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

Momentum-dependent group velocity of surface plasmon polaritons in two-dimensional metallic nanohole array

Open Access Open Access

Abstract

We determine the momentum-dependent group velocities of ( ± 1,0) and (0, ± 1) Bloch-like surface plasmon polaritons (SPPs) in two-dimensional Au nanohole array by measuring their propagation lengths and decay lifetimes at different SPP propagation length via angle- and polarization-resolved reflectivity spectroscopy and real- and Fourier-space microscopy. We find the decay length and lifetime, as well as group velocity, are highly dependent on the propagation direction. In particular, close to the Γ-M direction where two SPPs begin to interfere, the group velocity decreases due to the increase of the standing wave character. More importantly, the two SPPs are strongly interacted with each other at the Γ-M direction, resulting in forming the dark and bright modes. We find the group velocity of the dark mode is higher that of the bright mode despite its higher quality factor, or longer decay lifetime. We attribute such difference to the distinct field symmetries of dark and bright modes, yielding different effective indices. While bright mode has fields mostly concentrated at the flat metal region to produce higher effective index and therefore lower velocity, the fields of the dark mode are located near the air hole, resulting in higher velocity.

© 2016 Optical Society of America

1. Introduction

Decay is the key process in plasmonics since it defines many intriguing properties of surface plasmon polaritons (SPPs) [1]. The decay dynamics is defined by two parameters, which are the propagation length (LD) and decay rate (Γtot). For example, LD governs how far SPPs can travel before complete dissipation and a long LD is favorable for waveguiding [2]. On the other hand, field enhancement has been indicated to hinge on the interplay between radiative (Γrad) and absorption (Γabs) decay rates [3, 4]. More importantly, the interplay between LD and Γtot defines group velocity (vg = LDΓtot) [5, 6], which is related to the density-of-states (DOS) of the system [7]. Because of their significance, several methods based on spatial-, time-, and frequency-domains have been developed to study the decay process. Kim et al measure the propagating length of SPPs on 2D Au nanohole arrays by using near-field scanning optical microscopy and deduce the length to be 2-3 μm for (1,0) SPPs [8]. On the other hand, Ropers et al have used the time-resolved method to examine the transient of the electric field transmitted through the 1D Au grating and conclude the lifetime, which is 1/Γtot, is ~32 fs for −2 SPPs [9]. They also report the coupling of two degenerate −1 and + 1 SPP modes at normal incidence results in the formation of dark and bright modes due to the strong modification of the SPP radiation damping rates [9]. Dark mode is nonradiative with longer decay lifetime whereas bright mode is highly radiative and thus has shorter lifetime. We have deployed a frequency-domain method to examine the dependence of decay rate on wavelength and hole geometry in 2D nanohole arrays [10]. In particular, we use temporal coupled mode theory (CMT) to express the reflectivity spectrum explicitly on decay rates and have measured the absorption and radiative decay rates of (−1,0) SPPs [11]. More importantly, we find both the spectral profile and the field enhancement are associated with the interplay between the absorption and radiative decay rates [4].

To date, for periodic arrays, the decay rates or lengths have so far been measured in one single angle or direction, mostly along the Γ-X direction for different resonance wavelengths [4, 8–11]. However, it is known that SPPs can propagate in different directions and each could have different decay dynamics. Therefore, other than the spectral dependence, the knowledge of angle- or momentum-dependent LD and Γtot, as well as vg = LDΓtot, is also of great importance. In addition, the angular dependence is expected to play a major role in controlling the directionality of SERS and fluorescence, which are essential directional emission [12–15]. Therefore, the understanding the angle-dependent decay process of SPPs presents not only a scientific interest but also a practical incentive.

In this paper, we have measured the momentum-dependent LD and Γtot of ( ± 1,0) and (0, ± 1) Bloch-like SPPs in a 2D Au nanohole array by using angle- and polarization-resolved reflectivity spectroscopy and real- and Fourier-space microscopy. While reflectivity spectrum directly yields Γtot, we find the linewidth obtained from the equi-energy Fourier-space contour image is associated with LD corresponding to a particular propagation direction, as verified by finite-difference time-domain (FDTD) simulation. Our results show two quantities, together with vg, vary strongly with the propagation direction. In particular, when the direction is changing from Γ-X to Γ-M, vg decreases due to the increase proportion of the standing wave character arising from the coupling of two SPPs. Along the Γ-M direction where two SPPs are degenerate, in addition to the formation of plasmonic gap, two dark and bright modes are formed, giving rise to two distinct velocities. We find although the dark mode has smaller Γtot, its vg is higher than that of the bright mode. As two modes have different field symmetry, we elucidate the fact that the fields of the dark mode are mostly concentrated near the air hole region and thus produce a lower effective index neff and higher vg. More importantly, the distinct vg arising from the dark and bright modes imply different DOS, which could affect Raman and fluorescence emissions.

