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

Collective plasmonic oscillations in gold nanostrips arrays

Open Access Open Access

Abstract

Excitation of collective plasmonic modes and their effect on optical behavior are experimentally and theoretically studied in 1D arrays of gold nanostrips in comparison with continuous gold films with periodically modulated profile. In strips, the angular dependence of the reflectivity demonstrates a peak at the resonance condition as opposed to a dip observed in continuous sine wave gratings. In addition, an extremely narrow feature in the reflection is observed in strips and tentatively ascribed to the bright Wood-Rayleigh anomaly. Theoretical calculations based on the combined transfer-matrix coupled-wave analysis and coordinate transformation method are shown to fit the experimental angular and spectral behavior of the plasmonic resonances. The effects are also discussed in terms of a simple equivalent circuit model.

© 2017 Optical Society of America

1. Introduction

Over the last few decades, surface plasmons, i. e. collective oscillations of free electrons at metal-dielectric interfaces, have attracted much scientific attention and been studied for various potential applications, such as light manipulation [1], biomedical sensing [2], energy conversion [3], emission enhancement [4], nonlinear optics [5], etc. Depending on parameters and geometry of the system, different forms of surface plasmons can be excited, ranging from localized surface plasmons (LSP) observed in individual metal nanoparticles to propagating surface plasmon polaritons (SPPs) [6,7] at continuous metal dielectric interfaces. For the same frequency of the electromagnetic wave, the k-vector of SPP, kspp, is higher than k vector of the photon, k0, and can be found as [7]

ksppk0=ξ=εmεdεm+εd,
where εm and εd are dielectric permittivities of the metal and dielectric, respectively.

Metal surfaces with periodically modulated profiles, such as a continuous sine-wave grating, can provide matching conditions for SPP excitation [6] at

kspp=nG+kx,
where n≠0 is an integer, G = 2π/d is the grating vector, d is the grating period, kx = k0 sinθ is the projection of the optical k-vector onto the interface plane, and θ is the angle of incidence. The excitation of SPPs in continuous metal surfaces with profile modulation is well studied in the literature (including gratings with different depth of modulations, sinusoidal and non-sinusoidal profiles [6–10]). Discontinuous systems such as one-dimensional arrays of slits in metal [11,12], nanohole arrays [13–15], or arrays of plasmonic nanostructures [16] can provide additional opportunities for sensing applications, allowing one to collect both reflected and transmitted light. In addition, such structures can be of interest for tunable optics if the strips are deposited on stretchable substrates or with adding an additional conducting layer [17]. As was discussed [16], discontinuous systems such as arrays of metal nanostructures can exhibit collective modes with Q factors significantly exceeding Q factors of localized plasmon resonances of individual nanostructures. In [11], the arrays of slits in metal were discussed theoretically in terms of photonic band structures. In our work we experimentally test arrays of gold nanostrips for excitation of collective plasmonic waves and discuss their optical behavior in close comparison with continuous systems with similar period of modulation.

2. Experiment

In our experiments we use an array of gold nanostrips with periodicity ds = 358 ± 1 nm, height hs = 100 ± 10 nm and width ws = 179 ± 1 nm. For comparison purposes, we use continuous gold gratings with sine-wave profile modulation with various periodicity dg, depth of modulation hg = 40 - 60 nm, and thickness of gold film ls = 60 ± 3 nm. The schematic of the strips and gratings is shown in Figs. 1(a) and 1(b) correspondingly.

 figure: Fig. 1

Fig. 1 Schematics of (a) strips array, (b) sine-wave grating, (c) SEM image of the strips array.

