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

Spontaneous emission enhancement by rotationally-symmetric optical nanoantennas: impact of radially and axially propagating surface plasmon polaritons

Open Access Open Access

Abstract

The excitation and radiation properties of rotationally-symmetric optical nanoantennas are independent of the azimuth angle, which enables great convenience and superior performances in practical applications. However, for rotationally-symmetric nanoantennas, the physical mechanisms behind their resonance properties remain to be clarified. In this paper, firstly, for a simple single-nanocylinder-on-mirror antenna (S-antenna), we establish a first-principles-based semianalytical model of surface plasmon polariton (SPP) by considering an intuitive multiple-scattering process of the radially-propagating gap surface plasmon (RGSP) in the nanogap and the axially-propagating surface plasmon (ASP) on the nanocylinder. The model can comprehensively reproduce all the radiation properties of the S-antenna such as the total and radiative emission rates, SPP excitation rates, and far-field radiation pattern. The model indicates that when the antenna radius is small (respectively, large), the enhancement of spontaneous emission mainly results from the resonance of ASP (respectively, RGSP). To show the wide applicability of the SPP model along with its unveiled decisive role of the RGSP and ASP in the spontaneous emission enhancement for other rotationally-symmetric nanoantennas of cylindrical shapes, we extend the SPP model to a more complex ring-nanocylinder-on-mirror antenna (R-antenna) that supports two ASPs. Moreover, to provide an explicit explanation of the resonance properties of the R-antenna, we further establish a semianalytical model for the resonant modes (called quasinormal modes, QNMs) supported by the R-antenna based on the SPP model, which quantitatively reveals the role of the RGSP and ASP in forming the antenna resonant modes and the resultant enhancement of spontaneous emission.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Owing to the support of surface plasmon polariton (SPP), optical nanoantennas can localize the electromagnetic field within a nanoscale region far beyond the diffraction limit [1]. This enhances the light-matter interaction with an enhancement of electromagnetic field [2] and spontaneous emission [37], and has important applications in nano-optoelectronics [3,810], quantum information processing [11,12] and so on.

At present, optical nanoantennas with rotational symmetry, such as circular patch antennas [46,12,13], concentric ring antennas [1418], etc. have attracted extensive attention. Under the illumination of a far-field uniform plane wave, the enhancement factor of electromagnetic field is only related to the elevation angle of the incident direction, but not to the azimuth angle because of the rotational symmetry of the antennas, which provides great convenience for practical applications, such as solar cells [19], refractive index sensing [20,21], photodetection [22,23] and surface enhanced Raman scattering [13,24]. In addition, rotationally-symmetric nanoantennas illuminated by a radially-polarized cylindrical vector beam can achieve an optimal focusing of SPP [2527]. Reciprocally, for coupled emitters such as molecules [17,18] or quantum dots [4,5], the rotational symmetry of the antenna resonant modes enables the rotational symmetry of the far-field radiation pattern of emitters [46,11,15,17,18,2830]. This is beneficial for detecting and improving the objective collection efficiency of the radiation field [46,11,15,17,18,2830], for instance, for single-photon sources [5,11,15] and single-molecule fluorescence detection [17,18].

Optical antennas with metallic nanogaps can achieve a drastic enhancement of electromagnetic field and spontaneous emission rate due to the support of gap SPPs with nanometric mode volumes [7,3134]. Compared with the uncontrollable size of nanogap between nanoparticles [31,32], the nanoparticle-on-mirror (NPoM) antennas [7,33,34] have unique advantages of easy and accurate formation of metallic nanogaps down to sub-nanometer sizes. A combination of the rotationally-symmetric antennas and the NPoM antennas can combine the advantages of these two kinds of antennas [4,5,24,28,29,35,36].

At present, there have been rich theoretical and experimental works on the rotationally-symmetric nano-antennas, which are devoted to an understanding of the resonance-induced enhancement of electromagnetic field and spontaneous emission for guiding the design of such antenna devices. Theoretically, the resonant frequencies, resonant modes [16,37] and the radiation and scattering properties [3840] of the rotationally-symmetric nanoantennas with arbitrary shapes can be calculated with the numerical solvers of Maxwell's equations but with the lack of physical intuitiveness. For antennas with special shapes, such as spheres [4144], cylinders [20,4547], slender body [48,49], and torus [50,51], the resonant frequencies [20,4151], resonant modes [4244,4649,51] and scattering properties [20,41,4850] can be analytically solved with the method of separation of variables and sometimes under the electrostatic approximation [20,41,4851]. For composite structures composed of elementary structures of special shapes, such as concentric nanoshells [52,53], torus [54] and nanosphere-on-mirror antennas [55], their resonant frequencies can be obtained analytically with the plasmon hybridization model [5255]. The above analytical formalisms can describe the macroscopic resonance properties of some antennas with special shapes. At a microscopic level, the antenna resonance should be inherently related to SPPs defined as guided modes excited in the antenna [35,36,5658]. In this aspect, for a multilayer-stacked circular patch antenna, a Bessel-type standing-wave resonator model is established by regarding the antenna as a Fabry-Perot resonator of the radially-propagating surface plasmon (RSP) [56]. The model can predict the resonant condition of the RSP, and provides an approximate analytical expression for the reflection coefficient of the RSP at the edge of the patch antenna. Different from the circular patch antennas, the resonance of cylindrical antennas with a certain height is not only related to the resonance of RSP, but also related to the resonance of axially-propagating surface plasmon (ASP) [35,36,57]. For cylindrical dipole antennas [57] and nanocylinder-on-mirror antennas [35,36], the independence [57] or hybridization [35,36] between the RSP-related radial gap mode in the nanogap and the ASP-related axial antenna mode on the antenna arm are investigated numerically, showing their respective contributions in reducing the mode volume and enhancing the far-field radiation. However, to the best of our knowledge, analytical-level explorations of the impact of both the RSP and the ASP on the resonance properties of rotationally-symmetric nanoantennas, which are crucial for a transparent understanding and efficient design of such antenna devices, are still lacking.

In experiment, the development of microscopic imaging technologies renews our practical knowledge on the resonant modes and the resultant enhancement effects of rotationally-symmetric nanoantennas. The spatial distribution of resonant modes of rotationally-symmetric nanoantennas, such as circular-grating antennas [59], circular patch and ring antennas [37], has been measured by scanning near-field optical microscope (SNOM). In addition, electron energy loss spectroscopy (EELS) [6063] and cathodoluminescence (CL) imaging spectroscopy [28,61,64,65] have been used to study the resonant frequencies and the field distribution of the resonant modes of metallic circular patch antennas, with a spatial resolution reaching several nanometers [60]. EELS signal corresponds to the total local density of states (LDOS) of electromagnetic field, so it can be used to distinguish the radial resonant modes (breathing modes) and edge modes supported by circular patch nanoantennas [60,61], and to measure the dispersion curves of these two modes [63]. In contrast, the CL measurement corresponds to the radiative part of the LDOS, and therefore can achieve an imaging of the radial resonant modes only when the antenna radius is large [28,61,65]. All these experimental advances have raised a huge demand on a theoretical understanding of the formation of the resonant modes in rotationally-symmetric nanoantennas.

In this paper, we will build up semianalytical models aiming at clarifying the role of the RSP and ASP in forming the resonance and the resultant enhancement of spontaneous emission for rotationally-symmetric nanoantennas especially of cylindrical shapes. We study a nanocylinder-on-mirror antenna with a certain height [35,36], which combines the advantages of rotationally-symmetric antennas [6,13,56,65] and NPoM antennas [7,33,34]. The circular patch antenna with a height of only tens of nanometers on metallic substrate [4,5,28,29] can be seen as a special case of the nanocylinder-on-mirror antenna. Firstly, we consider an elementary structure, namely a single-nanocylinder-on-mirror antenna (S-antenna). We establish a semianalytical SPP model by considering an intuitive multiple-scattering process of the radially-propagating gap surface plasmon (RGSP) in the nanogap and the ASP on the nanocylinder. The model can comprehensively and quantitatively predict all the radiation properties of the antenna (such as enhancement factors of total and radiative emission rates, SPP excitation rates and far-field radiation pattern). Different from the Bessel-type standing-wave resonator model [56] that only considers the resonance of RSP, our model further considers the coupling between the RGSP and ASP, and shows that the ASP resonance plays a dominant role in the enhancement of spontaneous emission when the antenna radius is small. With the increase of the antenna radius, the ASP resonance gradually weakens, and the RGSP resonance gradually plays a major role. Based on this model, two phase-matching conditions can be derived to predict the resonance of the RGSP and ASP, respectively. Besides, all physical quantities (such as the SPP scattering coefficients) used in the SPP model are rigorously calculated based on the first principles of Maxwell’s equations, without any fitting or artificial setting of model parameters, which ensures that the model has a solid electromagnetic foundation and thus can provide quantitative predictions.

To show the wide applicability of the SPP model along with its unveiled decisive role of the RGSP and ASP in forming the resonance and the resultant spontaneous emission enhancement for other rotationally-symmetric nanoantennas of cylindrical shapes, we further extend the SPP model to a more complex ring-nanocylinder-on-mirror antenna (R-antenna) that supports two ASPs. Moreover, to provide an explicit explanation of the resonance properties of the R-antenna, we further establish a semianalytical model for the resonant modes (termed quasinormal modes, QNM) [58,6669] supported by the R-antenna based on the SPP model. Compared with previous works in which the QNMs are commonly obtained via a full-wave numerical solution of source-free Maxwell’s equations [35,66,67,70], our QNM model can provide analytical expressions for the complex eigenfrequencies and electromagnetic fields of the antenna QNMs using the RGSP and ASP modes and their scattering coefficients, thus quantitatively revealing the role of the RGSP and ASP in forming the antenna resonant modes and the resultant enhancement of spontaneous emission. In experimental studies of the rotationally-symmetric antennas of cylindrical shapes [28,37,6065], it is difficult to quantify the contributions of different types of SPPs to the measured field of resonant modes. For this problem, the proposed QNM model can be a useful tool for an in-depth analysis.

This article is organized as follows. In Sec. 2, the SPP model is established and used to analyze the spontaneous emission properties of the S-antenna. In Sec. 3, the SPP model is extended to the R-antenna, and the QNM model is further established. Conclusions are summarized in Sec. 4.

2. Analysis of single-nanocylinder-on-mirror antenna (S-antenna)

2.1 Rigorous data for the spontaneous emission enhancement by the S-antenna

In this section, we consider the elementary S-antenna. As shown in Fig. 1(a1), the S-antenna is composed of a gold nanocylinder (radius R, length L) located on a gold substrate, separated by a nanogap of polymethylmethacrylate (PMMA) dielectric layer (refractive index nd = 1.5, thickness w), and in an environment of air. A z-polarized electric-current point source, which represents a fluorescent molecule [17,18] or quantum dot [35], is placed at the center of the nanogap. The coordinate origin O is set at the position of the point source. The refractive index of gold at different wavelengths can be obtained from the experimental data [71].

 figure: Fig. 1.

Fig. 1. (a1) Sketch of the S-antenna excited by a z-polarized point source (red dot) at the center the PMMA nanogap. (a2) Definitions of the unknown SPP coefficients a, b, c, d, p to be solved in the SPP model. (b1) Distribution of the main electric-field component |Ez| of the fundamental RGSP mode. (b2) Distribution of the main electric-field component |Eρ| of the fundamental ASP mode. (b3) Distribution of the main electric-field component |Ez| of the fundamental RISP mode, with the insert showing an enlarged view in the PMMA layer. The vertical red dashed lines in (b1)-(b3) show the interfaces between different media. The results in (b) are obtained with the a-FMM at wavelength λ=0.7µm for a PMMA-layer thickness w = 10 nm and an antenna radius R = 82 nm.

Download Full Size | PDF

To obtain rigorous data of the radiation of the antenna, we use a full-wave aperiodic Fourier modal method (a-FMM) [72] extended to rotationally-symmetric structures [73,74]. The z-polarized point source in the nanogap can be expressed as an electric-current density J = zδ(x,y,z), with δ the Dirac function, and z the unit vector along the z direction. Under the cylindrical coordinate system (see the inset in Fig. 1), the point source can be approximately expressed as a circular line source with a minimal radius ρδ, i.e., J = z(2πρδ)−1δ(ρρδ)δ(z) (see Supplement 1 Sec. S1B, Fig. S2). The total spontaneous emission rate of the point source is proportional to the total emission power [75] and is expressed as [76] Γtot = −Re(Ez,source)/2, where Re(Ez,source) is the real part of the z-component of the electric-field vector at the position of the point source (or the circular line source). Γtot contains a radiative emission rate $\Gamma_\textrm {rad}$ into the far field and a nonradiative rate Γnr due to the ohmic loss of the metal. Since the S-antenna has a gold substrate, Γrad contains not only a radiative emission rate $\Gamma^{\prime}_\textrm {rad}$ of photons into the free space, but also an excitation rate ΓRISP of the radially-propagating interface surface plasmon (RISP) [7780] along the gold substrate surface outside the antenna. The field distribution of the RISP mode is shown in Fig. 1(b3), and its analytical expression is given by Eq. (1) in Sec. 2.2. The radiative emission rate Γrad can be expressed as ${\Gamma _{\textrm{rad}}} = \mathop{{\oiint}}\nolimits_A {\textbf{S} \cdot \textbf{n}da}$, where A is a closed surface near the antenna and encompassing the antenna and the point source, S is the time-averaged Poynting vector of the electromagnetic field excited by the point source, and n is the out-pointing unit normal vector on A. The photon-radiation emission rate can be calculated as an integral ${\Gamma ^{\prime}_{\textrm{rad}}}\textrm{ = }\int\!\!\!\int_C {\textbf{S} \cdot \textbf{n}dxdy}$, where C is an infinitely-large plane above the antenna (z coordinate being constant) and n = z. The excitation rate of RISP can be expressed as ${\Gamma _{\textrm{RISP}}} = {|p |^2}\int\!\!\!\int_\Sigma {{\textbf{S}_{\textrm{RISP}}} \cdot \textbf{n}ds}$, where Σ is an infinitely-high cylindrical surface encompassing the antenna (near and coaxial with the antenna), p is the complex amplitude coefficient of the RISP [as shown in Fig. 1(a2)], which can be calculated with the mode orthogonality theorem [7981] or reciprocity theorem [58,79,81] (see Supplement 1 Sec. S3A for details), SRISP is the time-averaged Poynting vector of the RISP with unit coefficient, and n is the out-pointing unit normal vector on Σ.