The paper is organized as follows. Section 2 discusses how one can determine LD from the equi-energy contour by using spatial CMT. FDTD simulation on a simple 1D grating is provided for verifying CMT. In section 3, on an actual 2D nanohole array, we present the measurements of Γtot, LD, and vg as a function of wavelength by using angle-resolved reflectivity spectroscopy along the Γ-X direction. The results serve as a testbed for our approach. Section 4 reports LD at λ = 633 nm as a function of SPP propagation direction by Fourier-space microscopy. Finally, in section 5, we summarize our findings on the dependences of Γtot and vg on SPP propagation direction, with special attention along the Γ-M direction.

2. Analytical and numerical calculations

Both LD and Γtot are two measurable quantities. While the measurement of Γtot has been discussed in earlier works, we focus on the relationship between LD and the linewidth in the equi-energy contour here.

A simple case of 1D metallic periodic nanoslit array is considered for proof of the concept. The nanoslit array is illuminated by p-polarized plane wave with constant real in-plane momentum k//, along the Γ-X direction. We assume the nanoslit array is optically thick and it supports only specular reflection. In analogy to temporal CMT, we write the spatial dynamics of a SPP mode amplitude, a, at a particular frequency ω as: daωdx=ik˜sppaω+κωs+ω=iksppaω12LDaω+κωs+ω, where κω is the in-coupling constant, s+ω is the momentum resolved amplitude of the incident wave power, k˜spp=kspp+i2LD is the complex resonant momentum indicating the SPP mode decays exponentially with x. Here k˜spp is assumed a scalar form but in fact is a vector when we consider the actual nanohole array later. The momentum dependent s+ω is related to kspp by phase matching condition under empty lattice approximation given as kspp=Re(2πλεε+1)=k//+2nπP, where ε is metal permittivity, P is the period of the nanoslit array and n is integer characterizing Bragg scattering order. As a result, one can solve for aω to be: aω=κωsω+i(k//+2nπ/Pkspp)+(1/2LD). Under conservation of energy, the outgoing power sω=Cs+ω+χωaω, where C is the nonresonant scattering constant and χωis the outcoupling constant for SPPs. sω will undergo phase-matching condition k//=kspp2nπP again for far-field radiation. Therefore, in momentum space at a single frequency, we are able to write k-solved reflectivity as:

RPω(k//)=|rpω+χωκωi(k//+2nπ/Pkspp)+(1/2LD)|2,
where rpω=C is the reflection coefficient. It is also expected from Eq. (1) that the denominator of the second term, which is the radiative decay of SPPs, remains unchanged even when the SPPs is not propagating along the Γ-X direction. In other words, by measuring the specular reflectivity of in k-space along the propagating direction, which is perpendicular to the equi-energy contour [5, 7], one could deduce LD from the linewidth.

We conduct FDTD simulations to verify the CMT result. For simplicity, we choose 1D grating as we believe the underlying physics is essentially identical to 2D cases. Two types of simulations have been performed and their simulation cells are shown in Figs. 1(a) and 1(c). The first type has unit cell with P = 510 nm, groove width and depth = 60 and 40 nm. Bloch boundary condition is used at two sides and perfectly matched layer (PML) is set on the top and at the bottom of the cell [4, 11]. We use a plane wave source to illuminate the grating at different k// along the Γ-X direction. Therefore, the incident plane is always perpendicular to the slits. The source has wavelength centered at 725 nm and bandwidth = 150 nm to excite only the −1 SPP mode. The specular reflection is calculated by using the near-to-far-field projection method given earlier [4, 11]. On the other hand, the second type involves a supercell with total cell length = 100 μm, which contains 180 grooves. The geometry of each single groove is identical to that of the first cell and PML is applied on the whole structure. Instead of using a plane wave source, total-field/scattered-field approach is used to exclude the input excitation and record only the scattered field. A power line monitor is set at 4 nm above the metal surface to examine the near-field distribution across the entire surface. Assuming the cell is sufficiently large to allow the SPPs to decay, this real space near-field profile directly yields LD. We first calculate the p-polarized reflectivity as a function of k// = 2πsinθ/λ at several wavelengths from λ = 710 – 770 nm by using the first cell and the results are shown in Fig. 1(b). They all exhibit Fano-like reflection dips positioning at different k//. Both ω and k// match well with the phase-matching equation, revealing −1 SPP mode is excited. By best-fitting the graphs using Eq. (1), LD is extracted and plotted in Fig. 1(e) against resonant wavelength, showing LD increases steadily with wavelength from 5 to 9 μm. This increase is expected because longer wavelength experiences weaker radiative scattering and lower Ohmic absorption, thus resulting in longer lifetime and decay length [1,10]. After that, we then examine the spatial decay of SPPs for the corresponding resonant wavelengths based on the second type simulation. The incident momentum for each wavelength is determined from k// in Fig. 1(b). The intensity profiles in logarithmic scale are shown in Fig. 1(d) and they all display exponential decrease with distance. LD for different wavelengths are then determined from the inverse of the slopes and plotted in Fig. 1(e) for comparison. Apparently, two results agree very well with each other with discrepancy smaller than 2.3%, confirming the applicability of determining LD in k-space.

 figure: Fig. 1

Fig. 1 (a) The FDTD unit cell for 1D groove grating. (b) The simulated reflectivity as a function of incident wavevector for different wavelengths indicating (−1,0) SPP mode is excited at the reflection dips. (c) The FDTD supercell for 1D groove grating. (d) The spatial profile of the scattered near field for different wavelengths in logarithmic scale, showing exponentially decrease of intensity. (e) The plot of propagation length against resonant wavelength deduced from two methods.