Download Full Size | PDF

The sinusoidal gratings were fabricated using the laser interference lithography technique [18,19]. The technique relies on recording the interference of two coherent beams with the photosensitive compound (photoresist). First, a glass substrate was spin coated with S1805 photoresist at a thickness of ~100 nm followed by 1 minute soft bake at 98° C. The sample was exposed to an interference pattern using He-Cd laser (λ = 332 nm) at an exposure dose of ~1.2 J/cm2. Then the sample was developed by MF-26 for 30 second to create the sinusoidal polymer grating. Finally, a thin film of gold (60 nm) was deposited with the thermal evaporation method. Arrays of gold strips on glass substrates were fabricated using a similar method. First with the interferometric technique a photoresist pattern in SU-8 was formed on top of a metal film and then the pattern was transferred into the metal with a reactive ion etching (dry etch).

The reflectivity of the samples, R, is measured using He-Ne laser at the wavelength of 632.8 nm and p polarization. The samples are oriented with the direction of strips/grooves vertical (perpendicular to the incidence plane, see the setup in Fig. 2(a)). The reflected beam intensity is recorded as a function of the incidence angle θ. In the continuous gratings, a dip is observed, Fig. 2(b). This dip is a typical signature of the SPP excitation [7]. Note that in our case dg < λ and the SPP is excited “backward”, in the direction opposite to the projection of the optical k-vector on the sample plane. The angular position of the dip is well described with the Eq. (2) with n = - 1 as

sinθspp= λd ξ,
which predicts the angle of 18° and 51° in gratings with dg = 463 nm and 350 nm correspondingly, assuming using standard optical parameters for gold [20].

 figure: Fig. 2

Fig. 2 (a) Setup for angular reflectivity measurement; Reflectivity vs angle of incidence in (b) gratings with dg = 463 nm (red) and 350 nm (blue), and (c) strips; (d) Narrow feature in the reflectivity in more detail: experiment (points), fitting with Lorentzian line with 2 δ=0.03° (solid trace).

Download Full Size | PDF

In the strips, the angular dependence of the reflectivity demonstrates quite a different behavior. No pronounced minimum is observed: Instead, the reflectivity demonstrates a maximum which can be approximated with a sum of a broad peak at ~44-46° and an additional extremely narrow peak at 49.77° (see Figs. 2(c) and 2(d)). The angular width of the narrow feature obtained from the fitting with the Lorentzian line with the angular halfwidth of 0.03°, which is on the order of the divergence angle (~1 mrad) for a standard He-Ne laser [21]. Note that a very narrow feature of the opposite character (a sharp dip in reflection) is predicted in the continuous square-wave gratings [10] resulting from strong enhancement of plasmonic fields at certain conditions of excitation. However, we believe that in our case, the origin of the sharp peak is associated with the bright Wood-Rayleigh anomaly expected at the condition [15].

ωcεd=|k0 sin θG|, 

A peak of this origin can be seen in the gratings as well, see the reflectivity of the gratings with dg = 463 nm, Fig. 2(b), however with a much smaller magnitude and a broader width.

In nanohole arrays in gold and palladium films [14,15], intense peaks in optical transmission spectra with a relatively narrow spectral width of 45 nm were observed and ascribed to the combined effect of the Rayleigh anomaly and SPP resonance on the opposite sides of the metal film. In order to better understand the origin of the peaks in our strips, we measure the spectral dependence of the absorption in the strip array and continuous grating with similar periodicity, dg = 350 nm, collecting all reflected, scattered and transmitted light. The reflection spectra at different angles are obtained with the sample placed in the center of the integrating sphere (Fig. 3(a)) at the horizontal light polarization and vertical orientation of the strips/grooves. Typical results are presented in Figs. 3(b) and 3(c). No narrow features are observed. Spectral dependences of the total collected intensity show relatively broad dips in the both grating and strips, associated with an increased absorption at resonance conditions. From the width of the dips, Q-factors of the resonance are estimated and presented in Fig. 3(d). For the strips they are approximately half that of the grating for the most of optical range. From the dependence of the dip position on the angle of incidence, one can find the dispersion relationship ω (k) for the plasmonic waves in the strips and gratings, estimating its k-vector as k = G - k0 sin θ.

 figure: Fig. 3

Fig. 3 (a) Schematic of the setup in spectral measurements; total reflection spectra for (b) gratings and (c) strips; (d) Q-factor and (e) dispersion relationship for the grating and strips as indicated. Solid curve is the theoretical prediction for a flat gold-air interface.

