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

Single-mode plasmonic waveguiding properties of metal nanowires with dielectric substrates

Open Access Open Access

Abstract

Single-mode plasmonic waveguiding properties of metal nanowires with dielectric substrates are investigated using a finite-element method. Au and Ag are selected as plasmonic materials for nanowire waveguides with diameters down to 5-nm-level. Typical dielectric materials with relatively low to high refractive indices, including magnesium fluoride (MgF2), silica (SiO2), indium tin oxide (ITO) and titanium dioxide (TiO2), are used as supporting substrates. Basic waveguiding properties, including propagation constants, power distributions, effective mode areas, propagation distances and losses are obtained at the typical plasmonic resonance wavelength of 660 nm. Compared to that of a freestanding nanowire, the mode area of a substrate-supported nanowire could be much smaller while maintaining an acceptable propagation length. For example, the mode area and propagation length of a 100-nm-diameter Ag nanowire with a MgF2 substrate are about 0.004 μm2 and 3.4 μm, respectively. The dependences of waveguiding properties on geometric and material parameters of the nanowire-substrate system are also provided. Our results may provide valuable references for waveguiding dielectric-supported metal nanowires for practical applications.

©2012 Optical Society of America

1. Introduction

Owing to their capability of manipulating electromagnetic fields on the deep-subwavelength scale by converting light into surface plasmon polaritons (SPPs), plasmonic metal nanostructures have inspired a variety of potentials ranging from ultra-compact optoelectronic circuits [1,2], optical sensors [3,4] to quantum electrodynamics research [58]. So far, various types of metal nanostructures have been proposed for guiding SPPs [915], among which Au and Ag nanowires are typical one-dimension waveguide structures with relatively low losses at visible and near-infrared spectral ranges. In the past years, waveguiding properties of metal nanowires have been extensively investigated theoretically [9,1619], but mostly on free-standing nanowires with symmetric dielectric surroundings. In experimental cases [2022], a supporting substrate is usually indispensable for nanowire manipulation and characterization, and therefore waveguiding properties of nanowires with proper substrates are of great importance for practical applications.

Recently, several groups reported the plasmon modes of Ag nanowires coupled with a silica or a silicon substrate [23,24], but mostly focused on nanowire-substrate system with a certain gap, and the propagation lengths obtained in Ref. [23] showed relatively large discrepancy with both theoretical and experimental results [2527]. More recent works compared the propagation lengths of Au and Ag nanowires with glass or indium tin oxide (ITO) substrates [28,29], but the power distribution, effective mode area have not been studied. In contrast to the well-studied free-standing nanowires, waveguiding properties of substrate-supported SPP nanowires have not been adequately investigated.

In this paper, we investigate waveguiding properties of metal nanowires with dielectric substrates using a Comsol Multiphysics finite element method. Aiming for operating SPP nanowires with tight confinement, here we focus on nanowires with subwavelength diameters down to 5-nm-level, and study only the lowest order mode in proposed nanowires under the following considerations: (1) single-mode operation is favorable or required in most practical applications; (2) the fundamental mode plays the central role and dominates the propagation properties [9]; (3) experimentally, although both of the m = 0 and 1 modes have no cutoff diameter, the m = 1 mode is difficult to excite [30]. Also, single-mode SPP waveguiding nanowires can be realized by controlling the incident polarization [31].

Using waveguiding systems with Au and Ag nanowires and dielectric substrates including magnesium fluoride (MgF2), silica (SiO2), indium tin oxide (ITO) and titanium dioxide (TiO2), we obtained single-mode waveguiding properties of these nanowires including propagation constants, power distributions, effective mode areas, propagation distances and losses at the typical plasmonic resonance wavelength of 660 nm. The propagation distances and losses reported here agree well with experimental results [22,27,29].

2. Basic model

The mathematical model for our numerical simulation is illustrated in Fig. 1 , in which an infinite long and straight nanowire with a diameter of D is placed on a dielectric substrate.

 figure: Fig. 1

Fig. 1 Mathematic model for simulation of a nanowire-substrate waveguiding system.

Download Full Size | PDF

The propagation constant (β) and propagation length (Lm) of the nanowire are defined as [32]

β=Re(β)+iIm(β),
Lm=12Im(β).