Download Full Size | PDF

3. Testbed results

In this section, we begin with measuring Γtot, LD, and vg at different resonant wavelengths by using the angle- and polarization-resolved reflectivity spectroscopy. We have fabricated 2D Au nanohole array by interference lithography [4, 10, 11]. Its scanning electron microscopy image is shown in the inset of Fig. 2(a), showing the nanohole array has period P = 510 nm, hole depth and radius = 200 and 50 nm, respectively. After fabrication, the sample is mounted on a computer controlled goniometer for p-polarized angle-dependent reflectivity mapping [10]. For 2D nanohole array, the phase matching equation should be written in vectorial form as:

kspp=k//+2nπPx^+2mπPy^=(ωcsinθcosφ+2nπP)x^+(ωcsinθsinφ+2mπP)y^,
where φ is the azimuthal angle, n and m are integers and |kspp|=Re(2πλεε+1) [4, 10, 11]. The mapping taken in the Γ-X direction where φ = 0° is shown in Fig. 2(a) which simultaneously provides both the frequency and momentum domains for Γtot = Γrad + Γabs and LD determination. A dispersive reflection band is observed and it is identified as the (−1,0) SPP mode by using the phase-matching equation illustrated as a dashed line. The band at ~2.4 eV is due to the absorption of Au. Several cuts at constant photon energy and wavevectors from the same SPP modes are extracted and then plotted in Figs. 2(b) and 2(c). They all exhibit Fano-like features and are best-fitted for determining Γtot and LD as shown in Fig. 2(d). The group velocities, given as ΓtotLD, are plotted in the inset [5, 6, 16]. We find vg varies from 0.69c to 0.95c. We check this by calculating vg defined as ∂ω/∂k from the mapping and it is plotted in Fig. 2(d) as a dot line for reference, showing vg remains constant at low photon energy but decreases steadily at higher energy due to the opening of the band gap. In fact, the band edge slowly levels off and thus reduces vg in the Γ-X direction. Nevertheless, both ΓtotLD and ∂ω/∂k determined vg are consistent with each other.

 figure: Fig. 2

Fig. 2 (a) The p-polarized wavevector-resolved reflectivity mapping of 2D Au nanohole array taken in Γ-X direction. The dashed line is calculated by using phase-matching equation, indicating the excitation of (−1,0) SPP mode. Inset: the SEM image of the nanohole array. The scale bar is 500nm. (b) The corresponding reflectivity spectra for different incident wavevectors. The dashed lines are the best fits. (c) The corresponding reflectivity plots for different photon energies. The dashed lines are the best fits. (d) The deduced group velocities for different resonant wavelength. The dot line is ∂ω/∂k from the mapping.

Download Full Size | PDF

4. Real- and Fourier-space microscopy for LD measurement

We then turn our attention to measure LD as a function of SPP propagation direction. HeNe 633 nm laser and white light source are used separately for real- and Fourier-space imaging. The schematic of the microscope is shown in Fig. 3(a) [17]. For real-space imaging, the laser light is first coupled into a single-mode fiber before collimated by an achromatic lens. It is then focused onto the back focal plane of a 100X, numerical aperture NA = 0.9 air dry objective lens and finally is output from the lens at well-defined incident and azimuthal angles. The illumination optics is set on a motorized translation stage to vary θ from 0° to 63° with a step size of ~0.2° [17]. The sample is placed on a rotation stage to vary φ from 0° to 360°. A hexagon adjustable iris is positioned at the image plane in the incident path to reduce the beam size so that only part of the sample (i.e. 1/3 of the viewing area) is illuminated. As a result, SPPs could propagate to the un-illuminated region and radiate out to the far-field. The radiation from the sample is collected by the same objective lens and is either spatially imaged by an EMCCD camera or detected by a photodiode. On the other hand, for Fourier-space imaging, the white light is fed on the entire back focal plane so that the sample is fully illuminated at all possible θ and φ as defined by the NA of the objective lens. This configuration excites all allowable SPPs simultaneously with the same |kSPP| but different directions as defined by the phase-matching equation (Eq. (2)). The reflections, including both resonant and non-resonant components, are collected and projected onto the Fourier-space image to yield the equi-energy contour in kx and ky space [18]. A 633 nm laser line filter, with bandwidth FWHM of 3 ± 0.5 nm, is inserted in the detection path to collect the desired wavelength so that a clear and sharp equi-energy contour can be observed for analysis. The technique is similar to those of Regan et al [19], Angelini et al [20], and Wagner et al [21] by using a leakage radiation microscopy to image the angular emission from 2D nanohole arrays and 3D photonic crystals. However, our technique relies only on the reflections from the sample and no emission from fluorescent dyes is required, which often suffers from photobleaching. As a result, for a particular SPP mode along the equi-energy contour, we measure the linewidths normal to the contour to determine the momentum-dependent LD.