Download Full Size | PDF

In Fig. 3(e), the dispersion curve in the continuous grating is very similar to what is theoretically calculated for a flat gold-air interface using standard parameters for gold [20]. However for the strips, the dependence is noticeably different, featuring a lower slope (slower propagation of the plasmonic mode).

3. Theory

In our numerical simulations we consider the continuous sine-wave structure and the strips array having the same period of 358 nm.

We use the coordinate transformation method [22] to calculate electromagnetic fields in continuous sine-wave gratings, presenting the fields E and H as functions of coordinates u = x and v = z - a(x), where a(x) = hg sin(Gx) /2. The fields are calculated for different modulation depths, hg, as the functions of angle of incidence, θ, and photon frequency, ω. Figure 4(a) shows the reflectivity coded as color in the dependence of θ and ω obtained in the grating with 40 nm modulation which corresponds to our experimental sample. The SPP curve (seen as a relatively narrow feature in yellow, spreading from (1.95 eV, ~50°) to (2.3 eV to ~25°)) corresponds to the decreased reflectivity and can be well described with Eq. (2). The reflectivity profiles at 42° normalized to the reflectivity of flat gold, Rflat, are shown in Fig. 4(b).

 figure: Fig. 4

Fig. 4 (a) and (b) Reflectivity in gratings: (a) R as the function of incidence angle and photon frequency at 40 nm depth modulation; (b) spectral profiles normalized to that in gold at various modulation depths (curves are shifted vertically). (c) and (d) Reflectivity in strips: (c) R as the function of θ   and ω at hs = 100 nm; (d) spectral profiles of R at different strip heights, (e), and (f) snap shot of magnetic fields at the resonance conditions indicated with arrows.

Download Full Size | PDF

The combined coupled-wave/transfer matrix approach [23] is employed to numerically simulate fields in the array of strips. The calculated reflectivity is shown in Fig. 4(c) for our experimental strip height hs = 100 nm. In comparison with continuous gratings (Fig. 4(a)), it shows quite an opposite character. Instead of the SPP-related decrease in reflectivity clearly observed in Fig. 4(a), a curve with almost the same shape and position indicates an increase in R, Fig. 4(b). In similarity with the grating, it corresponds to the propagation of a plasmonic mode along the metal-air interface. The cross-section of the graph at θ = 42° for different hs is shown in Fig. 4(d). As one can see, starting from moderate heights, hs = 30 nm and higher, the reflectivity maximum is predicted. A “snap-shot” of the magnetic fields at the resonance conditions in the strips with hs = 100 nm is shown in Fig. 4(e). Calculations predict another branch of plasmons in the strips; propagating along gold-glass interface, see Fig. 4(f), which corresponds to an increased reflectivity as well (Fig. 4(c), low angles and frequencies).

Thus, the results of numerical simulations correspond to experimental observations and predict an enhanced reflection in the strips at the resonance conditions. The physical origin of this effect can be viewed in terms of the resonance-related enhancement of the extinction cross-section [24]. This effect can also be discussed in terms of an equivalent circuit model, which considers the energy exchange between free space and a plasmonic mode in terms of the energy exchange between a transmission line and a resonance cavity. As shown in [25], this model provides a reasonably good description of losses associated with plasmonic modes in sine-wave gratings, assuming the coupling constant proportional to the modulation amplitude squared. Here we consider a plasmonic mode as a single resonance cavity and make an attempt to account for additional channels of energy leakage related to the transmission, scattering and diffraction from the sample, including an additional active impedance, R2 as shown in Fig. 5. The reflectivity can be found as [26]

|Γ|2=|Z0ZZ0+Z|2
where Z0 = R0 is the impedance of the transmission line, and Z is the input impedance,
Z=iωLo+ωM12L1iω2ω02ω2+iωω0Q+R2,
ω0(L1 C1)-1/2, and Q = ω0L1/R1. In our model, we assume R0 >> L0.