The propagation loss (α) is inversely proportional to Lm as

α=10log(1/e)Lm4.343Lm.

The effective mode area (Am) is defined as [3337]

Am=Wmmax{W(r)},
where Wm is the total mode energy and W(r) is the energy density (per unit length flowed along the direction of propagation). For dispersive and lossy materials, the W(r) inside can be calculated as [3337]
W(r)=12(d(ε(r)ω)dω|E(r)|2+μ0|H(r)|2),
where ε(r) is the complex dielectric function. At the wavelength of 660 nm used in this work, the dielectric constants of the Au (εAu) and Ag (εAg) were assumed to be εAu = −13.6815-1.0356i [38] and εAg = −17.6986-1.1786i [39], and d(ε(r)ω) / is obtained by analytical models from Refs [40] and [41], respectively. For simplicity, we assume that the nanowire has a perfectly smooth sidewall without surface-roughness-induced scattering. The refractive index of air is assumed to be 1.0, and the indices of four substrates chosen in this work at 660 nm wavelength are listed in Table 1 .

Tables Icon

Table 1. Refractive indices of substrates at 660 nm wavelength

3. Modal profiles and power distributions

Based on the model set up in Section 2, we investigate waveguiding properties of the nanowire-substrate system using a Comsol Multiphysics finite element method. The computational domain is discretized into a triangular mesh with an element size of one tenth of the nanowire diameter (e.g., 10 nm for a 100 nm diameter nanowire), terminated by perfectly matched layer (PML) boundaries.

Calculated modal profiles in terms of energy density distributions for Au nanowires guiding 660-nm wavelength light are shown in Fig. 2 , in which Au nanowires with different diameters (50, 100, 200 nm) and substrates (air, MgF2, SiO2, ITO and TiO2) are considered. While the free-standing 50-nm-diameter Au nanowire shows a symmetric mode profile (Fig. 2(a)), the introduction of the substrates breaks the symmetry and confines the energy to the interface between the nanowire and the substrate (Figs. 2(b)-2(d)), as has been reported previously [42]. Moreover, the fractional energy confined around the interface increases with the increasing nanowire diameter (e.g., modal profiles in Fig. 2(d) vs. Fig. 2(c)). The energy density enhancement can be explained by polarized substrate (insets of Fig. 2(a) and Fig. 2(b)), which is similar to nanoparticle-substrate system [43,44]. Meanwhile, for Au nanowires with the same 100 nm diameter (Figs. 2(e)-2(h)), the energy confinement around the interface increases with the increasing substrate indices (e.g., modal profiles in Fig. 2(h) vs. Fig. 2(g)), which can be attributed to higher polarizability in the substrate with a higher refractive index.

 figure: Fig. 2

Fig. 2 Energy density distribution on the cross section of Au nanowires. (a) D = 50 nm, no substrate; inset, schematic illustration of the polarized fields. (b) D = 50 nm, MgF2 substrate; inset, schematic illustration of the polarized fields. (c) D = 100 nm, MgF2 substrate; (d) D = 200 nm, MgF2 substrate; (e) D = 100 nm, no substrate; (f) D = 100 nm, SiO2 substrate; (g) D = 100 nm, ITO substrate; (h) D = 100 nm, TiO2 substrate. The wavelength of light used here is 660 nm.

Download Full Size | PDF

The propagation constants (β) of Au and Ag nanowires are obtained from the eigenvalue equations by the COMSOL software. Figure 3 and Fig. 4 give the calculated Re(β) and Im(β) of Au and Ag nanowires, respectively. It shows that, at 660-nm wavelength, Re(β) is approximately a constant when D is larger than about 100 nm; however, when D goes below 100 nm, Re(β) starts to increase with the decreasing D. It is noticed that, when D is very small (e.g., kD << 1, k represents the wavevector), Re(β) exhibits a 1/D behavior, agrees well with previous works [30,45]. The small D case can be regarded as the quasi-static configuration of electric field and associated charge density wave [30,46], in which the substrate has little impact on the plasmon mode. For comparison, Fig. 5 gives calculated modal profiles of SiO2-substrate-support Ag nanowires with diameters of 10, 20, 50 and 100 nm, clearly showing that the substrate-induced impact on the modal profiles decreases with the nanowire diameters: the modal profiles are obviously asymmetric in 50 and 100-nm nanowires, while show better symmetries in 10 and 20-nm nanowire. Also, Re(β) is affected by refractive index of the substrate (nsub): for the same nanowire, the higher the nsub, the larger the Re(β), which can be attributed to the lower light propagation velocity in higher-nsub medium.

