12ej(βrβl)X+Zr2Zlej(βrβl)X,A12=12ej(βr+βl)XZr2Zlej(βr+βl)X
A21=12ej(βr+βl)XZr2Zlej(βr+βl)X,A22=12ej(βrβl)X+Zr2Zlej(βrβl)X

where X is the position of the discontinuity. l and r refer to the left side and the right side of the discontinuity in the heterostructure, respectively, as shown in Fig. 3(a). The total transmission spectra can be calculated by ΠiAi. Note that this characteristic impedance formulation is general and is applicable for both ThM and InM Bragg reflectors. Figure 3(b) compares the bandgap strength Δω/ωb calculated using Eq. 5 and FDTD simulations for two ThM and InM Bragg reflectors, where Δω and ωb are the FWHM of the reflection peak at the central gap frequency and the central gap frequency, respectively.

The first band gap Δω/ωbsin -1[(n eff,2-n eff,1)/(n eff,2+n eff,1)] in a 1D dielectric photonic crystal depends on the effective index contrast, where n eff,1 and n eff,2 are the effective indices associated with the two alternately stacked waveguides and the number of periods of the alternating waveguides (N) is assumed to be infinite [15]. Figure 3(c) demonstrates the variation of sin -1[|(Z1-Z2)/(Z1+Z2)|]. Interestingly, The bandgap strength in the case of the ThM and InM Bragg reflectors increases about 35% and 0%, as t 2 increases from 40 nm to 200 nm in both Figures 3(b) and (c). Therefore, the MIM characteristic impedance defined in Eq. 4 can be utilized to predict the bandgap strength of MIM Bragg reflectors.

4. Design of MIM Bragg reflectors

To optimize the material and geometric parameters of MIM Bragg reflector structures, the design problem must be formulated to maximize the objectives and meet the constraints while effectively exploiting the design space based on the analysis presented in Section 3. In MIM waveguides decreasing t or increasing n both result in higher effective index value, which can be used to create effective index contrast and wide bandgaps in MIM Bragg reflector, but they also increase the propagation loss inside the structure, as shown in Fig. 1(b). In order to facilitate the design optimization, we have developed a dimensionless figure of merit (FOM), the propagation length over the Bragg reflector length

F=d1+d22k0[d1Im(neff,1)+d2Im(neff,2)]×1N(d1+d2)=12Nk0[d1Im(neff,1)+d2Im(neff,2)],

An important characteristic for Bragg reflectors is to find a structure with acceptable reflectivity and the lowest loss, which can be formally posed as

MaximizeF(N,n1,n2,t1,t2,d1,d2)
SubjecttoR(N,n1,n2,t1,t2,d1,d2)>R0

Note that we did not include bandgap strength in Eq. 6, because based on the application, wide or narrow bandgap may be required. Therefore, the maximum peak reflectivity in the bandgap is more appropriate for the definition of the figure of merit.

Higher refractive index materials result in higher propagation loss inMIMwaveguides. However, the characteristic impedance contrast (CZ=Z1Z2Z1+Z2) does not change with n in the case of ThM Bragg reflectors, except at high frequencies, where Real{εm(ω)}≈-εd. Therefore, in order to increase F for ThM Bragg reflectors, low refractive index dielectric should be used.

 figure: Fig. 4.

Fig. 4. FOM variation as a function of number of periods and (a) n 2 for InM Bragg reflector (n 1=1.46, t 1=t 2=100 nm) and (b) t 2 for ThM Bragg reflector (n 1=n 2=1.46, t 1=100 nm) for fixed central wavelength of 1550 nm. Transmission (TMM and FDTD) and Reflection (TMM) spectra versus wavelength for optimum (c) InM (n 2=2.6) and (d) ThM (t 2=50 nm) Bragg reflectors with N=6.