 figure: Fig. 5

Fig. 5 (a) Equivalent circuit; (b) spectral profiles calculated at Q = 90, R2 = 0, and various M2/L1R0 = 0.0015 n2, where n = 1, 2, 3, 4 as indicated; (c) and (d) calculations at Q = 40, M2/L1R0 = 0.0375 and R2/R0 = 0, 0.05, 0.3, 0.5 and 1.

Download Full Size | PDF

In Fig. 5(b) the reflectivity is calculated at a relatively high Q-factor = 90, an absence of additional losses, R2 = 0, and a different coupling ratio M2/L1R0 varied in the proportion 1:4:9:16. The shape of the resonance dip is similar to what is observed and calculated for sine-wave gratings at the different surface profile modulation depths (compare with the curves presented in Fig. 4(b), where the modulation depth varied as 1:2:3:4). The major difference between curves in Figs. 4(b) and 5(b) is seen in the spectral range higher than the frequency of decoupling of the −1st diffraction order, which was not taken into account in the equivalent circuit calculations.

In order to model effects in the strips, let us assume a Q-factor = 40 and a high mutual inductance, M2/L1R0 = 0.0375. An additional energy leakage (related to transmission, scattering and diffraction), is taken into account assuming various R2 >0. In Fig. 5(c) adding of a small active resistance R2 = 0.05 R0 (low energy leakage) leads to an overall decrease in the reflectivity demonstrating a common resonance-related dip in the reflection. However, at high R2/R0, instead of a dip, a peak in reflection back to the transmission line is expected at the resonance conditions (Fig. 5(d)). Thus, this simple model predicts the existence of the resonance peak from the resonance conditions themselves without an account for the bright Rayleigh anomaly. In order to explain the origin of the sharp feature in the frames of the equivalent circuit model, one can speculate that in the strips, the major energy leakage mechanism is related to diffraction. (In fact, in our experiment, up to 55% of incoming power is diffracted to the −1th order of diffraction.) If this leakage channel does not exist (at the conditions defined by Eq. (4)), in the equivalent circuit model this situation corresponds to R2 = 0 and a high reflectivity only at this particular condition (as schematically shown with the dotted line in Fig. 5(d)).

In conclusion, we present detailed experimental and theoretical studies of optical effects in a discontinuous system of gold nanostrips in close comparison with a continuous sine-wave grating of the same periodicity. In spite of a great similarity between collective plasmonic modes in gold continuous gratings and those in arrays of strips, their optical signatures can be significantly different. In the strips, SPPs can be excited at both front and back surfaces, and manifest themselves as a maximum in reflectivity as the opposite to the common SPP-related dip in reflectivity in continuous films. The Q-factor of these modes varies from 10 to 66 in the visible range. The presence of an additional strong and extremely narrow (with the angular width of 0.03°) peak in the strips reflectivity is a very interesting feature and can present new opportunities for applications. The experiments are supported with numerical simulations, and a simple equivalent circuit model is considered which provides a qualitative description of the observed effects.

Funding

National Science Foundation (NSF), (1205457, 1646789, 0966188)

References and links

1. J. A. Schuller, E. S. Barnard, W. Cai, Y. C. Jun, J. S. White, and M. L. Brongersma, “Plasmonics for extreme light concentration and manipulation,” Nat. Mater. 9(3), 193–204 (2010). [CrossRef]   [PubMed]  

2. S. V. Jenkins, T. J. Muldoon, and J. Chen, “Plasmonic nanostructures for biomedical and sensing applications,” in Metallic Nanostructures, Y. Xiong, X. Lu, eds. (Springer, 2015).