 figure: Fig. 3

Fig. 3 Numerical solutions of the real part of propagation constants (Re(β)) of (a) Au nanowires, and (b) Ag nanowires. Insets, Re(β) of the nanowire diameter larger than 100nm. The wavelength of light used here is 660 nm.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Numerical solutions of the imaginary part of propagation constants (Im(β)) of (a) Au nanowires, and (b) Ag nanowires. Insets, Im(β) of the nanowire diameter larger than 100nm. The wavelength of light used here is 660 nm.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Energy density distribution on the cross section of Ag nanowires placed on SiO2. (a) D = 10 nm; (b) D = 20 nm; (c) D = 50 nm; (d) D = 100 nm. The wavelength of light used here is 660 nm.

Download Full Size | PDF

To quantify the power distribution of the fundamental modes inside the nanowire core and the substrates, we define the fractional power inside the core (ƞc) and the substrate (ƞs) as

ηc=coreW(r)dAWm,
ηs=substrateW(r)dAWm,
where dA = rdr, is the cross-section surface element of the nanowire in the cylindrical coordinates. Figure 6 and Fig. 7 are the D-dependent ƞc and ƞs of Au and Ag nanowires, respectively. It shows that, ƞc increases monotonically with decreasing D. When D is very small (e.g., <5 nm for Au nanowire and < 3 nm for Ag nanowire), ƞc approach a constant value of about 72% for Au nanowire and 52% for Ag nanowire. In contrast, with increasing D, ƞs increases monotonically resulting in higher fractional energy confined inside the substrate.

 figure: Fig. 6

Fig. 6 Fractional power of the plasmon mode of Au nanowires inside the (a) core (ƞc), and (b) substrate (ƞs). Inset of (a), a close-up view of the ƞc for Au nanowires with MgF2 and SiO2 substrates, respectively. The wavelength of light used here is 660 nm.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Fractional power of the plasmon mode of Ag nanowires inside the (a) core (ƞc), and (b) substrate (ƞs). Inset of (a), a close-up view of the ƞc for Ag nanowires with MgF2 and SiO2 substrates, respectively. The wavelength of light used here is 660 nm.

Download Full Size | PDF

4. Optical confinement and effective mode areas

Owing to their possibility of confining propagation light below the diffraction limits, optical confinement (or equivalently the effective mode area) is one of the most important merits of the plasmonic waveguides. To quantify the confinement, here we calculate the effective mode area (Am) of nanowire waveguides with Eq. (4). Figures 8(a and b) show calculated mode areas of Au and Ag nanowires with dielectric substrates, respectively. For reference, mode areas of free-standing Au and Ag nanowires (without substrate) are also given (black line). It shows that, unlike that of a tightly confined dielectric waveguides (e.g., a silicon nanowire), in which Am increases with decreasing D, Am of the plasmonic nanowire decreases monotonically with D making it possible to realize ultra-tightly confined waveguiding modes. For example, Am of a 50-nm-diameter Au nanowire is about 0.0026 μm2, which is about 0.6% of λ2; for comparison, the minimum Am of a silicon nanowire is about 3% of λ2 [47]. Moreover, compared to that of the free-standing nanowire, Am of the substrate-supported nanowire is significantly reduced. For example, for a 100-nm-diameter free-standing Ag nanowire, Am is about 0.021μm2. When it is supported by a SiO2 substrate, Am reduces to 0.003 μm2, much smaller than that of the free-standing one. It is also noticed that, for the same nanowire, the higher the index of the substrate, the smaller the Am. Besides, with the same diameter and substrate, Am of a Au nanowire is always smaller than a Ag nanowire. In addition, for reference, Fig. 9 gives the normalized amplitude of the electric filed along y direction (see inset) of a 200-nm-diameter Au nanowires. The electric filed is much stronger at the substrate side than that at the opposite interface. Also, the higher the index of the substrate, the higher peak intensity of the electric field.