 figure: Fig. 3

Fig. 3 (a) The schematic setup for real- and Fourier-space microscopy. (b) The plot of reflectivity against incident angle from 2D Au nanohole array for λ = 633 nm obtained from the microscopy. The dashed line is extracted from the goniometer for comparison. Three crosses are located at off- and on- SPP resonance regions labeled as (c)-(e). (c) & (e) The off-resonance real images. The hexagon indicates the region illuminated by light. The solid line indicates the region extracted for examining the spatial decay of light. (d) The on-resonance real image featuring long light tails extending around the hexagon in a particular direction. The arrows define the directions of incident light and SPP propagation. (f) The extracted spatial decay of light from (c)-(e) in logarithmic scale. Only the on-resonance decay shows exponential behavior. (g) The measured propagation length as a function of azimuthal angle.

Download Full Size | PDF

In analogy to the FDTD simulation, we use the real-space imaging to determine LD for (−1,0) SPP mode at several φ. In principle, if a collimated laser beam is incident on the nanohole array at θ and φ according to the phase-matching equation, SPPs propagate at a well-defined direction and subsequently undergo dissipation. The spatial distribution of the leakage radiation can be imaged for LD measurement [22]. We first identify the desired incident angle for exciting SPPs. Figure 3(b) shows the p-polarized angular reflectivity of the nanohole array at λ = 633 nm taken by CCD-based spectrometer along the Γ-X direction. The reflectivity from Fig. 2(b) is also extracted and superimposed on the plot as the dashed line for comparison. Two agree with each other, indicating (−1,0) SPP mode is excited at θ = 9.3°. We then select three incident angles marked by the crosses in Fig. 3(b), spanning from off- to on-resonance, and capture their images in Figs. 3(c)-3(e) for comparison. Apparently, one sees the image taken at on-resonance is completely different from the off-resonance counterparts. For the off-resonance images in Figs. 3(c) and (e), the hexagons, which are the opening of the iris, are bright due to the strong non-resonant reflection background. In contrast, at on-resonance, the hexagon is dark, corresponding to the low reflection dip (Fig. 3(d)). More importantly, we see from the on-resonance image that the edges of the hexagon are not sharp but featured with long light tails on both sides pointing in one distinct direction [22]. In fact, the direction of the tail indicates the direction where the (−1,0) SPP mode is propagating. In this case, the mode is traveling in the –x direction, which is opposite to the incident direction as indicated by the arrow in Fig. 3(d). The radiation decreases in intensity with increasing distance. The off-resonance edges are sharp and do not show such behavior. We then extract the intensity profiles along the incident plane from the figures and show them in Fig. 3(f) in logarithmic scale. We see only the resonance decay curve exhibits a single exponential decrease with distance and the slope yields LD = 3 μm. We then repeat the same procedure for azimuthal angles φ = 10° and 20° and their corresponding LD for (−1,0) SPPs are plotted in Fig. 3(g). We find LD varies weakly with φ.

For Fourier-space imaging, an unpolarized light is incident on the nanohole array and the reflection image at λ = 633 nm is shown in Fig. 4(a) displaying four low reflection arcs at well-defined angles [14]. The nanohole array is rotated such that its Γ-X direction is parallel with kx. A four-fold symmetry is expected as the nanohole array is square lattice. The background is due to the non-resonant reflection whereas the arcs arise from the excitation of SPPs. In fact, they can be traced by the phase-matching equation and the calculated curves are shown in Fig. 4(a) as solid lines identifying (1,0), (−1,0), (0,1), and (0,-1) as well as (1,1), (−1,1), (1,-1), and (−1,-1) SPP modes are excited [10, 11, 14]. Wood’s anomalies are also observed and plotted as dashed lines [23]. In fact, the image defines the plasmonic equi-energy contour with the first Brillouin zone given as the dot square with length = 12.32 μm−1 [24]. In analogy to electronic crystal, owing to the square lattice, only 1/8 of the Brillouin zone is required with the reduced Brillouin zone defined by the triangle which indicates the Γ-X-M region [25]. Therefore, the decay lengths measured in this region are sufficient to represent the entire system. To confirm our measurement, we also have mapped the equi-energy contour by using the angle-resolved goniometry and the region taken at λ = 633 nm is extracted in Fig. 4(b) for comparison. In fact, two contours are very similar and they share common features.

 figure: Fig. 4

Fig. 4 (a) The unpolarized Fourier reflection image taken at λ = 633 nm from 2D Au nanohole array taken by microscopy. The solid and dashed lines indicate the SPP modes and Wood’s anomalies calculated from the phase-matching equations. The dot line defines the first Brillouin zone together with the reduced zone given by Γ-X-M triangle. (b) The unpolarized k-space mapping at λ = 633 nm from 2D Au nanohole array taken by angle-resolved goniometry. (c) Several (−1,0) SPP cuts taken at different azimuthal angles are extracted from the unpolarized Fourier image. The dashed lines are the best-fits for determining the corresponding propagation lengths. (d) The deduced (■) propagation length as a function of azimuthal angle. The positions of dark and bright modes and the Wood’s anomaly are also indicated. The (●) LD obtained from the real-space imaging in Fig. 3(g) is overlaid for comparison.