3. S. K. Cushing and N. Wu, “Progress and perspectives of plasmon-enhanced solar energy conversion,” J. Phys. Chem. Lett. 7(4), 666–675 (2016). [CrossRef]   [PubMed]  

4. K. J. Russell, T.-L. Liu, S. Cui, and E. L. Hu, “Large spontaneous emission enhancement in plasmonic nanocavities,” Nat. Photonics 6(7), 459–462 (2012). [CrossRef]  

5. T. Hanke, G. Krauss, D. Träutlein, B. Wild, R. Bratschitsch, and A. Leitenstorfer, “Efficient nonlinear light emission of single gold optical antennas driven by few-cycle near-infrared pulses,” Phys. Rev. Lett. 103(25), 257404 (2009). [CrossRef]   [PubMed]  

6. H. Raether, Surface Plasmons on Smooth and Rough Surfaces and on Gratings (Springer Verlag, Berlin, 1988), Introduction.

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

8. A. C. R. Pipino and G. C. Schatz, “Surface-profile dependence of photon-plasmon-polariton coupling at a corrugated silver surface,” J. Opt. Soc. Am. B 11(10), 2036–2045 (1994). [CrossRef]  

9. R. A. Watts, A. P. Hibbins, and J. R. Sambles, “The influence of grating profile on SPP resonances recorded in different diffracted orders,” J. Mod. Opt. 46, 2157–2186 (1999). [CrossRef]  

10. P. Sheng, R. S. Stepleman, and P. N. Sanda, “Exact eigenfunctions for square-wave gratings: Application to diffraction and surface-plasrnon calculations,” Phys. Rev. B 26(6), 2907–2916 (1982). [CrossRef]  

11. R. Marani, M. Grande, V. Petruzzelli, and A. D’Orazio, “Plasmonic bandgaps in 1D arrays of slits on metal layers excited by out-of-plane sources,” Int. J. Opt. 2012, 146396 (2012). [CrossRef]  

12. J. M. Steele, C. E. Moran, A. Lee, C. M. Aguirre, and N. J. Halas, “Metallodielectric gratings with subwavelength slots: Optical properties,” Phys. Rev. B 68(20), 205103 (2003). [CrossRef]  

13. A. Hajiaboli, M. Kahrizi, and V.-V. Truong, “Optical behavior of thick gold and silver films with periodic circular nanohole arrays,” J. Phys. D Appl. Phys. 45(48), 485105 (2012). [CrossRef]  

14. J. M. McMahon, J. Henzie, T. W. Odom, G. C. Schatz, and S. K. Gray, “Tailoring the sensing capabilities of nanohole arrays in gold films with Rayleigh anomaly-surface plasmon polaritons,” Opt. Express 15(26), 18119–18129 (2007). [CrossRef]   [PubMed]  

15. H. Gao, J. M. McMahon, M. H. Lee, J. Henzie, S. K. Gray, G. C. Schatz, and T. W. Odom, “Rayleigh anomaly-surface plasmon polariton resonances in palladium and gold subwavelength hole arrays,” Opt. Express 17(4), 2334–2340 (2009). [CrossRef]   [PubMed]  

16. B. D. Thackray, V. G. Kravets, F. Schedin, G. Auton, P. A. Thomas, and A. N. Grigorenko, “Narrow collective plasmon resonances in nanostructure arrays observed at normal light incidence for simplified sensing in asymmetric air and water environments,” ACS Photonics 1(11), 1116–1126 (2014). [CrossRef]  

17. B. Thackray, V. G. Kravets, F. Schedin, R. Jalil, and A. N. Grigorenko, “Resistive coupling of localized plasmon resonances in metallic nanostripes through a graphene layer,” J. Opt. 15(11), 114002 (2013). [CrossRef]  