To characterize the enhancement of spontaneous emission rate of fluorescence emitter, γtottotair, γradradair, $\gamma^{\prime}_\textrm {rad}$= $\Gamma^{\prime}_\textrm {rad}$air and γRISPRISPair are defined as the enhancement factors of the total emission rate, radiative emission rate, photon-radiation emission rate and the excitation rate of RISP, respectively, where Γair=ηvac$k^{2}_{0}$na/(12π) is the total emission rate of the point source in the air (k0 = 2π/λ with λ being the wavelength in the vacuum, na = 1 being the refractive index of air, ηvac being the wave impedance in the vacuum). A high value of γtot, also known as the Purcell factor [3,82], can shorten fluorescence lifetime and form nanoscale light sources with high modulation rate, such as light-emitting diode [3,9], nanolaser [8,10] or single-photon source [3,5,11,15] applied in quantum information processing. For surface-enhanced fluorescence [17,18,31,33] or Raman-scattering [13,24] molecular sensing, a high value of the quantum yield of the emitter ηq=$\gamma^{\prime}_\textrm {rad}$/($\eta^{-1}_0$−1+γtot) is required to enhance the fluorescence intensity, where η0∈(0,1) represents the intrinsic quantum yield of the emitter. For the emitter with η0 = 1 (such as quantum dots [3]), the quantum yield can be simplified as ηqη′, where η′=$\gamma^{\prime}_\textrm {rad}$/γtot is defined as the photon-radiation efficiency of the antenna. For emitters with a very low η0 (such as some fluorescent molecules [31]), the quantum yield can be simplified as ηq$\gamma^\prime_\textrm{rad}$η0, proportional to $\gamma^{\prime}_\textrm {rad}$. The radiation efficiency and RISP excitation efficiency of the antenna are defined as η=γrad/γtotradtot and ηRISP=γRISP/γtotRISPtot, respectively. For on-chip integrated SPP-based devices, increasing ηRISP is required to improve the signal intensity of the SPP source [35,80].

Next, we investigate the spontaneous emission enhancement and far-field radiation properties of the S-antenna based on rigorous numerical calculation. As shown in Figs. 2(a)-(c), γtot, $\gamma^{\prime}_\textrm {rad}$ and η′ of the S-antenna obtained with the full-wave a-FMM are plotted as functions of the antenna radius R and height L at a fixed excitation wavelength λ=0.7µm for a thickness w = 10nm of the PMMA layer. According to the results, the antenna radius and height are respectively determined as R = 0.082µm and L = 0.368µm, for which $\gamma^{\prime}_\textrm {rad}$ = 892.3 takes the maximum value and η′=0.4459 is close to the maximum. Meanwhile, a directional radiation of the far field into a central angular zone can be realized [see Supplement 1 Sec. S4, Fig. S9(c)], which can improve the objective collection efficiency of the fluorescence [4,5,17,18,28].

 figure: Fig. 2.

Fig. 2. (a)-(c) Enhancement factors of total emission rate (γtot) and photon-radiation emission rate ($\gamma^{\prime}_\textrm {rad}$), and photon-radiation efficiency η′ plotted as functions of antenna radius R and length L. The results are obtained with the full-wave a-FMM at a fixed wavelength λ=0.7µm for a PMMA-layer thickness w = 10 nm.

Download Full Size | PDF

The impact of the refractive index nd of the nanogap on the radiation properties of the S-antenna is shown in Table 1, which is obtained with the full-wave a-FMM. For nd = 1.5 and 3.45 (i.e., GaAs) and fixed nanogap size w = 10 nm and wavelength λ=1µm, the R and L are optimized [corresponding to Q = 1 in Eq. (12) and P = 1 in Eq. (10)] so that $\gamma^{\prime}_\textrm {rad}$ takes the maximum value and η′ is close to the maximum. The results show that with the increase of nd, γtot and γRISP increase while γrad and $\gamma^{\prime}_\textrm {rad}$ decrease, along with a resultant decrease of the radiation efficiency η=γrad/γtot, photon-radiation efficiency η′=$\gamma^{\prime}_\textrm {rad}$/γtot and RISP excitation efficiency ηRISP=γRISP/γtot.

Tables Icon

Table 1. Impact of the refractive index nd of the nanogap on the radiation properties of the S-antenna.

2.2 SPP model of the S-antenna

To clarify the role of the RSP and ASP in forming the resonance and the resultant spontaneous emission enhancement of the S-antenna, here we will establish a first-principles-based semianalytical SPP model by considering the excitation and multiple scattering processes of the RSP and ASP in the S-antenna. Since the size of the nanogap is much smaller than the wavelength, only the fundamental RGSP mode is bounded (field decaying to null at infinity in transversal z direction) and propagative (propagation constant being almost real) [83] in the nanogap of the antenna, and thus is considered in the model. The contribution of all other higher-order modes can be neglected. The field distribution of the fundamental RGSP mode is shown in Fig. 1(b1). Since both the RGSP in the nanogap and the RISP outside the antenna propagate radially, they are all called RSP. The distribution of electric field E and magnetic field H of the out-going RSP can be expressed analytically as [7780],

$$ \mathbf{E}=(\boldsymbol{\mathrm{\rho}}, \boldsymbol{\mathrm{\varphi}}, \mathbf{z})\left[\begin{array}{c} \frac{i}{2}\left[H_{n+1}^{(1)}(k \rho)-H_{n-1}^{(1)}(k \rho)\right] E_{x, 0}(z) \\ \frac{1}{2}\left[H_{n+1}^{(1)}(k \rho)+H_{n-1}^{(1)}(k \rho)\right] E_{x, 0}(z) \\ H_{n}^{(1)}(k \rho) E_{z, 0}(z) \end{array}\right] \exp (i n \varphi) $$
$$ \mathbf{H}=(\boldsymbol{\mathrm{\rho}}, \boldsymbol{\mathrm{\varphi}}, \mathbf{z})\left[\begin{array}{c} \frac{-1}{2}\left[H_{n+1}^{(1)}(k \rho)+H_{n-1}^{(1)}(k \rho)\right] H_{y, 0}(z) \\ \frac{i}{2}\left[H_{n+1}^{(1)}(k \rho)-H_{n-1}^{(1)}(k \rho)\right] H_{y, 0}(z) \\ 0 \end{array}\right] \exp (i n \varphi) $$
where (ρ,φ,z) are cylindrical coordinates [as shown in the inset in Fig. 1(a)], (ρ,φ,z) are the unit vectors along ρ, φ and z directions, respectively, and k is the propagation constant of the RSP mode. $H^{(1)}_{n}$ denotes the Hankel function of the first kind, and n represents the angular quantum number (n = 0, ±1, ±2, …). In this paper, the emitter is located at ρ=0 of the rotationally-invariant structure, so that the radiation field is independent of φ, i.e., n = 0. For the case that the emitter is not at ρ=0, the RSPs with n≠0 should be considered. In view that the RSPs with different n cannot excite each other, the SPP model in this paper can be readily extended to the cases of n≠0. The dependence of the RSP field on the z coordinate is described by [Ex,0(z), Ez,0(z), Hy,0(z)] (with a normalization of Ez,0(0) = 1), which correspond to the field distribution of the mode (propagating along x direction with a propagation constant of k) of the waveguide with the same structure (stratified in z direction and invariant in y direction) under the Cartesian coordinate system. For the RGSP, k = kRGSP and [Ex,0(z), Ez,0(z), Hy,0(z)] can be calculated rigorously with the full-wave a-FMM [7274] or approximately with analytical expressions [83]. Similarly, for the RISP, k = kRISP and [Ex,0(z), Ez,0(z), Hy,0(z)] can also be calculated with the a-FMM. For the in-going RSP, the $\left[H_{n-1}^{(1)}, H_{n}^{(1)}, H_{n+1}^{(1)}\right]$ in Eq. (1) are simply replaced by the Hankel functions of the second kind $\left[H_{n-1}^{(2)}, H_{n}^{(2)}, H_{n+1}^{(2)}\right]$.

Since the nanocylinder diameter is much smaller than the wavelength, only the fundamental ASP mode is bounded [84] (field decaying to null at infinity in transversal ρ direction) and propagative on the nanocylinder, and thus is considered in the model. The contribution of all other higher-order modes can be neglected. The field distribution of the fundamental ASP mode is shown in Fig. 1(b2).

To establish the SPP model, firstly, we consider the multiple scattering process of the fundamental ASP mode. As shown in Fig. 1(a2), c and d represent the unknown coefficients of up-going and down-going ASPs, respectively, which satisfy a normalization of Ez(ρ=R) = 1 at the lower and upper terminations of the antenna arm, respectively (i.e., on the cross-sections where the ASPs begin to propagate). To solve the c and d, a set of coupled-ASP equations can be written,

$$c = \beta + du{\rho _\textrm{c}},$$
$$d = cur,$$
where u = exp(ik0neff,AL) is the phase-shift factor of the ASP accumulated over the antenna arm (k0= 2π/λ, neff,A being the dimensionless complex effective index of the ASP mode). The ASP scattering coefficients β, ρc and r in Eq. (2) are defined in Figs. 3(a)-(c), which can be calculated as scattering matrix elements [85] with the full-wave a-FMM under cylindrical coordinate system [7274,86] (see Supplement 1 Sec. S1A for more details of the calculation method). Equation (2) can be understood intuitively. For Eq. (2a), the coefficient c of the up-going ASP results from two contributions: the first one [β, see Fig. 3(a)] from the direct excitation of the source in the nanogap; the second one from the reflection [ρc, see Fig. 3(b)] of the damped (u) down-going ASP (with coefficient d) at the substrate. Equation (2b) can be understood similarly. Solving Eq. (2), one can obtain,
$$c = \frac{\beta }{{1 - {u^2}{\rho _\textrm{c}}r}},$$
$$d = \frac{{\beta ur}}{{1 - {u^2}{\rho _\textrm{c}}r}}.$$

 figure: Fig. 3.

Fig. 3. Definition of the derived scattering coefficients and electromagnetic fields used in the SPP model for the S-antenna. (a) For a semi-infinite gold nanocylinder on a gold substrate excited by a z-polarized point source in the nanogap, β denotes the excitation coefficient of up-going ASP, as and bs are the coefficients of out-going and in-going RGSPs, respectively, ps is the coefficient of out-going RISP, and Ψs denotes the field excited by the point source. (b) For the structure of (a) but illuminated by an incident down-going ASP with unit coefficient, ρc is the reflection coefficient of the ASP at the substrate, ac and bc are the coefficients of out-going and in-going RGSPs, respectively, pc is the coefficient of out-going RISP, and Ψc denotes the field excited by the down-going ASP. (c) Reflection coefficient r of the ASP at the top termination of the semi-infinite gold nanocylinder.

Download Full Size | PDF

Next, we will derive analytical expressions of the scattering coefficients β and ρc in Eq. (3) by further considering the coupling between the RGSP and ASP, so as to quantitatively provide the contribution of the RGSP to the antenna radiation. To derive the analytical expression of β, a set of coupled-SPP equations can be written for the scattering process in Fig. 3(a),

$$a_\textrm{s} = \beta_{0} + b_\textrm{s}r_0,$$
$${b_\textrm{s}} = {a_\textrm{s}}{r_1},$$
$$\beta = {a_\textrm{s}}\alpha ,$$
where as and bs are coefficients of the out-going and in-going RGSPs in the nanogap, respectively. The scattering coefficients β0, r0, r1 and α are defined in Figs. 4(a1)-(a2) and (b1). They describe the single excitation or scattering process of the RGSP, and thus are called elementary scattering coefficients. These elementary scattering coefficients are calculated rigorously based on the first principles of Maxwell’s equations (see Supplement 1 Sec. S1A for the calculation method) and can self-consistently satisfy the reciprocity theorem (see Supplement 1 Sec. S1B). Different from these elementary scattering coefficients, β and ρc in Eq. (3) contain the multiple scattering processes of the RGSP and thus are called derived scattering coefficients, which can be further expressed in terms of the elementary scattering coefficients [see Eqs. (5c) and (7c) below]. Equation (4) can be understood intuitively. For Eq. (4a), the coefficient as of the out-going RGSP results from two contributions: the first one (β0) from the direct excitation by the source in the nanogap; the second one from the reflection (r0) of the in-going RGSP (coefficient bs) at the center of the nanogap. Equation (4b) can be understood similarly. For Eq. (4c), the excitation coefficient (β) of the ASP by the point source results from the excitation (α) of the out-going RGSP (coefficient as). Different from Eq. (2), Eq. (4) does not include the phase-shift factor of the RGSP, because it has been included in the analytical expression of the RGSP field [see Eq. (1)]. Solving Eq. (4), one can obtain,
$${a_\textrm{s}}\textrm{ = }\frac{{{\beta _0}}}{{\textrm{1} - {r_0}{r_1}}},$$
$${b_\textrm{s}} = \frac{{{\beta _0}{r_1}}}{{\textrm{1} - {r_0}{r_1}}},$$
$$\beta = \frac{{{\beta _0}\alpha }}{{\textrm{1} - {r_0}{r_1}}}.$$

 figure: Fig. 4.

Fig. 4. Definition of the elementary scattering coefficients and electromagnetic field used in the SPP model for the S-antenna. (a1) For the point source in the MIM nanogap, β0 denotes the excitation coefficient of the out-going RGSP, and Ψs,0 represents the excited electromagnetic field. (a2) r0 denotes the reflection coefficient of the in-going RGSP at the center of the MIM nanogap. (b1) For a semi-infinite gold nanocylinder on a gold substrate illuminated by an incident out-going RGSP with unit coefficient, r1 is the coefficient of the reflected in-going RGSP at the termination of the gap, α and α0 are the coefficients of the excited up-going ASP and out-going RISP, respectively. (b2) For the structure in (b1) but under illumination by a down-going ASP with unit coefficient, α′ and χ denote the excitation coefficients of the in-going RGSP and the out-going RISP, respectively, and ρ0 is the elementary reflection coefficient of the ASP. Note that ρ0 is different from the ρc defined in Fig. 3(b): ρc contains the contribution of the ASP excited by the out-going RGSP (with coefficient ac), but ρ0 does not.

Download Full Size | PDF

Similarly, to obtain the analytical expression of ρc, a set of coupled-SPP equations can be written by considering the scattering process in Fig. 3(b),

$${a_\textrm{c}} = {b_\textrm{c}}{r_0},$$
$${b_\textrm{c}} = \alpha ^{\prime} + {a_\textrm{c}}{r_1},$$
$${\rho _\textrm{c}} = {\rho _0} + {a_\textrm{c}}\alpha ,$$
where ac and bc are the coefficients of the out-going and in-going RGSPs, respectively. The scattering coefficients α′ and ρ0 are defined in Fig. 4(b2). They describe the single scattering process of the ASP and thus belong to the elementary scattering coefficients like the β0, r0, r1 and α defined earlier. Details for their calculation method can be found in Supplement 1 Sec. S1A. Equations (6a)–(6b) can be understood intuitively. For Eq. (6b), the coefficient bc of the in-going RGSP results from two contributions: the first one (α′) from the direct excitation of the down-going ASP; the second one from the reflection (r1) of the out-going RGSP (coefficient ac) at the termination of the gap. Equation (6a) can be understood similarly. For Eq. (6c), ρc contains not only the reflection (ρ0) of the down-going ASP at the substrate, but also the excitation (α) by the out-going RGSP (coefficient ac). Solving Eq. (6), one can obtain,
$${a_\textrm{c}} = \frac{{\alpha ^{\prime}{r_0}}}{{1 - {r_0}{r_1}}},$$
$${b_\textrm{c}} = \frac{{\alpha ^{\prime}}}{{1 - {r_0}{r_1}}},$$
$${\rho _\textrm{c}} = {\rho _0} + \frac{{\alpha ^{\prime}{r_0}\alpha }}{{1 - {r_0}{r_1}}}.$$

Similar to β and ρc, the as, bs, ac and bc in Eqs. (5) and (7) contain the multiple scattering process of the RGSP, and thus belong to the derived scattering coefficients as well.