Download Full Size | PDF

When tracing along the (−1,0) SPP mode, we find the reflectivity contrast changes noticeably. In particular, we note from Figs. 4(a) and 4(b) that at the point where (−1,0) and (0,-1) SPP modes cross, a small plasmonic gap is found at k=(0.2x^+0.2y^)k0 in which propagation of SPPs is forbidden. At the same time, two new coupled modes are formed at different k vectors along the Γ-M direction [9, 26]. In fact, this hybridized pair signifies the formation of dark and bright modes [9, 26, 27]. Specifically, dark mode is weakly radiative whereas bright mode is highly radiative and they thus have different damping rates as well as propagation lengths. Therefore, the contrast clearly identifies the dark and bright modes are located at k=(0.213x^+0.213y^)k0and(0.174x^+0.174y^)k0, respectively. One also sees the reflectivity changes quite abruptly around the Wood’s anomalies in which the number of diffraction orders changes [23]. All these observations suggest that LD is not a constant with φ but changes continuously along the (−1,0) SPP resonance. We then evaluate LD as a function of φ or kSPP by extracting cuts perpendicular to the (−1,0) SPP contour at different azimuthal angles and several of them are displayed in Fig. 4(c). We fit them by using Eq. (2) and plot LD against φ in Fig. 4(d). The LD obtained from the real-space imaging in Fig. 3(g) is also overlaid for comparison. From the plot, one sees although LD remains almost unchanged for small azimuthal angles, it gradually decreases when approaching to the band gap region, yielding 2.77 and 1.33 μm for the dark and bright modes, respectively. LD increases again afterwards.

5. Dependences of Γtot and vg on SPP propagation direction

Before the determination of k-dependent vg, we measure the corresponding Γtot. Again, we measure the Γtot of (−1,0) SPP mode at λ = 633 nm by using goniometer. For different φ, we extract the reflectivity spectra at the corresponding θ and fit them by temporal CMT to determine Γtot [11] and the results are plotted in Fig. 5(a). We see Γtot varies strongly with SPP propagation direction. It remains almost constant at ~40-44 meV between φ = 0° - 30°, splits into two bright and dark values of 56 and 37 meV when under coupling at φ = 45°, and finally increases again at larger φ. To explain these variations, we find from φ = 0° - 30° that kSPP varies from 1.067x^k0 to (1.063x^+0.1y^)k0 indicating the propagation direction of (−1,0) SPPs rotates only by 5.4° from the Γ-X direction. Such small directional change does not alter the damping of SPPs much and thus leads to an almost constant Γtot. However, when φ approaches to 45°, two (−1,0) and (0,-1) SPP modes begin to interfere and the resulting standing waves significantly alter the propagation direction. More importantly, the formation of dark and bright modes gives rise to strong modification of the decay rates, with the rates of dark and bright modes lower and higher than the nondegenerate counterparts. In fact, the splitting can be explained within the framework of CMT. Consider two SPPs with mode amplitudes a1 and a2, their time variants can be written as [9, 28]:

ddt[a1a2]=[iωoΓtot/2iω12iω21iωoΓtot/2][a1a2],
where ω12is the complex coupling constant between modes. After solving the determinant of the square matrix, two new modes at ωo±Re[ω12] emerge and they have decay rates given as Γtot±2Im[ω12]. Therefore, by measuring the gap size to be ~34 meV and the difference between two decay rates to be 18.5 meV, we deduce the coupling constant to be ω12=17+i4.6meV. At larger φ, the increase of Γtot is due to the increase of the number of radiative decay channels after crossing the Wood’s anomalies located as 52.7°. Our FDTD results support the experiment. We simulate Γtot, Γrad, and Γabs of the (−1,0) SPP mode in 2D Au nanohole array with P = 510 nm, hole depth and radius = 200 and 80 nm as a function of φ by using the time-domain method as described previously [4, 11] and results are plotted in Fig. 5(b). Clearly, the simulated Γtot is consistent with experiment. Both Γrad and Γabs are constants at low φ and similar splitting is seen when approaching towards 45°. Finally, Γrad increases significantly after the Wood’s anomalies.

 figure: Fig. 5

Fig. 5 (a) The experimental total decay rate Γtot of (−1,0) SPPs plotted as a function of azimuthal angle φ. The solid line indicates the emergence of Wood’s anomaly. (b) The FDTD simulated (▲) total decay Γtot, (■) radiative decay Γrad, and (●) absorption Γabs rates of (−1,0) SPPs as a function of azimuthal angle φ. At φ = 45°, two (−1,0) and (0,-1) SPP modes couple, leading to the splitting of the rates.