 figure: Fig. 8

Fig. 8 Effective mode area (Am) of (a) Au nanowires, and (b) Ag nanowires. The wavelength of light used here is 660 nm.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Normalized amplitude of the electric filed along y direction of a 200-nm-diameter Au nanowire. The wavelength of light used here is 660 nm. Inset, coordinates on the cross section of the nanowire.

Download Full Size | PDF

5. Propagation lengths and waveguiding losses

Around the surface plasmonic resonance frequency, a waveguiding metal nanowire with subwavelength confinement usually suffers from high ohmic losses. By substituting the calculated Im(β) in Fig. 4 into the definition in Eq. (2), we obtained the propagation length (Lm) shown in Figs. 10(a) and 10(b), in which the propagation loss (α) is obtained by Eq. (3). It shows that, Lm increases monotonically with the increasing diameters (D), reflecting the trade-off relations between the confinement and the loss of the plasmonic waveguide. For the Au nanowire, Lm of the freestanding nanowire is always larger than those with the substrate; while for the Ag nanowire, this is only true when D is smaller than 120 nm. Moreover, Fig. 11 gives nsub-dependent Lm of Au and Ag nanowires with diameters of 100 and 200 nm, respectively. It shows that, for the Au nanowire, Lm decreases monotonically with nsub increasing from MgF2, SiO2, ITO to TiO2 for both the 100- and 200-nm diameter nanowires. For the Ag nanowire, when D = 100 nm, Lm decreases monotonically with nsub increasing from MgF2, SiO2, ITO to TiO2; however, when D = 200 nm, Lm increases with increasing nsub (from MgF2 to SiO2) until 1.7, and then drops down with nsub increasing from ITO to TiO2, which agrees well with those shown in Fig. 10(b).

 figure: Fig. 10

Fig. 10 Propagation lengths (Lm) and losses (α) of (a) Au nanowires, and (b) Ag nanowires. The wavelength of light used here is 660 nm.

Download Full Size | PDF

 figure: Fig. 11

Fig. 11 Substrate index (nsub)-dependent propagation length (Lm) of the plasmon mode for (a) Au nanowires, and (b) Ag nanowires. The wavelength of light used here is 660 nm.

Download Full Size | PDF

Figure 10(b) also shows that, for Ag nanowires with diameters larger than 120 nm, the substrate-supported nanowire exhibits a larger Lm (or equivalently lower loss) than the free-standing one, suggesting a possible explanation for the larger experimentally measured Lm of Ag nanowires (with substrates) [27] than that (without substrate) obtained by numerical calculation [26]. To verify the validity of the simulation results, we have also compared the calculated Lm of both Au and Ag nanowires with experimental results reported previously [22,27,29]. For example, Ma et al. reported a propagation length of 10.56 μm in a 260 nm diameter Ag nanowire supported by SiO2 substrates at 633 nm wavelength [27], agrees well with our calculation result of about 10.86 μm at 660 nm wavelength.

6. Conclusions

In summary, we have investigated the single-mode waveguiding properties of Au and Ag nanowires with typical dielectric substrates. Using a FEM method, propagation constants, power distributions, effective mode areas, propagation lengths and losses are obtained numerically. It shows that, for a substrate-supported nanowire, the larger the nanowire diameter, the more asymmetric the plasmon mode profile. Also, with increasing substrate index, the fractional energy confined around the interface increases, and the effective mode area decreases. Meanwhile, compared to a free-standing nanowire, the nanowire-substrate structure can simultaneously offer a much tighter confinement and a relatively large propagation length, when the parameters of the nanowire and the substrate are properly chosen. For example, for a 100-nm-diameter free-standing Ag nanowire, Am and Lm are 0.021 μm2 and 3.6 μm, respectively. When it is supported by a MgF2 substrate, Am and Lm are 0.004 μm2 and 3.4 μm, respectively. It should also be mentioned that, in our calculation, the nanowires are assumed to have smooth surfaces and uniform diameters, while real nanowires are usually not ideally uniform, and the dielectric constants may be slightly different from those used in this work. Compared with previous experimental results, waveguiding properties we obtained here are reasonable and meaningful. Particularly, for experimental study or practical applications, plasmonic nanowires are usually supported by a certain substrate (usually dielectric) and operated in single mode, therefore, the single-mode plasmonic waveguiding properties of a metal nanowire shown here could be very helpful for design, selection and utilization of plasmonic waveguides and building blocks for a variety of applications from nanophotonic/plasmonic passive waveguiding circuits to active devices.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Nos. 61036012, 61108048) and Fundamental Research Funds for the Central Universities.