The unknown RGSP coefficients a and b [see Fig. 1(a2)] can be obtained with the above solved d, as, bs, ac and bc,

$$a\textrm{ = }{a_\textrm{s}} + du{a_\textrm{c}},$$
$$b = {b_\textrm{s}} + du{b_\textrm{c}}.$$

The first and second terms in the right side of Eq. (8) represent the coefficients of the RGSPs excited by the point source and the down-going ASP, respectively.

Consequently, the electromagnetic field in the nanogap can be expressed as,

$${\boldsymbol{\Psi }_{\textrm{gap}}} = ({\boldsymbol{\Psi }_{\textrm{s,0}}} - {\beta _\textrm{0}}{\boldsymbol{\Psi }_{\textrm{RGSP, + }}}) + a{\boldsymbol{\Psi }_{\textrm{RGSP, + }}} + b{\boldsymbol{\Psi }_{\textrm{RGSP,} - }},$$
where Ψ=[E, H] represents both the electric-field vector E and the magnetic-field vector H. Ψs,0 represents the field excited by the point source in the metal-insulator-metal (MIM) nanogap as shown in Fig. 4(a1). ΨRGSP,+ and ΨRGSP,− represent the out-going and in-going fundamental RGSPs, respectively. According to Eqs. (5a) and (8a), the coefficient a in the right side of Eq. (9) has already contained the contribution of β0, so β0ΨRGSP,+ should be subtracted from the Ψs,0. With Eq. (9), the z component Ez,source of the electric field at the point-source position in the nanogap can be calculated, and then the total emission rate Γtot = −Re(Ez,source)/2 can be obtained.

In addition, by using the above SPP model, the electromagnetic field Ψrad in the free space outside the antenna can be obtained as well (see Eq. (S23) in Supplement 1 Sec. S2A), and then the radiative emission rate Γrad and the far-field radiation pattern of the antenna (see Fig. S(9c) in Supplement 1 Sec. S4) can be obtained. Moreover, by using the SPP model, the coefficient p and the resultant excitation rate ΓRISP of the RISP outside the antenna can be obtained (see Eq. (S33) in Supplement 1 Sec. S3B).

2.3 Analysis of the S-antenna with the SPP model

In this subsection, we will use the SPP model to analyze the impact of the RSP and ASP on the resonance and the resultant spontaneous emission enhancement of the S-antenna. Firstly, we will test the validity of the SPP model against full-wave a-FMM results. As shown in Figs. 5(a1)-(a3), the enhancement factors of total emission rate (γtottotair), radiative emission rate (γradradair), and excitation rate of RISP (γRISPRISPair) are plotted as functions of the antenna length L at wavelength λ=0.7µm with the antenna radius (R = 0.082µm) and the gap width (w = 10nm) determined in Sec. 2.1. The results show that the model predictions (red-solid curves) agree well with the full-wave a-FMM calculations (blue circles). This agreement confirms the validity of the model, and exhibits the dominant impact of the RGSP and ASP considered in the model on the antenna radiation. Meanwhile, the SPP model exhibits some error at the first resonant peak. This error implies that at the first resonant peak, besides the fundamental RGSP and ASP modes considered in the model, other neglected higher-order modes (called residual field here) also contribute to the antenna radiation [76].

 figure: Fig. 5.

Fig. 5. (a1)-(a3) Spontaneous-emission enhancement factors Γtotair, Γradair and ΓRISPair of the S-antenna plotted as functions of antenna length L. The results are obtained for R = 0.082µm, w = 10 nm and λ=0.7µm. The blue circles and red-solid curves show the full-wave a-FMM results and SPP model predictions, respectively. (b1)-(b3) Γtotair, Γradair and antenna radiation efficiency ηradtot predicted by the SPP model plotted as functions of the antenna radius R and length L. The vertical black-white lines show the L = Lres(R) at resonance determined by Eq. (10), corresponding to P = 0, 1, …, 5 from left to right. The horizontal black-white lines show the R = Rres at resonance determined by Eq. (12), corresponding to Q = 1, 2, …, 6 from bottom to top. (c1)-(c3) SPP scattering coefficients |β|, |ρc| and |r| plotted as functions of antenna radius R. The vertical red-dashed lines in (c1)-(c2) show the R = Rres. The results in (c) are obtained with the full-wave a-FMM at wavelength λ=0.7µm.

Download Full Size | PDF

Next, benefiting from the analyticity of the model, we will analyze the physical mechanism of the spontaneous emission enhancement of the antenna. The prediction of the SPP model is summarized in Fig. 5(b1)-(b3), which show the radiation enhancement factors γtottotair, γradradair and the radiation efficiency ηradtot plotted as functions of the antenna radius R and height L.

First of all, Figs. 5(b1)-(b2) show that when R is small (R < 0.081µm), the spontaneous emission enhancement of the antenna (i.e., γtot and γrad reaching the maximum much large than 1) strongly depends on L. In the following analysis, we will show that in this case, the spontaneous emission enhancement mainly comes from the resonance of the ASP. To this end, the fact that γtot reaches the maximum results from the fact that the RGSP coefficients |a| and |b| reach the maximum [see Eq. (9)], which further results from the fact that the down-going ASP coefficient |d| reaches the maximum [Eq. (8)] under the following phase-matching condition of the ASP [Eq. (3b)],

$$2{k_0}{\textrm{Re}} ({n_{\textrm{eff,A}}})L\textrm{ + }\arg ({\rho _\textrm{c}})\textrm{ + }\arg (r) = 2P\pi ,$$
where Re(neff,A) denotes the real part of the complex effective index neff,A of the ASP mode, arg() denotes the argument, and P is an integer corresponding to different orders of the ASP resonance. Equation (10) is obtained by minimizing the denominator in Eq. (3b) and with the following considerations. When R is small, the scattering coefficients |ρc| and |r| in the denominator of Eq. (3b) are smaller than and close to 1 [see Figs. 5(c2)-(c3)]. In addition, since the ASP mode is propagative, there is Im(neff,A)≈0 (see Fig. S(4a) in Supplement 1 Sec. S1B), so that there is |u|=exp[−k0Im(neff,A)L]≈1. Therefore, when L satisfies Eq. (10), the denominator of Eq. (3b) is equal to 1−|u|2|ρc||r| which is close to 0. Then the ASP coefficient |d| takes the maximum value, and is much larger than the |as|, |ac|, |bs| and |bc| in Eq. (8).

For a given R, L = Lres(R) determined by Eq. (10) are shown by the vertical black-white solid lines in Fig. 5(b). Lres(R) is dependent on R since the neff,A (see Fig. S(4a) in Supplement 1 Sec. S1B), ρc and r [see Figs. 5(c2)-(c3)] in Eq. (10) are all dependent on R. Figure 5(b1) shows that when R is small (R < 0.081µm), the position where Γtot reaches the maximum can be accurately predicted by the Lres, which confirms that the enhancement of spontaneous emission is mainly due to the resonance of the ASP.

Equation (10) shows that when the antenna length L = Lres, the phase shift accumulated by the ASP propagating over one round on the antenna arm [the left side of Eq. (10)] is multiples of 2π, which results in a constructive interference of the multiple-scattered ASPs, and thus forms a Fabry-Perot resonance of the ASP over the antenna arm. Equations (8)-(9) show that the resonantly enhanced ASP (with coefficient d) on the antenna arm further excites the RGSP (coefficients a, b), thus enhances the electric field at the point-source position, and finally enhances the total emission rate Γtot = −Re(Ez,source)/2. In addition, Fig. 5(b1) shows that the smaller R is, the smaller the difference between two adjacent resonance positions Lres [which is λ/[2Re(neff,A)] according to Eq. (10)] will be. This is due to the increase of Re(neff,A) with the decrease of R (see Fig. S4(a) in Supplement 1 Sec. S1B).

On the other hand, Figs. 5(b1)-(b2) show that when R is large (R≥0.081µm), the spontaneous emission enhancement of the antenna (i.e., γtot and γrad reaching the maximum much large than 1) mainly depends on R. The following analysis will show that in this case, the spontaneous emission enhancement mainly comes from the resonance of the RGSP. To this end, for γtot we notice that with the increase of R, the reflection coefficient |r| at the top termination of the antenna decreases rapidly [see Fig. 5(c3)], which will make the denominator of Eq. (3b) no longer close to 0 when L = Lres. Therefore, the maximum value of the ASP coefficient |d| decreases, i.e., the ASP resonance becomes weak. We thus infer that the spontaneous emission enhancement of the antenna is mainly due to the resonance of the RGSP in the nanogap, i.e., the RGSP coefficients |a| and |b| reaching the maximum mainly results from that the |as|, |bs|, |ac| and |bc| reach the maximum under the following phase-matching condition of the RGSP,

$$\arg ({r_0}) + \arg({r_1}) = 2Q\pi ,$$
where Q is an integer corresponding to different orders of the RGSP resonance. Equation (11) is obtained by minimizing the denominator in Eqs. (5a)–(5b) and (7a)–(7b). Since the RGSP is totally reflected at the center of the gap, i.e., r0 = 1, Eq. (11) can be simplified as arg(r1) = 2. r1 can be further expressed as follows. Considering an incident out-going RGSP with Ez = 1 at the center of the gap termination, the Ez component at the center of the gap termination for the reflected in-going RGSP is denoted by $r ^{\prime}_1$ (called the Ez-component reflection coefficient of the RGSP). Then there is $r^{\prime}_1=r_1{H^{(2)}_0} {(k_\textrm {RGSP}R)}/H^{(1)}_0 {(k_\textrm {RGSP}R)}$ with kRGSP = k0neff,G, neff,G being the complex effective index of the RGSP mode. Figures S3(a1)-(a2) in Supplement 1 Sec. S1B show that $r ^{\prime}_1$ is a slowly varying function of the antenna radius R and approaches a constant with the increase of R. This is due to the fact that in the limit case of R→∞, the reflection of the out-going RGSP at the gap termination becomes the reflection of the planar-waveguide SPP at a flat termination. Using the asymptotic expressions of Hankel functions, one obtains $r_1$$r ^{\prime}_1$exp[i2(k0neff,GRπ/4)], substituting which into Eq. (11), one can obtain,
$$\arg ({r^{\prime}_1}) + 2[{k_0}{\textrm{Re}} ({n_{\textrm{eff,G}}})R - \frac{\pi }{4}] = 2Q\pi .$$

Equation (12) can be used to determine the antenna radius R = Rres at resonance. It can be rewritten as R = f(R) with f(R) = [2−arg($r ^{\prime}_1$)+π/2]/[2k0Re(neff,G)]. The transcendental equation R = f(R) can be solved with a linear-interpolation iterative method [73,87] to determine the R = Rres. Besides, since arg($r ^{\prime}_1$) is a slowly varying function of R (as shown in Fig. S3(a2) in Supplement 1 Sec. S1B) and neff,G is independent of R, f(R) is a slowly varying function of R and even approaches a constant with the increase of R, so that f(R) can be approximately treated as an analytical expression of Rres. The R = Rres are shown by the horizontal black-white lines in Fig. 5(b), and can accurately predict the resonance-peak positions of Γtot. This confirms that when R is large (R≥0.081µm), the spontaneous emission enhancement of the antenna is mainly due to the resonance of the RGSP.

In Eq. (12), k0neff,GRπ/4 is the phase shift accumulated by the RGSP propagating from the center to the termination of the nanogap. Therefore, when R = Rres, the phase shift accumulated by the RGSP propagating over one round in the nanogap [the left side of Eq. (12)] is multiples of 2π, which leads to a Fabry-Perot resonance of the RGSP. Compared with the resonance conditions of RGSP such as those in Refs. [36,56,57], Eq. (12) logically derives from the requirement of spontaneous emission enhancement, exhibiting an automatic inclusion of the reflection-induced phase change (the first term) and the propagation-induced phase shift (the second term).

Equation (8) shows that the resonance of the RGSP makes the |as|, |bs|, |ac| and |bc| take the maximum value, and thus enhances the RGSP (with coefficients a and b) in the nanogap. The latter enhances the electric field at the point-source position [see Eq. (9)], and thereby enhances the total emission rate Γtot = −Re(Ez,source)/2. In addition, by comparing Fig. 5(b) with Figs. S11(a) and (b) (see Supplement 1 Sec. S5A), corresponding to gap widths w = 10, 5 and 2nm, respectively, it can be seen that the smaller w is, the smaller the difference between two adjacent Rres [which is λ/[2Re(neff,G)] according to Eq. (12)] will be. This is due to the increase of Re(neff,G) with the decrease of w (see Fig. S4(b) in Supplement 1 Sec. S1B).

Figure 5(b1) shows that with the increase of Rres, Γtot at R = Rres decreases gradually. The reason is that when R = Rres, the denominator of as, bs, ac and bc in Eqs. (5a)–(5b) and (7a)–(7b) becomes,

$$ \left(1-r_{0} r_{1}\right)_{R=R_{\mathrm{res}}}=1-\left|r_{1}^{\prime}\right| \exp \left[-k_{0} \operatorname{Im}\left(n_{\mathrm{eff}, \mathrm{G}}\right) 2 R_{\text {res }}\right] $$

Equation (13) shows that when R = Rres and with the increase of Rres, the denominator of as, bs, ac and bc increases, which results in a decrease of |as|, |bs|, |ac| and |bc|, and consequently a decrease of |a| and |b| [see Eq. (8)], and finally a decrease of Γtot [see Eq. (9)].

Similar to the total emission rate Γtot, for a smaller R (R < 0.081µm), the radiative emission rate Γrad reaches the maximum when L = Lres(R). The reason is that when L = Lres(R), the ASP coefficients |c| and |d| reach the maximum [see Eq. (3)], which then results in the maximum of Γrad (see Eq. (S23) in Supplement 1 Sec. S2A). With the increase of R (R≥0.081µm), Γrad reaches the maximum when R = Rres. The reason is that when R = Rres, the excitation coefficient |β| of up-going ASP reaches the maximum [Eq. (5c)], which results in the maximum of the ASP coefficients |c| and |d| [Eq. (3)], and finally the maximum of Γrad (Eq. (S23) in Supplement 1 Sec. S2A).