Download Full Size | PDF

After discussing the decay rates, we move on to determine vg = LDΓtot and the results are shown in Fig. 6(a). We see vg gradually decreases from 0.7c to 0.6c from φ = 0° - 30° as the SPP propagation direction remains almost unchanged. However, above 30° vg decreases dramatically as the standing wave begins to take off. More interestingly is at the Γ-M direction where two dark and bright modes form, the dark mode has higher vg despite its lower Γtot. To confirm it, we display in Fig. 6(b) the unpolarized incident momentum-resolved reflectivity mapping taken at φ = 45°. The dark and bright modes are clearly seen and their dispersions are weakly dependent on momentum. As a result, two modes are expected to have slow group velocities compared to the (−1,0) nondegenerate SPPs. The vg of dark and bright modes are then determined to be 0.43c and 0.47c, respectively, which are consistent with the k-space measurement.

 figure: Fig. 6

Fig. 6 (a) The plots of vg of (−1,0) SPPs as a function of azimuthal angle φ. At φ = 45°, two (−1,0) and (0,-1) SPP modes couple, leading to the splitting of vg. (b) The unpolarized incident angle-resolved reflectivity mapping taken at φ = 45°, indicating the presence of dark and bright modes. The dispersions are weak.

Download Full Size | PDF

To address the distinct vg for the dark and bright modes, we use FDTD to simulate the plane-view field patterns superimposed with the Poynting vector maps of the bright and dark modes in Figs. 7(a) and 7(b). The Poynting vectors unambiguously indicate the energy flows in the Γ-M direction for two cases. From the field patterns, the crests and troughs do not represent the wavefronts of a traveling plane wave but the profile of a standing wave. While two coupled SPP coupled modes travel in the Γ-M direction, they are standing perpendicular to it. In fact, two modes display different field symmetry in the y-direction. The dark mode is asymmetric whereas the bright mode is symmetric [9, 29]. This difference results in different field localizations. As the fields of the dark mode are more localized near the air hole region, it produces lower neff, which gives rise to higher vg. On the other hand, the bright mode has larger neff and thus lower vg.

 figure: Fig. 7

Fig. 7 The plane-view field patterns |E| and the Poynting vector maps for (a) bright and (b) dark modes simulated by FDTD for λ = 633 nm at φ = 45°.

Download Full Size | PDF

6. Conclusion

In summary, we have employed angle- and polarization-resolved reflectivity spectroscopy and real- and Fourier-space microscopy to study the momentum-dependent Γtot and LD, and vg, of Bloch-like SPPs in 2D Au nanohole array. The Γtot has been determined by reflectivity spectroscopy at different SPP propagating direction. On the other hand, by using spatial coupled mode theory and FDTD simulations, we show the LD is associated with the linewidth of the reflectivity profile in Fourier-space. Therefore, by mapping out the equi-energy contour for a given wavelength, the momentum-deponent LD can be determined accordingly. We find both the decay rate and length are strongly dependent on the propagation direction of SPPs. In particular, they are strongly modified when two degenerate modes undergoing coherent coupling, yielding nontrivial and distinct vg for two dark and bright modes. We believe vg is mostly governed by neff. Dark and bright modes experience different neff and thus vg due to their field symmetry.

7. Acknowledgment

This research was supported by the Chinese University of Hong Kong through RGC General Research Fund (402812 and 403310), Direct Research Grant 4053077, Collaborative Research Fund CUHK1/CRF/12G, Area of Excellence AoE/P-02/12, and Innovative Technology Fund Tier 3 (ITS/196/13).

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. S. I. Bozhevolnyi, V. S. Volkov, E. Devaux, J.-Y. Laluet, and T. W. Ebbesen, “Channel plasmon subwavelength waveguide components including interferometers and ring resonators,” Nature 440(7083), 508–511 (2006). [CrossRef]   [PubMed]  

3. T. J. Seok, A. Jamshidi, M. Kim, S. Dhuey, A. Lakhani, H. Choo, P. J. Schuck, S. Cabrini, A. M. Schwartzberg, J. Bokor, E. Yablonovitch, and M. C. Wu, “Radiation engineering of optical antennas for maximum field enhancement,” Nano Lett. 11(7), 2606–2610 (2011). [CrossRef]   [PubMed]  

4. Z. Cao, L. Zhang, C.-Y. Chan, and H.-C. Ong, “Interplay between absorption and radiative decay rates of surface plasmon polaritons for field enhancement in periodic arrays,” Opt. Lett. 39(3), 501–504 (2014). [CrossRef]   [PubMed]  

5. G. Isić and R. Gajić, “Lifetime and propagation length of light in nanoscopic metallic slots,” J. Opt. Soc. Am. B 31(2), 393–399 (2014). [CrossRef]  

6. G. Isić, R. Gajić, and S. Vuković, “Plasmonic lifetimes and propagation lengths in metallodielectric superlattices,” Phys. Rev. B 89(16), 165427 (2014). [CrossRef]  

7. J. G. Pedersen, S. Xiao, and N. A. Mortensen, “Limits of slow light in photonic crystals,” Phys. Rev. B 78(15), 153101 (2008). [CrossRef]  