References and links

1. E. Ozbay, “Plasmonics: merging photonics and electronics at nanoscale dimensions,” Science 311(5758), 189–193 (2006). [CrossRef]   [PubMed]  

2. X. Guo, M. Qiu, J. M. Bao, B. J. Wiley, Q. Yang, X. N. Zhang, Y. G. Ma, H. K. Yu, and L. M. Tong, “Direct coupling of plasmonic and photonic nanowires for hybrid nanophotonic components and circuits,” Nano Lett. 9(12), 4515–4519 (2009). [CrossRef]   [PubMed]  

3. S. Lal, S. Link, and N. J. Halas, “Nano-optics from sensing to waveguiding,” Nat. Photonics 1(11), 641–648 (2007). [CrossRef]  

4. E. Hendry, T. Carpy, J. Johnston, M. Popland, R. V. Mikhaylovskiy, A. J. Lapthorn, S. M. Kelly, L. D. Barron, N. Gadegaard, and M. Kadodwala, “Ultrasensitive detection and characterization of biomolecules using superchiral fields,” Nat. Nanotechnol. 5(11), 783–787 (2010). [CrossRef]   [PubMed]  

5. A. V. Akimov, A. Mukherjee, C. L. Yu, D. E. Chang, A. S. Zibrov, P. R. Hemmer, H. Park, and M. D. Lukin, “Generation of single optical plasmons in metallic nanowires coupled to quantum dots,” Nature 450(7168), 402–406 (2007). [CrossRef]   [PubMed]  

6. A. Cuche, O. Mollet, A. Drezet, and S. Huant, “Deterministic quantum plasmonics,” Nano Lett. 10(11), 4566–4570 (2010). [CrossRef]   [PubMed]  

7. A. Ridolfo, O. Di Stefano, N. Fina, R. Saija, and S. Savasta, “Quantum plasmonics with quantum dot-metal nanoparticle molecules: influence of the Fano effect on photon statistics,” Phys. Rev. Lett. 105(26), 263601 (2010). [CrossRef]   [PubMed]  

8. J. Zuloaga, E. Prodan, and P. Nordlander, “Quantum plasmonics: optical properties and tunability of metallic nanorods,” ACS Nano 4(9), 5269–5276 (2010). [CrossRef]   [PubMed]  

9. J. Takahara, S. Yamagishi, H. Taki, A. Morimoto, and T. Kobayashi, “Guiding of a one-dimensional optical beam with nanometer diameter,” Opt. Lett. 22(7), 475–477 (1997). [CrossRef]   [PubMed]  

10. R. M. Dickson and L. A. Lyon, “Unidirectional plasmon propagation in metallic nanowires,” J. Phys. Chem. B 104(26), 6095–6098 (2000). [CrossRef]  

11. S. A. Maier, P. G. Kik, H. A. Atwater, S. Meltzer, E. Harel, B. E. Koel, and A. A. Requicha, “Local detection of electromagnetic energy transport below the diffraction limit in metal nanoparticle plasmon waveguides,” Nat. Mater. 2(4), 229–232 (2003). [CrossRef]   [PubMed]  

12. D. F. P. Pile, T. Ogawa, D. K. Gramotnev, T. Okamoto, M. Haraguchi, M. Fukui, and S. Matsuo, “Theoretical and experimental investigation of strongly localized plasmons on triangular metal wedges for subwavelength waveguiding,” Appl. Phys. Lett. 87(6), 061106 (2005). [CrossRef]  