Besides, Figs. 5(b1)-(b2), Figs. S11(a1)-(a2) and (b1)-(b2) in Supplement 1 Sec. S5A (with w = 10, 5, 2nm, respectively) show that when R = Rres, Γtot and Γrad still change slightly with L. This is especially obvious when R = Rres at the first resonance [Q = 1 in Eq. (12)], for which Γrad reaches the maximum (see Figs. S10(a2)-(c2) in Supplement 1 Sec. S5A) while counterintuitively, Γtot approximately reaches the minimum when L = Lres(Rres) [Figs. S10(a1)-(c1)]. This minimum of Γtot is due to a destructive interference of the two terms in the right side of Eq. (8) (see detailed explanations in Supplement 1 Sec. S5A).

As shown in Fig. 5(b3) and Figs. S11(a3), (b3) in Supplement 1 Sec. S5A (corresponding to gap widths w = 10, 5 and 2nm, respectively), the SPP model predicts that the radiation efficiency ηradtot reaches the maximum when R = Rres and L = Lres(Rres), and η decreases with the decrease of w. Detailed explanation of this impact of w on η can be found in Supplement 1 Sec. S5B.

For a narrow gap (w < 10nm), the antenna optical response may be affected by quantum effects such as anomalous skin effect (or surface-enhanced Landau damping), nonlocality and electron spill-out [8890], which can be described by nonclassical electromagnetic formalisms such as the one using the Feibelman d-parameters [89]. For instance, for an Ag-nanodisk NPoM antenna with a gap width of 5nm and an ultrasmall protuberance inside the gap, the dominant impact of quantum effects is a 10% shift of resonance frequency and a modest broadening of the resonance peak, while the enhancement factor γtot of total emission rate and the photon-radiation efficiency η′ at resonance are only slightly changed [90]. As for the limit case of gap width approaching 1nm, the resonance-wavelength shift caused by quantum effects is about 30% for an Au-nanodisk NPoM antenna [89]. Referring to these results, we expect that quantum corrections only moderately affect our results for the antenna with a narrow gap (w < 10nm).

3. Analysis of the ring-nanocylinder-on-mirror antenna (R-antenna)

3.1 Rigorous data for the spontaneous emission enhancement by the R-antenna

To show the wide applicability of the SPP model along with its unveiled decisive role of the RGSP and ASP in the spontaneous emission enhancement for other rotationally-symmetric nanoantennas of cylindrical shapes, in this section, we will consider a more complex R-antenna that supports two ASPs. As shown in Fig. 6(a1), the R-antenna is formed by adding a concentric gold nanoring (thickness h = 55nm) out of the gold nanocylinder of the S-antenna. The size of the air nanogap between the nanoring and the nanocylinder is also set to be h. The R-antenna supports two bounded and propagative fundamental ASP modes, which will be considered in the model. The electromagnetic fields of the two ASPs are mainly distributed inside and outside the antenna, respectively, and thus are called inner axially-propagating surface plasmon (i-ASP) and outer axially-propagating surface plasmon (o-ASP), respectively. At wavelength λ=0.7µm, the complex effective indices of the i-ASP and o-ASP are neff,1 = 1.510 + 0.02361i and neff,2 = 1.089 + 0.006250i, respectively, and the distributions of their dominant electric-field component |Eρ| are shown in Figs. 6(b1) and (b2), respectively. The coordinate origin is still set at the position of the point source which is located at the center of the PMMA nanogap.

 figure: Fig. 6.

Fig. 6. (a1) Sketch of the R-antenna excited by a z-polarized point source (red dot) at the center the PMMA nanogap. (a2) Definitions of the unknown SPP coefficients a, b, c1, c2, d1, d2 and p to be solved in the SPP model. (b1)-(b2) Distributions of the dominant electric-field component |Eρ| of the fundamental i-ASP and o-ASP modes, which are obtained with the a-FMM at wavelength λ=0.7µm. The vertical red dashed lines in (b1)-(b2) show the interfaces between different media.

Download Full Size | PDF

Next, we will investigate the spontaneous emission enhancement of the R-antenna based on rigorous numerical calculations. As shown in Figs. 7(a1)-(a3), the enhancement factors of total emission rate (γtottotair) and photon-radiation emission rate ($\gamma^{\prime}_\textrm {rad}$=$\Gamma^{\prime}_\textrm {rad}$air), and the photon-radiation efficiency ($\eta^{\prime}$=$\Gamma^{\prime}_\textrm {rad}$tot) of the R-antenna obtained with the full-wave a-FMM are plotted as functions of the antenna radius R and length L at a fixed excitation wavelength λ=0.7µm for a thickness w = 10nm of the PMMA layer. According to the results, the antenna radius and length are respectively determined as R = 0.092µm and L = 0.314µm, for which η′ takes the maximum value and $\gamma^{\prime}_\textrm {rad}$ is close to the maximum. Then we calculate the frequency spectrum of the γtot and $\gamma^{\prime}_\textrm {rad}$ for the S-antenna and R-antenna with the full-wave a-FMM. The results are shown in Figs. 7(b1)-(b2), where ω/(2πc) = 1/λ with ω being the angular frequency and c being the light speed in vacuum. For the S-antenna, the range of 1/λ at half maximum of γtot is [1.355, 1.468]µm−1. For the R-antenna, this range is [1.274, 1.422]µm−1, which is 1.31 times of that of the S-antenna. Since the fluorescence emission wavelength of molecules or quantum dots usually covers a range of tens to hundreds of nanometers [35,31], it is of great importance to design optical nanoantennas with a broadband enhancement of the spontaneous emission rate [9193]. In addition, like the S-antenna, the R-antenna can also realize a directional radiation of the far field into a central angular zone (see Fig. S9(c) in Supplement 1 Sec. S4), which improves the objective collection efficiency of the fluorescence [4,5,17,18,28].

 figure: Fig. 7.

Fig. 7. (a1)-(a3) Enhancement factors of total emission rate (γtottotair) and photon-radiation emission rate ($\gamma^{\prime}_\textrm {rad}$=$\Gamma^{\prime}_\textrm {rad}$air), and photon-radiation efficiency ($\eta^{\prime}$=$\Gamma^{\prime}_\textrm {rad}$tot) plotted as functions of the radius R and length L of the R-antenna. (b1)-(b2) Frequency spectra of Γtotair and $\Gamma^{\prime}_\textrm {rad}$air for the R-antenna (blue-solid curves) and S-antenna (red-dashed curves), respectively. The structural parameters are R = 0.092µm and L = 0.314µm for the R-antenna, and are R = 0.082µm and L = 0.368µm for the S-antenna. The results in (a)-(b) are calculated with the full-wave a-FMM.

Download Full Size | PDF

3.2 SPP model of the R-antenna

To show the decisive role of the RGSP and ASP in the spontaneous emission enhancement by the R-antenna, in this subsection, we will extend the SPP model to the R-antenna. As shown in Fig. 6(a2), c1 and d1 (c2 and d2) represent the unknown coefficients of the up-going and down-going i-ASPs (o-ASPs), respectively, which satisfy the same normalization as that of the ASPs of the S-antenna. By considering the excitation and multiple scattering processes of the i-ASPs and o-ASPs, a set of coupled-mode equations can be written,

$${c_1} = {\beta _1} + {d_1}{u_1}{\rho _{11}} + {d_2}{u_2}{\rho _{12}},$$
$${c_2} = {\beta _2} + {d_1}{u_1}{\rho _{21}} + {d_2}{u_2}{\rho _{22}},$$
$${d_1} = {c_1}{u_1}{r_{11}} + {c_2}{u_2}{r_{12}},$$
$${d_2} = {c_1}{u_1}{r_{21}} + {c_2}{u_2}{r_{22}},$$
where u1 = exp(ik0neff,1L) and u2 = exp(ik0neff,2L) are phase-shift factors of the i-ASP and o-ASP accumulated over the antenna arm, neff,1 and neff,2 are dimensionless complex effective indices of the i-ASP and o-ASP, respectively. The scattering coefficients βp and ρpq (p, q = 1, 2) are defined in Figs. 8(a1)-(a3), and rpq (p, q = 1, 2) are defined in Figs. 8(b1)-(b2). These scattering coefficients can be calculated as scattering matrix elements [85] with the full-wave a-FMM under the cylindrical coordinate system [7274,86]. Besides, βp and ρpq can be analytically expressed with the elementary scattering coefficients defined in Figs. 4(a1)-(a2) and Fig. S(5) in Supplement 1 Sec. S1C by considering the coupling among the RGSP, i-ASP and o-ASP (the whole process is similar to that of Eqs. (4)-(7), see Supplement 1 Sec. S1C). Similar to the β and ρc in Eq. (2a), βp and ρpq are also called derived scattering coefficients. And Eq. (14) can be understood in a similar way as Eq. (2).

 figure: Fig. 8.

Fig. 8. Definitions of the derived scattering coefficients and electromagnetic fields used in the SPP model of the R-antenna. (a1) For a semi-infinite gold nanoring-cylinder on a gold substrate excited by a z-polarized point source in the nanogap, β1 and β2 denote the excitation coefficients of up-going i-ASP and o-ASP, respectively, as and bs are coefficients of out-going and in-going RGSPs, respectively, ps is the coefficient of out-going RISP, and Ψs denotes the field excited by the point source. (a2) For the same structure as (a1) but illuminated by a down-going i-ASP with unit coefficient, ρ11 and ρ21 are coefficients of the reflected up-going i-ASP and o-ASP, respectively, ac,1 and bc,1 are coefficients of the out-going and in-going RGSPs, respectively, pc,1 is the coefficient of out-going RISP, and Ψc,i denotes the field excited by the down-going i-ASP. (a3) For the same structure as (a1) but illuminated by a down-going o-ASP with unit coefficient, ρ12 and ρ22 are the coefficients of the reflected up-going i-ASP and o-ASP respectively, ac,2 and bc,2 are coefficients of the out-going and in-going RGSPs, respectively, pc,2 is the coefficient of out-going RISP, and Ψc,o denotes the field excited by the down-going o-ASP. (b1) For an up-going i-ASP at the top termination of a semi-infinite gold nanoring-cylinder, r11 and r21 are coefficients of the reflected down-going i-ASP and o-ASP, respectively. (b2) For an up-going o-ASP at the top termination of the same structure as (b1), r12 and r22 are coefficients of the reflected down-going i-ASP and o-ASP, respectively.

Download Full Size | PDF

Similar to Eq. (8), the RGSP coefficients a and b [as shown in Fig. 6(a2)] can be expressed as,

$$a = {a_\textrm{s}} + {d_1}{u_1}{a_{\textrm{c,1}}} + {d_2}{u_2}{a_{\textrm{c,2}}},$$
$$b = {b_\textrm{s}} + {d_1}{u_1}{b_{\textrm{c,1}}} + {d_2}{u_2}{b_{\textrm{c,2}}}.$$

The as, bs, ac,p and bc,p (p = 1, 2) in Eq. (15) are defined in Figs. 8(a1)-(a3). They can be expressed analytically with the elementary scattering coefficients defined in Figs. 4(a1)-(a2) and Fig. S5 in Supplement 1 Sec. S1C [the whole process is similar to Eqs. (4)-(7)]. The expressions of as and bs are the same as Eqs. (5a)–(5b), and the expressions of ac,p and bc,p are provided in Supplement 1 Sec. S1C. Similar to the as, bs, ac and bc in Eqs. (4)-(7), the as, bs, ac,p and bc,p in Eq. (15) are also called derived scattering coefficients.

Next, the SPP coefficients can be obtained by solving Eqs. (14)-(15). Substituting Eqs. (14a)–(14b) into Eqs. (14c)–(14d) to eliminate c1 and c2, one can obtain a set of linear equations about d1 and d2,

$$\left[ {\begin{array}{cc} {1 - (u_1^2{r_{11}}{\rho_{11}} + {u_1}{u_2}{r_{12}}{\rho_{21}})}&{ - ({u_2}{u_1}{r_{11}}{\rho_{12}} + u_2^2{r_{12}}{\rho_{22}})}\\ { - (u_1^2{r_{21}}{\rho_{11}} + {u_1}{u_2}{r_{22}}{\rho_{21}})}&{1 - ({u_1}{u_2}{r_{21}}{\rho_{12}} + u_2^2{r_{22}}{\rho_{22}})} \end{array}} \right]\left[ {\begin{array}{c} {{d_1}}\\ {{d_2}} \end{array}} \right] = \left[ {\begin{array}{c} {{\beta_1}{u_1}{r_{11}} + {\beta_2}{u_2}{r_{12}}}\\ {{\beta_1}{u_1}{r_{21}} + {\beta_2}{u_2}{r_{22}}} \end{array}} \right],$$
i.e., Kd=β, where K is the coefficient matrix, d = [d1, d2]T, and β=[β1u1r11+β2u2r12, β1u1r21+β2u2r22]T. Solving d1 and d2 from Eq. (16) and substituting them into Eqs. (14a)–(14b) and Eq. (15), one can obtain c1, c2 and a, b. The above process can be expressed with matrices as,
$$\left[ {\begin{array}{c} a\\ b\\ {{c_1}}\\ {{c_2}}\\ {{d_1}}\\ {{d_2}} \end{array}} \right] = \left[ {\begin{array}{cc} {{a_{c,1}}{u_1}}&{{a_{\textrm{c,}2}}{u_2}}\\ {{b_{c,1}}{u_1}}&{{b_{\textrm{c,}2}}{u_2}}\\ {{u_1}{\rho_{11}}}&{{u_2}{\rho_{12}}}\\ {{u_1}{\rho_{21}}}&{{u_2}{\rho_{22}}}\\ 1&0\\ 0&1 \end{array}} \right]\left[ {\begin{array}{c} {{d_1}}\\ {{d_2}} \end{array}} \right]\textrm{ + }\left[ {\begin{array}{c} {{a_\textrm{s}}}\\ {{b_\textrm{s}}}\\ {{\beta_1}}\\ {{\beta_2}}\\ 0\\ 0 \end{array}} \right],$$
i.e., b = Fd + s, where b = [a, b, c1, c2, d1, d2]T, F is the coefficient matrix, and s is the excitation term.

The electromagnetic field in the nanogap of the R-antenna can be expressed as Eq. (9), where the RGSP coefficients a and b are given by Eqs. (16)–(17). Finally, the total emission rate Γtot = −Re(Ez,source)/2 can be obtained.

By using the above SPP model, the electromagnetic field Ψrad in the free space outside the antenna can be obtained as well (see Eq. (S26) in Supplement 1 Sec. S2B), from which the radiative emission rate Γrad and the far-field radiation pattern of the antenna (see Fig. S9(c) in Supplement 1 Sec. S4) can be obtained. Moreover, with the SPP model, the coefficient p and the resultant excitation rate ΓRISP of the RISP outside the antenna can be also obtained (see Eq. (S34) in Supplement 1 Sec. S3B).