8. D. S. Kim, S. C. Hohng, V. Malyarchuk, Y. C. Yoon, Y. H. Ahn, K. J. Yee, J. W. Park, J. Kim, Q. H. Park, and C. Lienau, “Microscopic origin of surface-plasmon radiation in plasmonic band-gap nanostructures,” Phys. Rev. Lett. 91(14), 143901 (2003). [CrossRef]   [PubMed]  

9. C. Ropers, D. J. Park, G. Stibenz, G. Steinmeyer, J. Kim, D. S. Kim, and C. Lienau, “Femtosecond light transmission and subradiant damping in plasmonic crystals,” Phys. Rev. Lett. 94(11), 113901 (2005). [CrossRef]   [PubMed]  

10. D. Y. Lei, J. Li, A. I. Fernández-Domínguez, H. C. Ong, and S. A. Maier, “Geometry dependence of surface plasmon polariton lifetimes in nanohole arrays,” ACS Nano 4(1), 432–438 (2010). [CrossRef]   [PubMed]  

11. Z. Cao, H.-Y. Lo, and H.-C. Ong, “Determination of absorption and radiative decay rates of surface plasmon polaritons from nanohole array,” Opt. Lett. 37(24), 5166–5168 (2012). [CrossRef]   [PubMed]  

12. L. Langguth, D. Punj, J. Wenger, and A. F. Koenderink, “Plasmonic band structure controls single-molecule fluorescence,” ACS Nano 7(10), 8840–8848 (2013). [CrossRef]   [PubMed]  

13. K. G. Lee, X. W. Chen, H. Eghlidi, P. Kukura, R. Lettow, A. Renn, V. Sandoghdar, and S. Gotzinger, “A planar dielectric antenna for directional single-photon emission and near-unity collection efficiency,” Nat. Photonics 5(3), 166–169 (2011). [CrossRef]  

14. Z. L. Cao and H. C. Ong, “Determination of coupling rate of light emitter to surface plasmon polaritons supported on nanohole array,” Appl. Phys. Lett. 102(24), 241109 (2013). [CrossRef]  

15. C. Y. Chan, Z. L. Cao, and H. C. Ong, “Study of coupling efficiency of molecules to surface plasmon polaritons in surface-enhanced Raman scattering (SERS),” Opt. Express 21(12), 14674–14682 (2013). [CrossRef]   [PubMed]  

16. P. Y. Chen, R. C. McPhedran, C. M. de Sterke, C. G. Poulton, A. A. Asatryan, L. C. Botten, and M. J. Steel, “Group velocity in lossy periodic structured media,” Phys. Rev. A 82(5), 053825 (2010). [CrossRef]  

17. Z. L. Cao and H. C. Ong, “Direct imaging of radiative decay of surface plasmon polaritons in nanohole arrays by cross-polarization microscopy,” Appl. Phys. Lett. 102(9), 093108 (2013). [CrossRef]  

18. A. Drezet, A. Hohenau, A. L. Stepanov, H. Ditlbacher, B. Steinberger, N. Galler, F. R. Aussenegg, A. Leitner, and J. R. Krenn, “How to erase surface plasmon fringes,” Appl. Phys. Lett. 89(9), 091117 (2006). [CrossRef]  

19. C. J. Regan, A. Krishnan, R. Lopez-Boada, L. Grave de Peralta, and A. A. Bernussi, “Direct observation of photonic Fermi surfaces by plasmon tomography,” Appl. Phys. Lett. 98(15), 151113 (2011). [CrossRef]  

20. A. Angelini, E. Enrico, N. De Leo, P. Munzert, L. Boarino, F. Michelotti, F. Giorgis, and E. Descrovi, “Fluorescence diffraction assisted by Bloch surface waves on a one-dimensional photonic crystal,” New J. Phys. 15(7), 073002 (2013). [CrossRef]  

21. R. Wagner and F. Cichos, “Fast measurement of photonic stop bands by back focal plane imaging,” Phys. Rev. B 87(16), 165438 (2013). [CrossRef]  

22. A. W. Sanders, D. A. Routenberg, B. J. Wiley, Y. Xia, E. R. Dufresne, and M. A. Reed, “Observation of Plasmon Propagation, Redirection, and Fan-Out in Silver Nanowires,” Nano Lett. 6(8), 1822–1826 (2006). [CrossRef]   [PubMed]  

23. J. Li, H. Iu, D. Y. Lei, J. T. K. Wan, J. B. Xu, H. P. Ho, M. Y. Waye, and H. C. Ong, “Dependence of surface plasmon lifetimes on the hole size in two-dimensional metallic arrays,” Appl. Phys. Lett. 94(18), 183112 (2009). [CrossRef]  

24. J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, Photonic Crystals: Molding the Flow of Light, 2nd ed. (Princeton University, 2008).