Download Full Size | PDF

Figures 4(a) and (b) demonstrate the variations of F as N and the characteristic impedance contrast change in InM and ThM Bragg reflectors for a fixed central wavelength λb=1550 nm. As shown in Figures 4(a) and (b), for a fixed C Z, the number of periods has an indirect relationship on F due to the increase in the propagation loss. In both ThM and InM Bragg reflectors, the maximum achievable F increases with the characteristic impedance contrast. In the InM Bragg reflector, when the minimum reflection criteria is satisfied, a change in C Z for a fixed N results in an insignificant change in F. Therefore, the Bragg performance is not degraded with the discrete values of available dielectric constants. However, in the ThM Bragg reflector, F rapidly decreases with decreasing t 2 (increasing C Z), which is due to increasing loss in narrower MIM waveguides as depicted in Fig. 1(b). The transmission and reflection spectra of the optimum designs with N=6 and λb=1550 nm are shown in Figures 4(c) and (d) for InM (F≈13) and ThM (F≈11) Bragg reflectors, respectively. The two Bragg reflector structures demonstrate almost the same performance. In addition, the TMM calculations based on MIM waveguide characteristic impedance and FDTD simulations are in good agreement.

5. Conclusion

In this paper, we presented a characteristic impedance modeling for plasmonic waveguides with Metal-Insulator-Metal (MIM) configuration. Leveraging the MIM impedance model, we presented a general Transfer Matrix Method model for MIM Bragg reflectors and validated our model against FDTD simulations. The results indicate that both periodically stacked dielectric layers of different thickness or different material can achieve the same performance in terms of propagation loss and minimum transmission.

References and links

1. M. Rattier, H. Benisty, R. Stanley, J.-F. Carlin, R. Houdre, U. Oesterle, C. Smith, C. Weisbuch, and T. Krauss, “Toward ultrahigh-efficiency aluminum oxide microcavitylight-emitting diodes: guided mode extraction by photonic crystals,” IEEE J. Sel. Top. Quantum. Electron. 8, 238–247 (2002). [CrossRef]  

2. J. Wierer, D. Kellogg, and N. Holonyak, “Tunnel contact junction native-oxide aperture and mirror vertical-cavity surface-emitting lasers and resonant-cavity light-emitting diodes,” Appl. Phys. Lett. 74, 926–928 (1999). [CrossRef]  

3. D. Zhao, K. Chan, Y. Liu, L. Zhang, and I. Bennion, “Wavelength-switched optical pulse generation in a fiber ring laser with a Fabry-Perot semiconductor modulator and a sampled fiber Bragg grating,” IEEE Photon. Technol. Lett. 13, 191–193 (2001). [CrossRef]  

4. J. Dionne, L. Sweatlock, and H. Atwater, “Plasmon slot waveguides: Towards chip-scale propagation with subwavelength-scale localization,” Phys. Rev. B 73, 035407 (2006). [CrossRef]  

5. G. Veronis and S. Fan, “Subwavelength plasmonic waveguide structures based on slots in thin metal films,” Proceed. SPIE 6123 (2006). [CrossRef]  

6. P. Hobson, S. Wedge, J. Wasey, I. Sage, and W. Barnes, “Surface plasmon mediated emission from organic light-emitting diodes,” Adv. Mater. 14, 1393–1396 (2005). [CrossRef]  

7. A. Tredicucci, C. Gmachl, F. Capasso, A. Hutchinson, D. Sivco, and A. Cho, “Single-mode surface-plasmon laser,” Appl. Phys. Lett. 76, 2164–2166 (2000). [CrossRef]  

8. C. Sirtori, J. Faist, F. Capasso, D. Sivco, A. Hutchinson, and A. Cho, “Quantum cascade laser with plasmonenhanced waveguide operating at 8.4 u m wavelength,” Appl. Phys. Lett. 66, 3242–3244 (1995). [CrossRef]  

9. H. Ditlbacher, J. Krenn, G. Schider, A. Leitner, and F. Aussenegg, “Two-dimensional optics with surface plasmon polaritons,” Appl. Phys. Lett. 81, 1762–1764 (2002). [CrossRef]  

10. J.-C. Weeber, Y. Lacroute, A. Dereux, E. Devaux, T. Ebbesen, C. Girard, M. U. Gonzalez, and A.-L. Baudrion, “Near-field characterization of Bragg mirrors engraved in surface plasmon waveguides,” Phys. Rev. B 70, 235406 (2004). [CrossRef]  