3.3 Analysis of the R-antenna with the SPP model

In this subsection, we will use the SPP model to analyze the impact of the RGSP and ASP on the spontaneous emission enhancement of the R-antenna. Firstly, we will test the validity of the SPP model against full-wave a-FMM results. Figures 9(a1)-(a3) show the frequency spectra of the enhancement factors of total emission rate (γtottotair), radiative emission rate (γradradair) and the RISP excitation rate (γRISPRISPair), which are obtained for the structural parameters determined in Sec. 3.1 (R = 0.092µm, L = 0.314µm, h = 55nm, w = 10nm). The model predictions (red-solid curves) agree well with the full-wave a-FMM calculations (blue circles), which confirms the validity of the SPP model. This agreement indicates that the RGSP and ASP play a decisive role in the spontaneous emission enhancement of the R-antenna.

 figure: Fig. 9.

Fig. 9. Frequency spectra of the radiation enhancement factors (a) and of the SPP energy flux ΦSPP,i (b), where ω/(2πc) = 1/λ with ω being the angular frequency and c being the light speed in vacuum. (a1)-(a3) show the enhancement factors of total emission rate (γtottotair), radiative emission rate (γradradair) and RISP excitation rate (γRISPRISPair), respectively. The curves with different colors in (b) show ΦSPP,i (normalized by Γair) of different SPP modes with coefficients a, b, c1, c2, d1, d2 (corresponding to i = 1, 2, …, 6, respectively). The circles, solid curves and crosses in (a)-(b) show the results of the full-wave a-FMM, SPP model and QNM model, respectively. The vertical green-dashed lines show the Re($\tilde{\omega }$m) obtained by solving Eq. (20) of the QNM model.

Download Full Size | PDF

Besides, the SPP model can predict the energy flux of different SPP modes in the R-antenna, so as to clarify the contributions of different SPP modes to the antenna near field and its determined far-field radiation properties [34,93,94]. The energy flux of each SPP mode can be expressed as,

$${\Phi _{\textrm{SPP},i}} = {|{{b_i}} |^2}\int\!\!\!\int_\Sigma {{\textbf{S}_{\textrm{SPP},i}} \cdot \textbf{n}} ds,$$
where bi (i = 1, 2, …, 6) is the i-th element of b = [a, b, c1, c2, d1, d2]T in Eq. (17) and represents the coefficient of a SPP mode, SSPP,i represents the Poynting vector of a normalized SPP mode corresponding to coefficient bi, Σ is a cross-section perpendicular to the propagation direction of the mode, and n is the unit vector along the propagation direction of the mode. The frequency spectrum of ΦSPP,i (normalized by Γair) is shown in Fig. 9(b). The predictions of the SPP model (solid curves) are in good agreement with the full-wave a-FMM results (circles). The frequencies where ΦSPP,i takes large values are consistent with those where the total emission rate Γtot takes larger values, which further confirms that the RGSP and ASP play a major role in the spontaneous emission enhancement of the R-antenna.

3.4 QNM model of the R-antenna

Figure 9(a) shows that the R-antenna can achieve a broadband enhancement of the spontaneous emission rate which is formed by several resonance peaks close to each other. These resonance properties, however, are not easy to be explained explicitly with the SPP model. To provide an explicit explanation of these resonance properties, we will further establish a semianalytical model for the resonant modes of the R-antenna based on the SPP model, as stated below.

Theoretically, the antenna can be regarded as an open resonator, so that its supported resonant mode (i.e., localized surface plasmon resonance, LSPR) can be rigorously defined as the QNM [58,6669], which is the eigensolution of the source-free Maxwell’s equations and satisfies the out-going wave condition at infinity. Due to the energy loss of the resonator, the QNM corresponds to a complex eigen/resonant frequency, and its electromagnetic field diverges at infinity [67,68]. After solving the QNMs, the electromagnetic field excited by external source can be expanded upon the basis of QNMs. This QNM expansion can provide analytical expressions for the dependence of the source-excited field on the frequency, polarization and distribution of the excitation source [58,66,67,69,95].

In the following, we will establish a semianalytical model (called QNM model) for the QNMs supported by the R-antenna based on the SPP model. In previous works, the QNMs are commonly obtained via a full-wave numerical solution of source-free Maxwell’s equations [35,66,67,70], which conceals the physical mechanism in forming the QNMs. With the QNM model, analytical expressions for the complex eigenfrequencies and electromagnetic fields of the QNMs can be obtained in terms of the scattering coefficients and electromagnetic fields of SPPs, which will be helpful to clarify the physical mechanism on the formation of the QNMs. With the QNM expansion, the QNM model can further provide analytical expressions for the dependence of the source-excited electromagnetic field and the resultant spontaneous emission rate on the frequency, which can achieve an explicit explanation of the resonance properties of the R-antenna.

First of all, similar to Eq. (9) of the SPP model of the S-antenna, the source-excited electromagnetic field on the R-antenna can be expressed as,

$$\boldsymbol{\Psi }(\textbf{r},\omega ) = [{\boldsymbol{\Psi }_{\textrm{s,0}}}(\textbf{r},\omega ) - {\beta _0}(\omega ){\boldsymbol{\Psi }_{\textrm{RGSP, + }}}(\textbf{r},\omega )] + {\boldsymbol{\Psi }_{\textrm{SPP}}}(\textbf{r},\omega )\textbf{b}(\omega ),$$
where ΨSPP(r,ω) = (ΨRGSP,+, ΨRGSP,−, Ψi‐ASP,+, Ψo‐ASP,+, Ψi‐ASP,−, Ψo‐ASP,−), ΨRGSP,+ and ΨRGSP,− denote the fields of out-going and in-going RGSPs in the nanogap, respectively, Ψi‐ASP,+ and Ψi‐ASP,− (Ψo‐ASP,+, Ψo‐ASP,−) denote the fields of up-going and down-going i‐ASPs (o‐ASPs) on the antenna arm, respectively, and b = [a, b, c1, c2, d1, d2]T are the coefficients of SPPs [defined after Eq. (17)].

For the QNMs, since they are eigensolutions of the source-free Maxwell’s equations, one needs to remove the excitation terms β1, β2 and as, bs in the coupled-SPP Eqs. (14)-(15). Accordingly, Eq. (16) becomes a set of homogeneous linear equations Kd = 0. To ensure the existence of the nontrivial solutions that correspond to the QNMs, the determinant of the coefficient matrix should be zero,

$$\det [\textbf{K}(\omega )] = 0.$$

The complex eigenfrequencies $\tilde{\omega }$m (m = 1, 2, …) of QNMs can be obtained by solving Eq. (20) with ω as the unknown. The transcendental Eq. (20) can be solved with a linear-interpolation iterative method [73,87]. Then the unknown SPP coefficients d = dm can be determined by solving the non-zero solution of K($\tilde{\omega }$m)d = 0. Substituting d = dm into Eq. (17) (setting the excitation term s = 0), one then obtains all the unknown SPP coefficients b = bm = F($\tilde{\omega }$m)dm. Substituting the SPP coefficients bm into Eq. (19) (removing the excitation term Ψs,0β0ΨRGSP,+), one can obtain the electromagnetic field of the m-th QNM,

$${\boldsymbol{\Psi }_{\textrm{QNM},m}}(\textbf{r}) = {\boldsymbol{\Psi }_{\textrm{SPP}}}(\textbf{r},{\tilde{\omega }_m}){\textbf{b}_m} = {\boldsymbol{\Psi }_{\textrm{SPP}}}(\textbf{r},{\tilde{\omega }_m})[\textbf{F}({\tilde{\omega }_m}){\textbf{d}_m}].$$

In the following, we will consider the case with the presence of the excitation source, and expand the source-excited electromagnetic field and SPP coefficients upon the basis of QNMs, so as to provide analytical expressions for their dependence on the excitation frequency. Solving Eq. (16) under the point-source excitation, the SPP coefficients d1 and d2 can be expressed as,

$$\textbf{d}(\omega ) = \frac{{{\textbf{K}^\ast }(\omega )\boldsymbol{\mathrm{\beta}}(\omega )}}{{\det [\textbf{K}(\omega )]}},$$
where K−1 = K*/det(K), with K* representing the adjugate matrix of K. Inserting Eq. (22) into Eq. (17), all the unknown SPP coefficients b(ω) can be determined as,
$$\textbf{b}(\omega ) = \textbf{F}(\omega )\frac{{{\textbf{K}^\ast }(\omega )\boldsymbol{\mathrm{\beta}}(\omega )}}{{\det [\textbf{K}(\omega )]}} + \textbf{s}(\omega ).$$

Inserting Eq. (23) into Eq. (19), one can obtain the source-excited electromagnetic field,

$$\boldsymbol{\Psi }(\textbf{r},\omega ) = [{{\boldsymbol{\Psi }_{\textrm{s,0}}}(\textbf{r},\omega ) - {\beta_0}(\omega ){\boldsymbol{\Psi }_{\textrm{RGSP, + }}}(\textbf{r},\omega )} ]+ {\boldsymbol{\Psi }_{\textrm{SPP}}}(\textbf{r},\omega )\{ \textbf{F}(\omega )\frac{{{\textbf{K}^\ast }(\omega )\boldsymbol{\mathrm{\beta}}(\omega )}}{{\det [\textbf{K}(\omega )]}} + \textbf{s}(\omega )\} .$$

Equations (24) shows that as the excitation frequency ω approaches the complex eigenfrequency $\tilde{\omega }$m of QNMs, the source-excited electromagnetic field will tend to infinity in view of det[K($\tilde{\omega }$m)] = 0 [Eq. (20)]. Therefore, $\tilde{\omega }$m is the complex frequency pole of Ψ(r,ω). We assume that Ψ(r,ω) is a meromorphic function of ω, with $\tilde{\omega }$m being the first-order and nonzero complex pole, and that Ψ(r,ω) is bounded as ω→∞ in the complex plane. Then with the use of the complex-pole expansion theorem (Mittag-Leffler theorem) [58,93,95,96], Ψ(r,ω) can be expressed as,

$$\boldsymbol{\Psi }(\textbf{r},\omega ) \approx \boldsymbol{\Psi }(\textbf{r},0) + \sum\limits_{m = 1}^M {\frac{{\omega /{{\tilde{\omega }}_m}}}{{\omega - {{\tilde{\omega }}_m}}}{\textbf{p}_m}(\textbf{r})} ,$$
where the approximate equality will become equality when M→∞. There is Ψ(r,0)≈0 in view that the scattering effect of the finite-size resonator will vanish for an infinitely large wavelength. ${\bf p}_{m}({\bf r})=\lim_{\omega\rightarrow{\tilde\omega_m}}(\omega-\tilde{\omega }_{m})\boldsymbol{\Psi}({\bf r},\omega)$ is the residue of Ψ(r,ω) at the complex pole $\tilde{\omega }$m and can be calculated by using Eq. (24),
$$\begin{aligned} {\textbf{p}_m}(\textbf{r}) &= \mathop {\lim }\limits_{\omega \to {{\tilde{\omega }}_m}} (\omega - {{\tilde{\omega }}_m}){\boldsymbol{\Psi }_{\textrm{SPP}}}(\textbf{r},\omega )\{ \textbf{F}(\omega )\frac{{{\textbf{K}^\ast }(\omega )\boldsymbol{\mathrm{\beta}}(\omega )}}{{\det [\textbf{K}(\omega )]}}\} \\ &= {\boldsymbol{\Psi }_{\textrm{SPP}}}(\textbf{r},{{\tilde{\omega }}_m})\frac{{\textbf{F}({{\tilde{\omega }}_m}){\textbf{d}_m}\textbf{c}_m^{\textrm {T}} \boldsymbol{\mathrm{\beta}}({{\tilde{\omega }}_m})}}{{{{\{ {{\partial \det [\textbf{K}(\omega )]} / {\partial \omega }}\} }_{\omega = {{\tilde{\omega }}_m}}}}} = \frac{{\textbf{c}_m^{\textrm {T}} \boldsymbol{\mathrm{\beta}}({{\tilde{\omega }}_m})}}{{{{\{ {{\partial \det [\textbf{K}(\omega )]} / {\partial \omega }}\} }_{\omega = {{\tilde{\omega }}_m}}}}}{\boldsymbol{\Psi }_{\textrm{QNM},m}}(\textbf{r}). \end{aligned}$$

The first equality in Eq. (26) is obtained in view that the Ψs,0β0ΨRGSP,+ and ΨSPP(r,ω)s(ω) in Eq. (24) are non-resonant terms, which then yields $\lim_{\omega \rightarrow \tilde{\omega }_{m}}(\omega-\tilde{\omega }_{m})(\boldsymbol{\Psi}_{{\textrm s},0}-\beta_0\boldsymbol{\Psi}_{{\rm RGSP},+})={\textbf 0}$ and ${{\textrm{lim}}_{\omega\rightarrow{\tilde{\omega}_{m} }}}$(ω$\tilde{\omega }$m)ΨSPP(r,ω)s(ω) = 0. For the derivation of the second equality in Eq. (26), there is K*($\tilde{\omega }$m)=dm$\textbf{c}^{\textrm{T}}_{m}$ when ω=$\tilde{\omega }$m, where $\textbf{c}^{\textrm{T}}_{m}$ is a row vector and dm is the nontrivial solution of K($\tilde{\omega }$m)d = 0. Here we assume that the QNMs are nondegenerate, i.e., there is only one linearly independent solution for K($\tilde{\omega }$m)d = 0, which then yields rank[K($\tilde{\omega }$m)] = dim[K($\tilde{\omega }$m)]−1 and resultantly rank[K*($\tilde{\omega }$m)] = 1. The third equality in Eq. (26) makes use of Eq. (21). Substituting Eq. (26) into Eq. (25), one then obtains the source-excited field Ψ(r,ω) expressed as an expansion upon the basis of the QNM fields ΨQNM,m(r),

$$\boldsymbol{\Psi }(\textbf{r},\omega ) \approx \sum\limits_{m = 1}^M {{\alpha _m}(\omega ){\boldsymbol{\Psi }_{\textrm{QNM,}m}}(\textbf{r})} ,$$
$${\alpha _m}(\omega ) = \frac{{\omega /{{\tilde{\omega }}_m}}}{{\omega - {{\tilde{\omega }}_m}}}\frac{{\textbf{c}_m^{\textrm {T}} \boldsymbol{\mathrm{\beta}}({{\tilde{\omega }}_m})}}{{{{\{ {{\partial \det [\textbf{K}(\omega )]} / {\partial \omega }}\} }_{\omega = {{\tilde{\omega }}_m}}}}}.$$

Equation (27b) shows that the QNM expansion coefficients αm(ω) can be expressed in terms of the SPP scattering coefficients at the complex resonant frequency $\tilde{\omega }$m. Note that the QNM expansion of Eq. (27) is fully analytical with respect to the excitation frequency ω.