13. S. I. Bozhevolnyi, V. S. Volkov, E. Devaux, and T. W. Ebbesen, “Channel plasmon-polariton guiding by subwavelength metal grooves,” Phys. Rev. Lett. 95(4), 046802 (2005). [CrossRef]   [PubMed]  

14. J. Jung, T. Søndergaard, and S. I. Bozhevolnyi, “Theoretical analysis of square surface plasmon-polariton waveguides for long-range polarization-independent waveguiding,” Phys. Rev. B 76(3), 035434 (2007). [CrossRef]  

15. D. Solis Jr, W. S. Chang, B. P. Khanal, K. Bao, P. Nordlander, E. R. Zubarev, and S. Link, “Bleach-imaged plasmon propagation (BlIPP) in single gold nanowires,” Nano Lett. 10(9), 3482–3485 (2010). [CrossRef]   [PubMed]  

16. C. A. Pfeiffer, E. N. Economou, and K. L. Ngai, “Surface polaritons in a circularly cylindrical interface: Surface plasmons,” Phys. Rev. B 10(8), 3038–3051 (1974). [CrossRef]  

17. H. Khosravi, D. R. Tilley, and R. Loudon, “Surface polaritons in cylindrical optical fibers,” J. Opt. Soc. Am. A 8(1), 112–122 (1991). [CrossRef]  

18. L. Novotny and C. Hafner, “Light propagation in a cylindrical waveguide with a complex, metallic, dielectric function,” Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Topics 50(5), 4094–4106 (1994). [CrossRef]   [PubMed]  

19. U. Schröter and A. Dereux, “Surface plasmon polaritons on metal cylinders with dielectric core,” Phys. Rev. B 64(12), 125420 (2001). [CrossRef]  

20. C. H. Dong, X. F. Ren, R. Yang, J. Y. Duan, J. G. Guan, G. C. Guo, and G. P. Guo, “Coupling of light from an optical fiber taper into silver nanowires,” Appl. Phys. Lett. 95(22), 221109 (2009). [CrossRef]  

21. R. X. Yan, P. Pausauskie, J. X. Huang, and P. D. Yang, “Direct photonic-plasmonic coupling and routing in single nanowires,” Proc. Natl. Acad. Sci. U.S.A. 106(50), 21045–21050 (2009). [CrossRef]   [PubMed]  

22. T. Shegai, Y. Z. Huang, H. X. Xu, and M. Kall, “Coloring fluorescence emission with silver nanowires,” Appl. Phys. Lett. 96(10), 103114 (2010). [CrossRef]  

23. C. L. Zou, F. W. Sun, Y. F. Xiao, C. H. Dong, X. D. Chen, J. M. Cui, Q. Gong, Z. F. Han, and G. C. Guo, “Plasmon modes of silver nanowire on a silica substrate,” Appl. Phys. Lett. 97(18), 183102 (2010). [CrossRef]  

24. Z. P. Li, K. Bao, Y. R. Fang, Z. Q. Guan, N. J. Halas, P. Nordlander, and H. X. Xu, “Effect of a proximal substrate on plasmon propagation in silver nanowires,” Phys. Rev. B 82(24), 241402 (2010). [CrossRef]  

25. H. Ditlbacher, A. Hohenau, D. Wagner, U. Kreibig, M. Rogers, F. Hofer, F. R. Aussenegg, and J. R. Krenn, “Silver nanowires as surface plasmon resonators,” Phys. Rev. Lett. 95(25), 257403 (2005). [CrossRef]   [PubMed]  

26. X. W. Chen, V. Sandoghdar, and M. Agio, “Highly efficient interfacing of guided plasmons and photons in nanowires,” Nano Lett. 9(11), 3756–3761 (2009). [CrossRef]   [PubMed]  

27. Y. G. Ma, X. Y. Li, H. K. Yu, L. M. Tong, Y. Gu, and Q. H. Gong, “Direct measurement of propagation losses in silver nanowires,” Opt. Lett. 35(8), 1160–1162 (2010). [CrossRef]   [PubMed]  