25. C. Kittel, Introduction to Solid State Physics, 8th ed. (John Wiley and Sons, 2005).

26. H.-Y. Lo and H.-C. Ong, “Decay rates modification through coupling of degenerate surface plasmon modes,” Opt. Lett. 37(13), 2736–2738 (2012). [CrossRef]   [PubMed]  

27. S. R. K. Rodriguez, A. Abass, B. Maes, O. T. A. Janssen, G. Vecchi, and J. Gómez Rivas, “Coupling Bright and Dark Plasmonic Lattice Resonances,” Phys. Rev. X 1(2), 021019 (2011). [CrossRef]  

28. H. A. Haus, Waves And Fields In Optoelectronics (Prentice-Hall, Englewood Cliffs, 1984).

29. C. Sauvan, C. Billaudeau, S. Collin, N. Bardou, F. Pardo, J.-L. Pelouard, and P. Lalanne, “Surface plasmon coupling on metallic film perforated by two-dimensional rectangular hole array,” Appl. Phys. Lett. 92(1), 011125 (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 (7)

Fig. 1
Fig. 1 (a) The FDTD unit cell for 1D groove grating. (b) The simulated reflectivity as a function of incident wavevector for different wavelengths indicating (−1,0) SPP mode is excited at the reflection dips. (c) The FDTD supercell for 1D groove grating. (d) The spatial profile of the scattered near field for different wavelengths in logarithmic scale, showing exponentially decrease of intensity. (e) The plot of propagation length against resonant wavelength deduced from two methods.
Fig. 2
Fig. 2 (a) The p-polarized wavevector-resolved reflectivity mapping of 2D Au nanohole array taken in Γ-X direction. The dashed line is calculated by using phase-matching equation, indicating the excitation of (−1,0) SPP mode. Inset: the SEM image of the nanohole array. The scale bar is 500nm. (b) The corresponding reflectivity spectra for different incident wavevectors. The dashed lines are the best fits. (c) The corresponding reflectivity plots for different photon energies. The dashed lines are the best fits. (d) The deduced group velocities for different resonant wavelength. The dot line is ∂ω/∂k from the mapping.
Fig. 3
Fig. 3 (a) The schematic setup for real- and Fourier-space microscopy. (b) The plot of reflectivity against incident angle from 2D Au nanohole array for λ = 633 nm obtained from the microscopy. The dashed line is extracted from the goniometer for comparison. Three crosses are located at off- and on- SPP resonance regions labeled as (c)-(e). (c) & (e) The off-resonance real images. The hexagon indicates the region illuminated by light. The solid line indicates the region extracted for examining the spatial decay of light. (d) The on-resonance real image featuring long light tails extending around the hexagon in a particular direction. The arrows define the directions of incident light and SPP propagation. (f) The extracted spatial decay of light from (c)-(e) in logarithmic scale. Only the on-resonance decay shows exponential behavior. (g) The measured propagation length as a function of azimuthal angle.
Fig. 4
Fig. 4 (a) The unpolarized Fourier reflection image taken at λ = 633 nm from 2D Au nanohole array taken by microscopy. The solid and dashed lines indicate the SPP modes and Wood’s anomalies calculated from the phase-matching equations. The dot line defines the first Brillouin zone together with the reduced zone given by Γ-X-M triangle. (b) The unpolarized k-space mapping at λ = 633 nm from 2D Au nanohole array taken by angle-resolved goniometry. (c) Several (−1,0) SPP cuts taken at different azimuthal angles are extracted from the unpolarized Fourier image. The dashed lines are the best-fits for determining the corresponding propagation lengths. (d) The deduced (■) propagation length as a function of azimuthal angle. The positions of dark and bright modes and the Wood’s anomaly are also indicated. The (●) LD obtained from the real-space imaging in Fig. 3(g) is overlaid for comparison.
Fig. 5
Fig. 5 (a) The experimental total decay rate Γtot of (−1,0) SPPs plotted as a function of azimuthal angle φ. The solid line indicates the emergence of Wood’s anomaly. (b) The FDTD simulated (▲) total decay Γtot, (■) radiative decay Γrad, and (●) absorption Γabs rates of (−1,0) SPPs as a function of azimuthal angle φ. At φ = 45°, two (−1,0) and (0,-1) SPP modes couple, leading to the splitting of the rates.
Fig. 6
Fig. 6 (a) The plots of vg of (−1,0) SPPs as a function of azimuthal angle φ. At φ = 45°, two (−1,0) and (0,-1) SPP modes couple, leading to the splitting of vg. (b) The unpolarized incident angle-resolved reflectivity mapping taken at φ = 45°, indicating the presence of dark and bright modes. The dispersions are weak.
Fig. 7
Fig. 7 The plane-view field patterns |E| and the Poynting vector maps for (a) bright and (b) dark modes simulated by FDTD for λ = 633 nm at φ = 45°.

Equations (3)

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

R P ω ( k // )= | r p ω + χ ω κ ω i( k // +2nπ/P k spp )+( 1/2 L D ) | 2 ,
k spp = k // + 2nπ P x ^ + 2mπ P y ^ =( ω c sinθcosφ+ 2nπ P ) x ^ +( ω c sinθsinφ+ 2mπ P ) y ^ ,
d dt [ a 1 a 2 ]=[ i ω o Γ tot /2 i ω 12 i ω 21 i ω o Γ tot /2 ][ a 1 a 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.