Similarly, as the excitation frequency ω of the point source approaches $\tilde{\omega }$m, the SPP coefficients b(ω) will tend to infinity [see Eq. (23)] due to det[K(ω)] = 0 [Eq. (20)]. Therefore, similar to Eq. (27), the expression Eq. (23) of the source-excited SPP coefficients b(ω) can be also expanded upon the basis of QNMs based on the Mittag-Leffler theorem [58,93,95,96] as follows,

$$\textbf{b}(\omega ) \approx \sum\limits_{m = 1}^M {{\alpha _m}(\omega )} {\textbf{b}_m},$$
where αm(ω) is given by Eq. (27b), and bm is the SPP coefficients contained in the m-th QNM, see Eq. (21).

Next, we will verify the validity of the QNM model. Since QNM corresponds to a complex eigenfrequency, it is required to obtain the refractive index ng(ω) of gold at complex-valued frequencies ω. Here ng(ω) at complex ω is obtained through an analytic continuation by using an analytical expression of ng(ω), which is obtained with a polynomial fitting of ng(ω) at real ω [71] (with a fitting wavelength range of [0.55, 1]µm and a quintic polynomial). For the R-antenna, we still use the structural parameters determined in Sec. 3.1 (R = 0.092µm, L = 0.314µm, h = 55nm, w = 10nm). Firstly, we compare the complex resonant frequencies $\tilde{\omega }$m of QNMs predicted by the QNM model with those calculated by the full-wave a-FMM. For the former, we can obtain $\tilde{\omega }$m/(2πc) = 1.301 − 0.03152i, 1.409 − 0.03581i, 1.468 − 0.02024iµm−1 by solving Eq. (20). For the latter, the results can be obtained by solving a nonlinear equation f(ω) = 1/F(ω) = 0 with the linear-interpolation iterative method [73,87], where F is a certain component of the complex electromagnetic field at a certain position and can be calculated by the full-wave a-FMM [7274,86]. The results are $\tilde{\omega }$m/(2πc) = 1.300 − 0.02787i, 1.406 − 0.03121i, 1.469 − 0.01932iµm−1, corresponding to quality factors [67] Qm = Re($\tilde{\omega }$m)/[−2Im($\tilde{\omega }$m)] = 23.30, 22.53, 38.02, respectively. It can be seen that the results of the two methods are consistent, which confirms the validity of the QNM model in reproducing the QNMs of the R-antenna.

Then we reproduce the frequency spectrum of the enhancement factor γtottotair of the total spontaneous emission rate with the QNM model [Eq. (27)]. Figure 9(a1) shows that the QNM model predictions (green crosses) are in good agreement with the full-wave a-FMM calculations (blue circles) and SPP model results (red-solid lines), which further verifies the validity of the QNM model in reproducing the source-excited electromagnetic field (and the resultant spontaneous emission rate) with the QNM expansion.

By virtue of the frequency analyticity of the QNM expansion [Eq. (27)] provided by the QNM model, it is possible to achieve an explicit explanation of the broadband enhancement of the spontaneous emission rate by the R-antenna [as shown in Fig. 9(a)]. For the three QNMs calculated above, their Im($\tilde{\omega }$m) are close to 0. Therefore, when the source excitation frequency matches Re($\tilde{\omega }$m), i.e., ω=Re($\tilde{\omega }$m), the expansion coefficient αm(ω) of these QNMs in Eq. (27b) and the resultant source-excited electromagnetic field Ψ(r,ω) in Eq. (27a) will take maximum values, i.e., resonance will occur. Consequently, there will appear three peaks at ω=Re($\tilde{\omega }$m) in the frequency spectra of spontaneous emission rates (as shown by the three vertical green-dashed lines in Fig. 9), which are close to each other and thus form a broadband enhancement of the spontaneous emission rate.

In addition, the QNM model [Eq. (28)] can also predict the frequency spectrum of the energy flux ΦSPP,i (i = 1, 2, …, 6) of each SPP mode excited by the point source. Figure 9(b) shows that the QNM model predictions (crosses) coincide well with the a-FMM rigorous calculations (circles) and the SPP model predictions (solid lines). Compared to the SPP model, the QNM model further provides a frequency analyticity for the ΦSPP,i, indicating that the resonance peaks of ΦSPP,i shown in Fig. 9(b) are at ω=Re($\tilde{\omega }$m) (shown by the vertical green-dashed lines) due to a resonant excitation of the QNMs.

More importantly, with the QNM model [Eq. (21)], the contribution of each SPP mode to the QNM field can be determined quantitatively, which provides theoretical recipes for analyzing the field components in the resonant modes measured in experiments [28,37,5965]. Substituting the bm and ΨSPP(r,$\tilde{\omega }$m) in Eq. (21) into Eq. (18), one can obtain the energy flux ΦSPP,i (i = 1, 2, …, 6, corresponding to SPP modes with coefficients a, b, c1, c2, d1, d2, respectively) of each SPP mode in the QNM field, which is normalized to satisfy ΦSPP,1 = 1. The results are provided in Table 2. It can be seen that for a given QNM, the energy fluxes of different SPP modes are different. For example, for the QNM corresponding to $\tilde{\omega }$3, ΦSPP,3 and ΦSPP,5 are much larger than the energy fluxes of other SPP modes. In addition, for different QNMs, the relative weights of the energy fluxes of SPP modes also become different. At each resonant-peak position ω=Re($\tilde{\omega }$m) shown in Fig. 9(b), since the total field is mainly composed of one dominant QNM, the proportions of different SPPs in the total field is close to those in the QNM field.

Tables Icon

Table 2. Energy flux ΦSPP,i (i = 1, 2, …, 6, corresponding to SPP modes with coefficients a, b, c1, c2, d1, d2, respectively) of each SPP mode in the QNM field predicted by the QNM model. The results are obtained for the R-antenna with R = 0.092µm, L = 0.314µm, w = 10nm and h = 55nm, and are normalized to satisfy ΦSPP,1 = 1.

4. Conclusion

In this paper, intuitive and first-principles-based semianalytical models are proposed for rotationally-symmetric optical nanoantennas of cylindrical shapes, aiming at clarifying the role of the RSP and ASP in forming the resonance and the resultant spontaneous emission enhancement by such antennas. Firstly, for a simple S-antenna, we establish a semianalytical SPP model by considering an intuitive multiple-scattering process of the RGSP in the nanogap and ASP on the vertical arm. All physical quantities (such as SPP scattering coefficients) used in the model are rigorously calculated based on the first principles of Maxwell’s equations, without any fitting or artificial setting of model parameters. This ensures that the model has a solid electromagnetic foundation and can provide quantitative predictions of all the antenna radiation properties (such as enhancement factors of total and radiative emission rates, SPP excitation rates, far-field radiation pattern, etc.). The model shows that for the point source located in the nanogap of the antenna, the total emission rate is determined by the RGSP and the radiative emission rate is determined by the ASP. When the antenna radius is small, the ASP is enhanced at resonance and is further coupled to the RGSP in the nanogap or radiated into the free space, so as to enhance the total or radiative emission rate, respectively. When the antenna radius is large, the ASP resonance is weak, while the RGSP is enhanced at resonance and results in an enhancement of the total spontaneous emission rate; meanwhile, the enhanced RGSP is coupled to the ASP on the antenna arm which further radiates into the free space, so as to enhance the radiative emission rate. Based on the model, two phase-matching conditions are derived for predicting the ASP and RGSP resonances in the above two cases.

To show the wide applicability of the SPP model along with its unveiled decisive role of the RGSP and ASP in the spontaneous emission enhancement for other rotationally-symmetric nanoantennas of cylindrical shapes, we extend the SPP model to a more complex R-antenna that supports two ASPs. Moreover, to provide an explicit explanation of the resonance properties of the R-antenna, a semianalytical model (called QNM model) is established for the QNMs supported by the R-antenna based on the SPP model. The QNM model provides analytical expressions for the complex eigenfrequencies and field distributions of QNMs in terms of SPP modes and their scattering coefficients, which quantitatively clarify the contribution of the RGSP and ASP to the formation of QNMs. Then, by expanding the source-excited electromagnetic field upon the basis of QNMs, we obtain an analytical description of the dependence of the source-excited field and the resultant spontaneous emission rate on the excitation frequency, which indicates that the broadband enhancement of the spontaneous emission rate by the R-antenna results from a resonant excitation of multiple QNMs with different resonance frequencies close to each other.

The proposed models clarify the impact of radially-propagating and axially-propagating SPPs on the spontaneous emission enhancement by the rotationally-symmetric optical nanoantennas of cylindrical shapes, which provides guidance for the design of such antennas. The present models can be extended to other rotationally-symmetric nanoantennas of cylindrical shapes that may support multiple RSPs and ASPs, such as concentric multi-grooves (i.e., bull-eye structure) [11,15,17,18], multilayer-disc structures [65], cylindrical dipole antennas [30], etc. Similar models can be developed for analyzing the reciprocal phenomenon of electromagnetic-field enhancement by such nanoantennas under far-field illumination such as uniform plane wave [14,56,57] or cylindrical vectorial beam [2527].

Funding

National Natural Science Foundation of China (62075104, 61775105).

Disclosures

The authors declare no conflicts of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

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

2. P. Muhlschlegel, H. J. Eisler, O. J. Martin, B. Hecht, and D. W. Pohl, “Resonant optical antennas,” Science 308(5728), 1607–1609 (2005). [CrossRef]  

3. M. Pelton, “Modified spontaneous emission in nanophotonic structures,” Nat. Photonics 9(7), 427–435 (2015). [CrossRef]  

4. C. Belacel, B. Habert, F. Bigourdan, F. Marquier, J. P. Hugonin, S. M. de Vasconcellos, X. Lafosse, L. Coolen, C. Schwob, C. Javaux, B. Dubertret, J. J. Greffet, P. Senellart, and A. Maitre, “Controlling spontaneous emission with plasmonic optical patch antennas,” Nano Lett. 13(4), 1516–1521 (2013). [CrossRef]  

5. A. R. Dhawan, C. Belacel, J. U. Esparza-Villa, M. Nasilowski, Z. Wang, C. Schwob, J. P. Hugonin, L. Coolen, B. Dubertret, P. Senellart, and A. Maitre, “Extreme multiexciton emission from deterministically assembled single-emitter subwavelength plasmonic patch antennas,” Light: Sci. Appl. 9(1), 33 (2020). [CrossRef]  

6. R. Esteban, T. V. Teperik, and J. J. Greffet, “Optical patch antennas for single photon emission using surface plasmon resonances,” Phys. Rev. Lett. 104(2), 026802 (2010). [CrossRef]  

7. G. M. Akselrod, C. Argyropoulos, T. B. Hoang, C. Ciracì, C. Fang, J. Huang, D. R. Smith, and M. H. Mikkelsen, “Probing the mechanisms of large Purcell enhancement in plasmonic nanoantennas,” Nat. Photonics 8(11), 835–840 (2014). [CrossRef]  

8. J. B. Khurgin and G. Sun, “Comparative analysis of spasers, vertical-cavity surface-emitting lasers and surface-plasmon-emitting diodes,” Nat. Photonics 8(6), 468–473 (2014). [CrossRef]  

9. K. L. Tsakmakidis, R. W. Boyd, E. Yablonovitch, and X. Zhang, “Large spontaneous-emission enhancements in metallic nanostructures: towards LEDs faster than lasers,” Opt. Express 24(16), 17916–17927 (2016). [CrossRef]  

10. R.-M. Ma, R. F. Oulton, V. J. Sorger, and X. Zhang, “Plasmon lasers: coherent light source at molecular scales,” Laser Photonics Rev. 7(1), 1–21 (2013). [CrossRef]  

11. D. Komisar, S. Kumar, Y. H. Kan, C. Wu, and S. I. Bozhevolnyi, “Generation of radially polarized single photons with plasmonic bullseye antennas,” ACS Photonics 8(8), 2190–2196 (2021). [CrossRef]  

12. L. Li, L. Wang, C. Du, Z. Guan, Y. Xiang, W. Wu, M. Ren, X. Zhang, A. Tang, W. Cai, and J. Xu, “Ultrastrong coupling of CdZnS/ZnS quantum dots to bonding breathing plasmons of aluminum metal-insulator-metal nanocavities in near-ultraviolet spectrum,” Nanoscale 12(5), 3112–3120 (2020). [CrossRef]  

13. K. H. Su, S. Durant, J. M. Steele, Y. Xiong, C. Sun, and X. Zhang, “Raman enhancement factor of a single tunable nanoplasmonic resonator,” J. Phys. Chem. B 110(9), 3964–3968 (2006). [CrossRef]  

14. H. J. Lezec, A. Degiron, E. Devaux, R. A. Linke, L. Martin-Moreno, F. J. Garcia-Vidal, and T. W. Ebbesen, “Beaming light from a subwavelength aperture,” Science 297(5582), 820–822 (2002). [CrossRef]  

15. F. Werschler, B. Lindner, C. Hinz, F. Conradt, P. Gumbsheimer, Y. Behovits, C. Negele, T. de Roo, O. Tzang, S. Mecking, A. Leitenstorfer, and D. V. Seletskiy, “Efficient emission enhancement of single CdSe/CdS/PMMA quantum dots through controlled near-field coupling to plasmonic bullseye resonators,” Nano Lett. 18(9), 5396–5400 (2018). [CrossRef]  

16. F. Hao, P. Nordlander, M. T. Burnett, and S. A. Maier, “Enhanced tunability and linewidth sharpening of plasmon resonances in hybridized metallic ring/disk nanocavities,” Phys. Rev. B 76(24), 245417 (2007). [CrossRef]  

17. H. Aouani, O. Mahboub, N. Bonod, E. Devaux, E. Popov, H. Rigneault, T. W. Ebbesen, and J. Wenger, “Bright unidirectional fluorescence emission of molecules in a nanoaperture with plasmonic corrugations,” Nano Lett. 11(2), 637–644 (2011). [CrossRef]  

18. H. Aouani, O. Mahboub, E. Devaux, H. Rigneault, T. W. Ebbesen, and J. Wenger, “Plasmonic antennas for directional sorting of fluorescence emission,” Nano Lett. 11(6), 2400–2406 (2011). [CrossRef]  

19. N. F. Chiu, C. H. Hou, C. J. Cheng, and F. Y. Tsai, “Plasmonic circular nanostructure for enhanced light absorption in organic solar cells,” Int. J. Photoenergy 2013(1), 1–7 (2013). [CrossRef]  

20. P. Hanarp, M. Käll, and D. S. Sutherland, “Optical properties of short range ordered arrays of nanometer gold disks prepared by colloidal lithography,” J. Phys. Chem. B 107(24), 5768–5772 (2003). [CrossRef]  