18. W. Fan, S. Zhang, K. J. Malloy, and S. R. J. Brueck, “Large-area infrared nanophotonic materials fabricated using interferometric lithography,” J. Vac. Sci. Technol. B 23(6), 2700–2704 (2005). [CrossRef]  

19. H. C. Guo, D. Nau, A. Radke, X. P. Zhang, J. Stodolka, X. L. Yang, S. G. Tikhodeev, N. A. Gippius, and H. Giessen, “Large-area metallic photonic crystal fabrication with interference lithography and dry etching,” Appl. Phys. B 81(2-3), 271–275 (2005). [CrossRef]  

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

21. He-Ne lasers specifications, https://www.thorlabs.com.

22. J. Chandezon, G. Raoult, and D. Maystre, “A new theoretical method for diffraction gratings and its numerical application,” J. Opt. 11(4), 235–241 (1980). [CrossRef]  

23. D. Keene and M. Durach, “Hyperbolic resonances of metasurface cavities,” Opt. Express 23(14), 18577–18588 (2015). [CrossRef]   [PubMed]  

24. J. Jung, T. Søndergaard, J. Beermann, A. Boltasseva, and S. I. Bozhevolnyi, “Theoretical analysis and experimental demonstration of resonant light scattering from metal nanostrips on quartz,” J. Opt. Soc. Am. B 26(1), 121–124 (2009). [CrossRef]  

25. N. Noginova, M. LePain, V. Rono, S. Mashhadi, R. Hussain, and M. Durach, “Plasmonic pressure in profile-modulated and rough surfaces,” New J. Phys. 18(9), 093036 (2016). [CrossRef]  

26. C. G. Montgomery, R. H. Dicke, and E. M. Purcell, “The reflectivity,” in Principles of Microwave Circuits (IET, 1948).

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 (5)

Fig. 1
Fig. 1 Schematics of (a) strips array, (b) sine-wave grating, (c) SEM image of the strips array.
Fig. 2
Fig. 2 (a) Setup for angular reflectivity measurement; Reflectivity vs angle of incidence in (b) gratings with dg = 463 nm (red) and 350 nm (blue), and (c) strips; (d) Narrow feature in the reflectivity in more detail: experiment (points), fitting with Lorentzian line with 2 δ=0.03° (solid trace).
Fig. 3
Fig. 3 (a) Schematic of the setup in spectral measurements; total reflection spectra for (b) gratings and (c) strips; (d) Q-factor and (e) dispersion relationship for the grating and strips as indicated. Solid curve is the theoretical prediction for a flat gold-air interface.
Fig. 4
Fig. 4 (a) and (b) Reflectivity in gratings: (a) R as the function of incidence angle and photon frequency at 40 nm depth modulation; (b) spectral profiles normalized to that in gold at various modulation depths (curves are shifted vertically). (c) and (d) Reflectivity in strips: (c) R as the function of θ   and ω at hs = 100 nm; (d) spectral profiles of R at different strip heights, (e), and (f) snap shot of magnetic fields at the resonance conditions indicated with arrows.
Fig. 5
Fig. 5 (a) Equivalent circuit; (b) spectral profiles calculated at Q = 90, R2 = 0, and various M2/L1R0 = 0.0015 n2, where n = 1, 2, 3, 4 as indicated; (c) and (d) calculations at Q = 40, M2/L1R0 = 0.0375 and R2/R0 = 0, 0.05, 0.3, 0.5 and 1.

Equations (6)

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

k spp k 0 =ξ= ε m ε d ε m + ε d ,
k spp =nG+ k x ,
sin θ spp =  λ d  ξ,
ω c ε d =| k 0  sin θG |, 
| Γ | 2 = | Z 0 Z Z 0 +Z | 2
Z=iω L o + ω M 1 2 L 1 i ω 2 ω 0 2 ω 2 + iω ω 0 Q + R 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.