11. B. Wang and G. P. Wang, “Plasmon Bragg reflectors and nanocavities on flat metallic surfaces,” Appl. Phys. Lett. 87, 013107 (2005). [CrossRef]  

12. A. Hosseini and Y. Massoud, “A low-loss metal-insulator-metal plasmonic bragg reflector,” Opt. Express 14, 318–323 (2006). [CrossRef]  

13. Z. Han, E. Forsberg, and S. He, “Surface plasmon Bragg gratings formed in metal-insulator-metal waveguides,” IEEE Photon. Technol. Lett. 19, 91–93 (2007). [CrossRef]  

14. G. Veronis and S. Fan, “Bends and splitters in metal-dielectric-metal subwavelength plasmonic waveguides,” Appl. Phys. Lett. 87, 131102 (2005). [CrossRef]  

15. I. Nusinsky and A. A. Hardy, “Band-gap analysis of one-dimensional photonic crystals and conditions for gap closing,” Phys. Rev. B 73, 125104 (2006). [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 (4)

Fig. 1.
Fig. 1. (a) Real part and (b) imaginary part of effective index (neff) variation as a function of wavelength for d=25 nm, 50 nm, 100 nm and 200 nm, for Ag/SiO 2/Ag (solid line) and Ag/Si/Ag (circled line) waveguides. The refractive indices (n) of SiO 2 and Si are assumed to be 1.46 and 3.4, respectively. A schematic of MIM waveguide is shown in (a) inset.
Fig. 2.
Fig. 2. FDTD simulation results for a thickness modulated waveguide designed based on (a) our characteristic impedance matching technique provided in Eq. 4 and (b) the characteristic impedance matching technique introduced in [14]. The schematic of both structures with actual physical dimensions, effective indices, and characteristic impedances are shown on top of each FDTD simulation result.
Fig. 3.
Fig. 3. (a) Schematic of a generalized Bragg reflector with dielectric index and thickness modulations. (b) Bandgap strength variation as a function of thickness for ThM Bragg reflector [t 2=t 1/2, n 1=n 2=1.46] and InM Bragg reflector [n 1=1.46, n 2=3.4, t 1=t 2] from TMM and FDTD simulations. (c) Variation of sin -1{abs[(Z1-Z2)/(Z1+Z2)]} versus t 2.
Fig. 4.
Fig. 4. FOM variation as a function of number of periods and (a) n 2 for InM Bragg reflector (n 1=1.46, t 1=t 2=100 nm) and (b) t 2 for ThM Bragg reflector (n 1=n 2=1.46, t 1=100 nm) for fixed central wavelength of 1550 nm. Transmission (TMM and FDTD) and Reflection (TMM) spectra versus wavelength for optimum (c) InM (n 2=2.6) and (d) ThM (t 2=50 nm) Bragg reflectors with N=6.

Equations (9)

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

ε d k x 2 + ε m k x 1 coth ( i k x 1 d 2 ) = 0 ,
k x 1 , 2 2 = ε d , m ( ω c ) 2 k 2 ,
ε m ( ω ) = ε ( ) ω p 2 ω ( ω + i γ ) ,
Z = E x d H y w = β d n 2 ω ε 0 w ,
A 11 = 1 2 e j ( β r β l ) X + Z r 2 Z l e j ( β r β l ) X , A 12 = 1 2 e j ( β r + β l ) X Z r 2 Z l e j ( β r + β l ) X
A 21 = 1 2 e j ( β r + β l ) X Z r 2 Z l e j ( β r + β l ) X , A 22 = 1 2 e j ( β r β l ) X + Z r 2 Z l e j ( β r β l ) X
F = d 1 + d 2 2 k 0 [ d 1 I m ( n eff , 1 ) + d 2 I m ( n eff , 2 ) ] × 1 N ( d 1 + d 2 ) = 1 2 N k 0 [ d 1 I m ( n eff , 1 ) + d 2 I m ( n eff , 2 ) ] ,
Maximize F ( N , n 1 , n 2 , t 1 , t 2 , d 1 , d 2 )
Subject to R ( N , n 1 , n 2 , t 1 , t 2 , d 1 , d 2 ) > R 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.