21. E. M. Larsson, J. Alegret, M. Kall, and D. S. Sutherland, “Sensing characteristics of NIR localized surface plasmon resonances in gold nanorings for application as ultrasensitive biosensors,” Nano Lett. 7(5), 1256–1263 (2007). [CrossRef]  

22. R. D. Bhat, N. C. Panoiu, S. R. Brueck, and R. M. Osgood, “Enhancing the signal-to-noise ratio of an infrared photodetector with a circular metal grating,” Opt. Express 16(7), 4588–4596 (2008). [CrossRef]  

23. F. F. Ren, W. Z. Xu, J. Ye, K. W. Ang, H. Lu, R. Zhang, M. Yu, G. Q. Lo, H. H. Tan, and C. Jagadish, “Second-order surface-plasmon assisted responsivity enhancement in germanium nano-photodetectors with bull's eye antennas,” Opt. Express 22(13), 15949–15956 (2014). [CrossRef]  

24. K. Ikeda, “Spectroscopy and photoelectrochemistry of organic monolayers within sphere-plane gold nano-gaps,” Electrochemistry 79(10), 768–772 (2011). [CrossRef]  

25. W. Chen, D. C. Abeysinghe, R. L. Nelson, and Q. Zhan, “Plasmonic lens made of multiple concentric metallic rings under radially polarized illumination,” Nano Lett. 9(12), 4320–4325 (2009). [CrossRef]  

26. S. A. Syubaev, A. Y. Zhizhchenko, D. V. Pavlov, S. O. Gurbatov, E. V. Pustovalov, A. P. Porfirev, S. N. Khonina, S. A. Kulinich, J. B. B. Rayappan, S. I. Kudryashov, and A. A. Kuchmizhak, “Plasmonic nanolenses produced by cylindrical vector beam printing for sensing applications,” Sci. Rep. 9(1), 19750 (2019). [CrossRef]  

27. G. M. Lerman, A. Yanai, and U. Levy, “Demonstration of nanofocusing by the use of plasmonic lens illuminated with radially polarized light,” Nano Lett. 9(5), 2139–2143 (2009). [CrossRef]  

28. A. Mohtashami, T. Coenen, A. Antoncecchi, A. Polman, and A. F. Koenderink, “Nanoscale excitation mapping of plasmonic patch antennas,” ACS Photonics 1(11), 1134–1143 (2014). [CrossRef]  

29. Y. Yang, B. Zhen, C. W. Hsu, O. D. Miller, J. D. Joannopoulos, and M. Soljacic, “Optically thin metallic films for high-radiative-efficiency plasmonics,” Nano Lett. 16(7), 4110–4117 (2016). [CrossRef]  

30. J. Paparone, J. Laverdant, G. Brucoli, C. Symonds, A. Crut, N. Del Fatti, J. M. Benoit, and J. Bellessa, “Vertical pillar nanoantenna for emission enhancement and redirection,” J. Phys. D: Appl. Phys. 51(4), 045301 (2018). [CrossRef]  

31. A. Kinkhabwala, Z. Yu, S. Fan, Y. Avlasevich, K. Müllen, and W. E. Moerner, “Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna,” Nat. Photonics 3(11), 654–657 (2009). [CrossRef]  

32. X. Zhou, C. Deeb, S. Kostcheev, G. P. Wiederrecht, P.-M. Adam, J. Béal, J. Plain, D. J. Gosztola, J. Grand, N. Félidj, H. Wang, A. Vial, and R. Bachelot, “Selective functionalization of the nanogap of a plasmonic dimer,” ACS Photonics 2(1), 121–129 (2015). [CrossRef]  

33. H. Wang, Y. Lin, P. Ma, Y. Zhong, and H. Liu, “Tunable fluorescence emission of molecules with controllable positions within the metallic nanogap between gold nanorods and a gold film,” J. Mater. Chem. C 7(43), 13526–13535 (2019). [CrossRef]  

34. A. Pors and S. I. Bozhevolnyi, “Quantum emitters near layered plasmonic nanostructures: decay rate contributions,” ACS Photonics 2(2), 228–236 (2015). [CrossRef]  

35. C. Zhang, J. P. Hugonin, J. J. Greffet, and C. Sauvan, “Surface plasmon polaritons emission with nanopatch antennas: enhancement by means of mode hybridization,” ACS Photonics 6(11), 2788–2796 (2019). [CrossRef]  

36. C. Tserkezis, R. Esteban, D. O. Sigle, J. Mertens, L. O. Herrmann, J. J. Baumberg, and J. Aizpurua, “Hybridization of plasmonic antenna and cavity modes: extreme optics of nanoparticle-on-mirror nanogaps,” Phys. Rev. A 92(5), 053811 (2015). [CrossRef]  

37. D. Denkova, N. Verellen, A. V. Silhanek, P. Van Dorpe, and V. V. Moshchalkov, “Lateral magnetic near-field imaging of plasmonic nanoantennas with increasing complexity,” Small 10(10), 1959–1966 (2014). [CrossRef]  

38. J. Aizpurua, L. Blanco, P. Hanarp, D. S. Sutherland, M. Käll, G. W. Bryant, and F. J. García de Abajo, “Light scattering in gold nanorings,” J. Quant. Spectrosc. Radiat. Transfer 89(1-4), 11–16 (2004). [CrossRef]  

39. L. A. Blanco and F. J. García de Abajo, “Spontaneous light emission in complex nanostructures,” Phys. Rev. B 69(20), 205414 (2004). [CrossRef]  

40. J. Aizpurua, P. Hanarp, D. S. Sutherland, M. Kall, G. W. Bryant, and F. J. Garcia de Abajo, “Optical properties of gold nanorings,” Phys. Rev. Lett. 90(5), 057401 (2003). [CrossRef]  

41. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (John Wiley & Sons, Inc., 1983).

42. E. A. Muljarov, W. Langbein, and R. Zimmermann, “Brillouin-Wigner perturbation theory in open electromagnetic systems,” EPL 92(5), 50010 (2010). [CrossRef]  

43. M. B. Doost, W. Langbein, and E. A. Muljarov, “Resonant-state expansion applied to three-dimensional open optical systems,” Phys. Rev. A 90(1), 013834 (2014). [CrossRef]  

44. Z. Sztranyovszky, W. Langbein, and E. A. Muljarov, “Optical resonances in graded index spheres: A resonant-state expansion study and analytic approximations,” arXiv:2109.10713 (2021).

45. J. M. Nápoles-Duarte, M. A. Chavez-Rojo, M. E. Fuentes-Montero, L. M. Rodríguez-Valdez, R. García-Llamas, and J. A. Gaspar-Armenta, “Surface plasmon resonances in Drude metal cylinders: radius dependence and quality factor,” J. Opt. 17(6), 065003 (2015). [CrossRef]  

46. F. Minkowski, F. Wang, A. Chakrabarty, and Q.-H. Wei, “Resonant cavity modes of circular plasmonic patch nanoantennas,” Appl. Phys. Lett. 104(2), 021111 (2014). [CrossRef]  

47. M. B. Doost, W. Langbein, and E. A. Muljarov, “Resonant state expansion applied to two-dimensional open optical systems,” Phys. Rev. A 87(4), 043827 (2013). [CrossRef]  

48. M. Ruiz and O. Schnitzer, “Slender-body theory for plasmonic resonance,” Proc. R. Soc. A 475(2229), 20190294 (2019). [CrossRef]  

49. M. Ruiz and O. Schnitzer, “Plasmonic resonances of slender nanometallic rings,” arXiv:2107.01716 (2021).

50. A. Mary, D. M. Koller, A. Hohenau, J. R. Krenn, A. Bouhelier, and A. Dereux, “Optical absorption of torus-shaped metal nanoparticles in the visible range,” Phys. Rev. B 76(24), 245422 (2007). [CrossRef]  

51. A. Mary, A. Dereux, and T. L. Ferrell, “Localized surface plasmons on a torus in the nonretarded approximation,” Phys. Rev. B 72(15), 155426 (2005). [CrossRef]  

52. E. Prodan, C. Radloff, N. J. Halas, and P. Nordlander, “A hybridization model for the plasmon response of complex nanostructures,” Science 302(5644), 419–422 (2003). [CrossRef]  

53. E. Prodan and P. Nordlander, “Plasmon hybridization in spherical nanoparticles,” J. Chem. Phys. 120(11), 5444–5454 (2004). [CrossRef]  

54. C. M. Dutta, T. A. Ali, D. W. Brandl, T. H. Park, and P. Nordlander, “Plasmonic properties of a metallic torus,” J. Chem. Phys. 129(8), 084706 (2008). [CrossRef]  

55. P. Nordlander and E. Prodan, “Plasmon hybridization in nanoparticles near metallic surfaces,” Nano Lett. 4(11), 2209–2213 (2004). [CrossRef]  

56. R. Filter, J. Qi, C. Rockstuhl, and F. Lederer, “Circular optical nanoantennas: an analytical theory,” Phys. Rev. B 85(12), 125429 (2012). [CrossRef]  

57. R. Esteban, G. Aguirregabiria, A. G. Borisov, Y. M. M. Wang, P. Nordlander, G. W. Bryant, and J. Aizpurua, “The morphology of narrow gaps modifies the plasmonic response,” ACS Photonics 2(2), 295–305 (2015). [CrossRef]  

58. H. Jia, F. Yang, Y. Zhong, and H. Liu, “Understanding localized surface plasmon resonance with propagative surface plasmon polaritons in optical nanogap antennas,” Photonics Res. 4(6), 293 (2016). [CrossRef]  

59. J. Qi, T. Kaiser, A. E. Klein, M. Steinert, T. Pertsch, F. Lederer, and C. Rockstuhl, “Enhancing resonances of optical nanoantennas by circular gratings,” Opt. Express 23(11), 14583–14595 (2015). [CrossRef]  

60. F. P. Schmidt, H. Ditlbacher, U. Hohenester, A. Hohenau, F. Hofer, and J. R. Krenn, “Dark plasmonic breathing modes in silver nanodisks,” Nano Lett. 12(11), 5780–5783 (2012). [CrossRef]  

61. F. P. Schmidt, A. Losquin, F. Hofer, A. Hohenau, J. R. Krenn, and M. Kociak, “How dark are radial breathing modes in plasmonic nanodisks,” ACS Photonics 5(3), 861–866 (2018). [CrossRef]  

62. S. Kadkhodazadeh, F. A. A. Nugroho, C. Langhammer, M. Beleggia, and J. B. Wagner, “Optical property–composition correlation in noble metal alloy nanoparticles studied with EELS,” ACS Photonics 6(3), 779–786 (2019). [CrossRef]  

63. F. P. Schmidt, H. Ditlbacher, U. Hohenester, A. Hohenau, F. Hofer, and J. R. Krenn, “Universal dispersion of surface plasmons in flat nanostructures,” Nat. Commun. 5(1), 3604 (2014). [CrossRef]  

64. M. Jeannin, N. Rochat, K. Kheng, and G. Nogues, “Cathodoluminescence spectroscopy of plasmonic patch antennas: towards lower order and higher energies,” Opt. Express 25(5), 5488–5500 (2017). [CrossRef]  

65. M. Kuttge, F. J. Garcia de Abajo, and A. Polman, “Ultrasmall mode volume plasmonic nanodisk resonators,” Nano Lett. 10(5), 1537–1541 (2010). [CrossRef]  

66. W. Yan, R. Faggiani, and P. Lalanne, “Rigorous modal analysis of plasmonic nanoresonators,” Phys. Rev. B 97(20), 205422 (2018). [CrossRef]  

67. P. Lalanne, W. Yan, K. Vynck, C. Sauvan, and J.-P. Hugonin, “Light interaction with photonic and plasmonic resonances,” Laser Photonics Rev. 12(5), 1700113 (2018). [CrossRef]  

68. E. S. C. Ching, P. T. Leung, A. Maassen van den Brink, W. M. Suen, S. S. Tong, and K. Young, “Quasinormal-mode expansion for waves in open systems,” Rev. Mod. Phys. 70(4), 1545–1554 (1998). [CrossRef]  

69. C. Sauvan, J. P. Hugonin, I. S. Maksymov, and P. Lalanne, “Theory of the spontaneous optical emission of nanosize photonic and plasmon resonators,” Phys. Rev. Lett. 110(23), 237401 (2013). [CrossRef]  

70. P. Lalanne, W. Yan, A. Gras, C. Sauvan, J. P. Hugonin, M. Besbes, G. Demesy, M. D. Truong, B. Gralak, F. Zolla, A. Nicolet, F. Binkowski, L. Zschiedrich, S. Burger, J. Zimmerling, R. Remis, P. Urbach, H. T. Liu, and T. Weiss, “Quasinormal mode solvers for resonators with dispersive materials,” J. Opt. Soc. Am. A 36(4), 686–704 (2019). [CrossRef]  

71. E. D. Palik, Handbook of Optical Constants of Solids-II (Academic, 1985).

72. J. P. Hugonin and P. Lalanne, “Perfectly matched layers as nonlinear coordinate transforms: a generalized formalization,” J. Opt. Soc. Am. A 22(9), 1844–1849 (2005). [CrossRef]  

73. Y. Li, H. Liu, H. Jia, F. Bo, G. Zhang, and J. Xu, “Fully vectorial modeling of cylindrical microresonators with aperiodic Fourier modal method,” J. Opt. Soc. Am. A 31(11), 2459–2466 (2014). [CrossRef]  

74. H. Liu, DIF CODE for Modeling Light Diffraction in Nanostructures (Nankai University, Tianjin, 2010).

75. L. Novotny and B. Hecht, Principles of Nano-optics (Cambridge University, 2006).

76. H. Jia, H. Liu, and Y. Zhong, “Role of surface plasmon polaritons and other waves in the radiation of resonant optical dipole antennas,” Sci. Rep. 5(1), 8456 (2015). [CrossRef]  

77. S. H. Chang, S. Gray, and G. Schatz, “Surface plasmon generation and light transmission by isolated nanoholes and arrays of nanoholes in thin metal films,” Opt. Express 13(8), 3150–3165 (2005). [CrossRef]  

78. S. Nerkararyan, K. Nerkararyan, N. Janunts, and T. Pertsch, “Generation of Hankel-type surface plasmon polaritons in the vicinity of a metallic nanohole,” Phys. Rev. B 82(24), 245405 (2010). [CrossRef]  

79. J. Yang, J.-P. Hugonin, and P. Lalanne, “Near-to-far field transformations for radiative and guided waves,” ACS Photonics 3(3), 395–402 (2016). [CrossRef]  

80. F. Bigourdan, J.-P. Hugonin, F. Marquier, C. Sauvan, and J.-J. Greffet, “Nanoantenna for electrical generation of surface plasmon polaritons,” Phys. Rev. Lett. 116(10), 106803 (2016). [CrossRef]  