28. B. Wild, L. Cao, Y. Sun, B. P. Khanal, E. R. Zubarev, S. K. Gray, N. F. Scherer, and M. Pelton, “Propagation lengths and group velocities of plasmons in chemically synthesized gold and silver nanowires,” ACS Nano 6(1), 472–482 (2012). [CrossRef]   [PubMed]  

29. P. Kusar, C. Gruber, A. Hohenau, and J. R. Krenn, “Measurement and reduction of damping in plasmonic nanowires,” Nano Lett. 12(2), 661–665 (2012). [CrossRef]   [PubMed]  

30. D. E. Chang, A. S. Sorensen, P. R. Hemmer, and M. D. Lukin, “Strong coupling of single emitters to surface plasmons,” Phys. Rev. B 76(3), 035420 (2007). [CrossRef]  

31. Z. P. Li, K. Bao, Y. R. Fang, Y. Z. Huang, P. Nordlander, and H. X. Xu, “Correlation between Incident and Emission Polarization in Nanowire Surface Plasmon Waveguides,” Nano Lett. 10(5), 1831–1835 (2010). [CrossRef]   [PubMed]  

32. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, New York, 2007).

33. R. F. Oulton, V. J. Sorger, D. A. Genov, D. F. P. Pile, and X. Zhang, “A hybrid plasmonic waveguide for subwavelength confinement and long-range propagation,” Nat. Photonics 2(8), 496–500 (2008). [CrossRef]  

34. R. F. Oulton, G. Bartal, D. F. P. Pile, and X. Zhang, “Confinement and propagation characteristics of subwavelength plasmonic modes,” New J. Phys. 10(10), 105018 (2008). [CrossRef]  

35. J. Zhang, L. K. Cai, W. L. Bai, Y. Xu, and G. F. Song, “Hybrid plasmonic waveguide with gain medium for lossless propagation with nanoscale confinement,” Opt. Lett. 36(12), 2312–2314 (2011). [CrossRef]   [PubMed]  

36. Y. Song, M. Yan, Q. Yang, L. M. Tong, and M. Qiu, “Reducing crosstalk between nanowire-based hybrid plasmonic waveguides,” Opt. Commun. 284(1), 480–484 (2011). [CrossRef]  

37. D. X. Dai, Y. C. Shi, S. L. He, L. Wosinski, and L. Thylen, “Gain enhancement in a hybrid plasmonic nano-waveguide with a low-index or high-index gain medium,” Opt. Express 19(14), 12925–12936 (2011). [CrossRef]   [PubMed]  

38. P. B. Johnson and R. W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6(12), 4370–4379 (1972). [CrossRef]  

39. E. D. Palik, Handbook of Optical Constants of Solids (Academic, New York, 1985).

40. P. G. Etchegoin, E. C. Le Ru, and M. Meyer, “An analytic model for the optical properties of gold,” J. Chem. Phys. 125(16), 164705 (2006). [CrossRef]   [PubMed]  

41. I. El-Kady, M. M. Sigalas, R. Biswas, K. M. Ho, and C. M. Soukoulis, “Metallic photonic crystals at optical wavelengths,” Phys. Rev. B 62(23), 15299–15302 (2000). [CrossRef]  

42. C. L. Zou, Y. F. Xiao, Z. F. Han, C. H. Dong, X. D. Chen, J. M. Cui, G. C. Guo, and F. W. Sun, “High-Q nanoring surface plasmon microresonator,” J. Opt. Soc. Am. B 27(12), 2495–2498 (2010). [CrossRef]  

43. J. J. Mock, R. T. Hill, A. Degiron, S. Zauscher, A. Chilkoti, and D. R. Smith, “Distance-dependent plasmon resonant coupling between a gold nanoparticle and gold film,” Nano Lett. 8(8), 2245–2252 (2008). [CrossRef]   [PubMed]  

44. H. J. Chen, T. Ming, S. Zhang, Z. Jin, B. C. Yang, and J. F. Wang, “Effect of the dielectric properties of substrates on the scattering patterns of gold nanorods,” ACS Nano 5(6), 4865–4877 (2011). [CrossRef]   [PubMed]  

45. R. Kolesov, B. Grotz, G. Balasubramanian, R. J. Stöhr, A. A. L. Nicolet, P. R. Hemmer, F. Jelezko, and J. Wrachtrup, “Wave-particle duality of single surface plasmon polaritons,” Nat. Phys. 5(7), 470–474 (2009). [CrossRef]  

46. M. I. Stockman, “Nanoplasmonics: past, present, and glimpse into future,” Opt. Express 19(22), 22029–22106 (2011). [CrossRef]   [PubMed]  

47. L. M. Tong, J. Y. Lou, and E. Mazur, “Single-mode guiding properties of subwavelength-diameter silica and silicon wire waveguides,” Opt. Express 12(6), 1025–1035 (2004). [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 (11)

Fig. 1
Fig. 1 Mathematic model for simulation of a nanowire-substrate waveguiding system.
Fig. 2
Fig. 2 Energy density distribution on the cross section of Au nanowires. (a) D = 50 nm, no substrate; inset, schematic illustration of the polarized fields. (b) D = 50 nm, MgF2 substrate; inset, schematic illustration of the polarized fields. (c) D = 100 nm, MgF2 substrate; (d) D = 200 nm, MgF2 substrate; (e) D = 100 nm, no substrate; (f) D = 100 nm, SiO2 substrate; (g) D = 100 nm, ITO substrate; (h) D = 100 nm, TiO2 substrate. The wavelength of light used here is 660 nm.
Fig. 3
Fig. 3 Numerical solutions of the real part of propagation constants (Re(β)) of (a) Au nanowires, and (b) Ag nanowires. Insets, Re(β) of the nanowire diameter larger than 100nm. The wavelength of light used here is 660 nm.
Fig. 4
Fig. 4 Numerical solutions of the imaginary part of propagation constants (Im(β)) of (a) Au nanowires, and (b) Ag nanowires. Insets, Im(β) of the nanowire diameter larger than 100nm. The wavelength of light used here is 660 nm.
Fig. 5
Fig. 5 Energy density distribution on the cross section of Ag nanowires placed on SiO2. (a) D = 10 nm; (b) D = 20 nm; (c) D = 50 nm; (d) D = 100 nm. The wavelength of light used here is 660 nm.
Fig. 6
Fig. 6 Fractional power of the plasmon mode of Au nanowires inside the (a) core (ƞc), and (b) substrate (ƞs). Inset of (a), a close-up view of the ƞc for Au nanowires with MgF2 and SiO2 substrates, respectively. The wavelength of light used here is 660 nm.
Fig. 7
Fig. 7 Fractional power of the plasmon mode of Ag nanowires inside the (a) core (ƞc), and (b) substrate (ƞs). Inset of (a), a close-up view of the ƞc for Ag nanowires with MgF2 and SiO2 substrates, respectively. The wavelength of light used here is 660 nm.
Fig. 8
Fig. 8 Effective mode area (Am) of (a) Au nanowires, and (b) Ag nanowires. The wavelength of light used here is 660 nm.
Fig. 9
Fig. 9 Normalized amplitude of the electric filed along y direction of a 200-nm-diameter Au nanowire. The wavelength of light used here is 660 nm. Inset, coordinates on the cross section of the nanowire.
Fig. 10
Fig. 10 Propagation lengths (Lm) and losses (α) of (a) Au nanowires, and (b) Ag nanowires. The wavelength of light used here is 660 nm.
Fig. 11
Fig. 11 Substrate index (nsub)-dependent propagation length (Lm) of the plasmon mode for (a) Au nanowires, and (b) Ag nanowires. The wavelength of light used here is 660 nm.

Tables (1)

Tables Icon

Table 1 Refractive indices of substrates at 660 nm wavelength

Equations (7)

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

β = Re ( β ) + i Im ( β ) ,
L m = 1 2 Im ( β ) .
α = 10 log ( 1 / e ) L m 4.343 L m .
A m = W m max { W ( r ) } ,
W ( r ) = 1 2 ( d ( ε ( r ) ω ) d ω | E ( r ) | 2 + μ 0 | H ( r ) | 2 ) ,
η c = c o r e W ( r ) d A W m ,
η s = s u b s t r a t e W ( r ) d A W m ,
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.