81. C. Vassallo, Optical Waveguide Concepts (Elsevier, 1991).

82. E. M. Purcell, “Spontaneous emission probabilities at radio frequencies,” Phys. Rev. 69(11-12), 681 (1946).

83. S. I. Bozhevolnyi, V. S. Volkov, E. Devaux, J. Y. Laluet, and T. W. Ebbesen, “Channelling surface plasmons,” Appl. Phys. A: Mater. Sci. Process. 89(2), 225–231 (2007). [CrossRef]  

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

85. L. Li, “Formulation and comparison of two recursive matrix algorithms for modeling layered diffraction gratings,” J. Opt. Soc. Am. A 13(5), 1024–1035 (1996). [CrossRef]  

86. L. Li, “Fourier modal method for crossed anisotropic gratings with arbitrary permittivity and permeability tensors,” J. Opt. A: Pure Appl. Opt. 5(4), 345–355 (2003). [CrossRef]  

87. J. M. Ortega and W. C. Rheinboldt, Iterative Solution of Nonlinear Equations in Several Variables (Academic, 1970).

88. S. I. Bozhevolnyi, L. Martin-Moreno, and F. Garcia-Vidal, eds., Quantum Plasmonics (Springer, 2017), Chap. 13.

89. Y. Yang, D. Zhu, W. Yan, A. Agarwal, M. Zheng, J. D. Joannopoulos, P. Lalanne, T. Christensen, K. K. Berggren, and M. Soljačić, “A general theoretical and experimental framework for nanoscale electromagnetism,” Nature 576(7786), 248–252 (2019). [CrossRef]  

90. T. Wu, W. Yan, and P. Lalanne, “Bright plasmons with cubic nanometer mode volumes through mode hybridization,” ACS Photonics 8(1), 307–314 (2021). [CrossRef]  

91. E. J. R. Vesseur, F. J. G. de Abajo, and A. Polman, “Broadband Purcell enhancement in plasmonic ring cavities,” Phys. Rev. B 82(16), 165419 (2010). [CrossRef]  

92. A. P. Edwards and A. M. Adawi, “Plasmonic nanogaps for broadband and large spontaneous emission rate enhancement,” J. Appl. Phys. 115(5), 053101 (2014). [CrossRef]  

93. X. Zhai, N. Wang, Y. Zhong, and H. Liu, “Broadband enhancement of the spontaneous emission by a split-ring nanoantenna: impact of azimuthally propagating surface plasmon polaritons,” IEEE J. Sel. Top. Quantum Electron. 27(1), 1–15 (2021). [CrossRef]  

94. T. H. Taminiau, F. D. Stefani, and N. F. van Hulst, “Optical nanorod antennas modeled as cavities for dipolar emitters: evolution of sub- and super-radiant modes,” Nano Lett. 11(3), 1020–1024 (2011). [CrossRef]  

95. E. A. Muljarov and W. Langbein, “Exact mode volume and Purcell factor of open optical systems,” Phys. Rev. B 94(23), 235438 (2016). [CrossRef]  

96. G. B. Arfken, H. J. Weber, and F. E. Harris, Mathematical Methods for Physicists, 7th ed. (Academic, 2012).

Supplementary Material (1)

NameDescription
Supplement 1       The Supplemental Document provides more details about the theoretical derivations and numerical results.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (9)

Fig. 1.
Fig. 1. (a1) Sketch of the S-antenna excited by a z-polarized point source (red dot) at the center the PMMA nanogap. (a2) Definitions of the unknown SPP coefficients a, b, c, d, p to be solved in the SPP model. (b1) Distribution of the main electric-field component |Ez| of the fundamental RGSP mode. (b2) Distribution of the main electric-field component |Eρ| of the fundamental ASP mode. (b3) Distribution of the main electric-field component |Ez| of the fundamental RISP mode, with the insert showing an enlarged view in the PMMA layer. The vertical red dashed lines in (b1)-(b3) show the interfaces between different media. The results in (b) are obtained with the a-FMM at wavelength λ=0.7µm for a PMMA-layer thickness w = 10 nm and an antenna radius R = 82 nm.
Fig. 2.
Fig. 2. (a)-(c) Enhancement factors of total emission rate (γtot) and photon-radiation emission rate ($\gamma^{\prime}_\textrm {rad}$), and photon-radiation efficiency η′ plotted as functions of antenna radius R and length L. The results are obtained with the full-wave a-FMM at a fixed wavelength λ=0.7µm for a PMMA-layer thickness w = 10 nm.
Fig. 3.
Fig. 3. Definition of the derived scattering coefficients and electromagnetic fields used in the SPP model for the S-antenna. (a) For a semi-infinite gold nanocylinder on a gold substrate excited by a z-polarized point source in the nanogap, β denotes the excitation coefficient of up-going ASP, as and bs are the coefficients of out-going and in-going RGSPs, respectively, ps is the coefficient of out-going RISP, and Ψs denotes the field excited by the point source. (b) For the structure of (a) but illuminated by an incident down-going ASP with unit coefficient, ρc is the reflection coefficient of the ASP at the substrate, ac and bc are the coefficients of out-going and in-going RGSPs, respectively, pc is the coefficient of out-going RISP, and Ψc denotes the field excited by the down-going ASP. (c) Reflection coefficient r of the ASP at the top termination of the semi-infinite gold nanocylinder.
Fig. 4.
Fig. 4. Definition of the elementary scattering coefficients and electromagnetic field used in the SPP model for the S-antenna. (a1) For the point source in the MIM nanogap, β0 denotes the excitation coefficient of the out-going RGSP, and Ψs,0 represents the excited electromagnetic field. (a2) r0 denotes the reflection coefficient of the in-going RGSP at the center of the MIM nanogap. (b1) For a semi-infinite gold nanocylinder on a gold substrate illuminated by an incident out-going RGSP with unit coefficient, r1 is the coefficient of the reflected in-going RGSP at the termination of the gap, α and α0 are the coefficients of the excited up-going ASP and out-going RISP, respectively. (b2) For the structure in (b1) but under illumination by a down-going ASP with unit coefficient, α′ and χ denote the excitation coefficients of the in-going RGSP and the out-going RISP, respectively, and ρ0 is the elementary reflection coefficient of the ASP. Note that ρ0 is different from the ρc defined in Fig. 3(b): ρc contains the contribution of the ASP excited by the out-going RGSP (with coefficient ac), but ρ0 does not.
Fig. 5.
Fig. 5. (a1)-(a3) Spontaneous-emission enhancement factors Γtotair, Γradair and ΓRISPair of the S-antenna plotted as functions of antenna length L. The results are obtained for R = 0.082µm, w = 10 nm and λ=0.7µm. The blue circles and red-solid curves show the full-wave a-FMM results and SPP model predictions, respectively. (b1)-(b3) Γtotair, Γradair and antenna radiation efficiency ηradtot predicted by the SPP model plotted as functions of the antenna radius R and length L. The vertical black-white lines show the L = Lres(R) at resonance determined by Eq. (10), corresponding to P = 0, 1, …, 5 from left to right. The horizontal black-white lines show the R = Rres at resonance determined by Eq. (12), corresponding to Q = 1, 2, …, 6 from bottom to top. (c1)-(c3) SPP scattering coefficients |β|, |ρc| and |r| plotted as functions of antenna radius R. The vertical red-dashed lines in (c1)-(c2) show the R = Rres. The results in (c) are obtained with the full-wave a-FMM at wavelength λ=0.7µm.
Fig. 6.
Fig. 6. (a1) Sketch of the R-antenna excited by a z-polarized point source (red dot) at the center the PMMA nanogap. (a2) Definitions of the unknown SPP coefficients a, b, c1, c2, d1, d2 and p to be solved in the SPP model. (b1)-(b2) Distributions of the dominant electric-field component |Eρ| of the fundamental i-ASP and o-ASP modes, which are obtained with the a-FMM at wavelength λ=0.7µm. The vertical red dashed lines in (b1)-(b2) show the interfaces between different media.
Fig. 7.
Fig. 7. (a1)-(a3) Enhancement factors of total emission rate (γtottotair) and photon-radiation emission rate ($\gamma^{\prime}_\textrm {rad}$=$\Gamma^{\prime}_\textrm {rad}$air), and photon-radiation efficiency ($\eta^{\prime}$=$\Gamma^{\prime}_\textrm {rad}$tot) plotted as functions of the radius R and length L of the R-antenna. (b1)-(b2) Frequency spectra of Γtotair and $\Gamma^{\prime}_\textrm {rad}$air for the R-antenna (blue-solid curves) and S-antenna (red-dashed curves), respectively. The structural parameters are R = 0.092µm and L = 0.314µm for the R-antenna, and are R = 0.082µm and L = 0.368µm for the S-antenna. The results in (a)-(b) are calculated with the full-wave a-FMM.
Fig. 8.
Fig. 8. Definitions of the derived scattering coefficients and electromagnetic fields used in the SPP model of the R-antenna. (a1) For a semi-infinite gold nanoring-cylinder on a gold substrate excited by a z-polarized point source in the nanogap, β1 and β2 denote the excitation coefficients of up-going i-ASP and o-ASP, respectively, as and bs are coefficients of out-going and in-going RGSPs, respectively, ps is the coefficient of out-going RISP, and Ψs denotes the field excited by the point source. (a2) For the same structure as (a1) but illuminated by a down-going i-ASP with unit coefficient, ρ11 and ρ21 are coefficients of the reflected up-going i-ASP and o-ASP, respectively, ac,1 and bc,1 are coefficients of the out-going and in-going RGSPs, respectively, pc,1 is the coefficient of out-going RISP, and Ψc,i denotes the field excited by the down-going i-ASP. (a3) For the same structure as (a1) but illuminated by a down-going o-ASP with unit coefficient, ρ12 and ρ22 are the coefficients of the reflected up-going i-ASP and o-ASP respectively, ac,2 and bc,2 are coefficients of the out-going and in-going RGSPs, respectively, pc,2 is the coefficient of out-going RISP, and Ψc,o denotes the field excited by the down-going o-ASP. (b1) For an up-going i-ASP at the top termination of a semi-infinite gold nanoring-cylinder, r11 and r21 are coefficients of the reflected down-going i-ASP and o-ASP, respectively. (b2) For an up-going o-ASP at the top termination of the same structure as (b1), r12 and r22 are coefficients of the reflected down-going i-ASP and o-ASP, respectively.
Fig. 9.
Fig. 9. Frequency spectra of the radiation enhancement factors (a) and of the SPP energy flux ΦSPP,i (b), where ω/(2πc) = 1/λ with ω being the angular frequency and c being the light speed in vacuum. (a1)-(a3) show the enhancement factors of total emission rate (γtottotair), radiative emission rate (γradradair) and RISP excitation rate (γRISPRISPair), respectively. The curves with different colors in (b) show ΦSPP,i (normalized by Γair) of different SPP modes with coefficients a, b, c1, c2, d1, d2 (corresponding to i = 1, 2, …, 6, respectively). The circles, solid curves and crosses in (a)-(b) show the results of the full-wave a-FMM, SPP model and QNM model, respectively. The vertical green-dashed lines show the Re($\tilde{\omega }$m) obtained by solving Eq. (20) of the QNM model.

Tables (2)

Tables Icon

Table 1. Impact of the refractive index nd of the nanogap on the radiation properties of the S-antenna.

Tables Icon

Table 2. Energy flux ΦSPP,i (i = 1, 2, …, 6, corresponding to SPP modes with coefficients a, b, c1, c2, d1, d2, respectively) of each SPP mode in the QNM field predicted by the QNM model. The results are obtained for the R-antenna with R = 0.092µm, L = 0.314µm, w = 10nm and h = 55nm, and are normalized to satisfy ΦSPP,1 = 1.

Equations (45)

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

E = ( ρ , φ , z ) [ i 2 [ H n + 1 ( 1 ) ( k ρ ) H n 1 ( 1 ) ( k ρ ) ] E x , 0 ( z ) 1 2 [ H n + 1 ( 1 ) ( k ρ ) + H n 1 ( 1 ) ( k ρ ) ] E x , 0 ( z ) H n ( 1 ) ( k ρ ) E z , 0 ( z ) ] exp ( i n φ )
H = ( ρ , φ , z ) [ 1 2 [ H n + 1 ( 1 ) ( k ρ ) + H n 1 ( 1 ) ( k ρ ) ] H y , 0 ( z ) i 2 [ H n + 1 ( 1 ) ( k ρ ) H n 1 ( 1 ) ( k ρ ) ] H y , 0 ( z ) 0 ] exp ( i n φ )
c = β + d u ρ c ,
d = c u r ,
c = β 1 u 2 ρ c r ,
d = β u r 1 u 2 ρ c r .
a s = β 0 + b s r 0 ,
b s = a s r 1 ,
β = a s α ,
a s  =  β 0 1 r 0 r 1 ,
b s = β 0 r 1 1 r 0 r 1 ,
β = β 0 α 1 r 0 r 1 .
a c = b c r 0 ,
b c = α + a c r 1 ,
ρ c = ρ 0 + a c α ,
a c = α r 0 1 r 0 r 1 ,
b c = α 1 r 0 r 1 ,
ρ c = ρ 0 + α r 0 α 1 r 0 r 1 .
a  =  a s + d u a c ,
b = b s + d u b c .
Ψ gap = ( Ψ s,0 β 0 Ψ RGSP, +  ) + a Ψ RGSP, +  + b Ψ RGSP, ,
2 k 0 Re ( n eff,A ) L  +  arg ( ρ c )  +  arg ( r ) = 2 P π ,
arg ( r 0 ) + arg ( r 1 ) = 2 Q π ,
arg ( r 1 ) + 2 [ k 0 Re ( n eff,G ) R π 4 ] = 2 Q π .
( 1 r 0 r 1 ) R = R r e s = 1 | r 1 | exp [ k 0 Im ( n e f f , G ) 2 R res  ]
c 1 = β 1 + d 1 u 1 ρ 11 + d 2 u 2 ρ 12 ,
c 2 = β 2 + d 1 u 1 ρ 21 + d 2 u 2 ρ 22 ,
d 1 = c 1 u 1 r 11 + c 2 u 2 r 12 ,
d 2 = c 1 u 1 r 21 + c 2 u 2 r 22 ,
a = a s + d 1 u 1 a c,1 + d 2 u 2 a c,2 ,
b = b s + d 1 u 1 b c,1 + d 2 u 2 b c,2 .
[ 1 ( u 1 2 r 11 ρ 11 + u 1 u 2 r 12 ρ 21 ) ( u 2 u 1 r 11 ρ 12 + u 2 2 r 12 ρ 22 ) ( u 1 2 r 21 ρ 11 + u 1 u 2 r 22 ρ 21 ) 1 ( u 1 u 2 r 21 ρ 12 + u 2 2 r 22 ρ 22 ) ] [ d 1 d 2 ] = [ β 1 u 1 r 11 + β 2 u 2 r 12 β 1 u 1 r 21 + β 2 u 2 r 22 ] ,
[ a b c 1 c 2 d 1 d 2 ] = [ a c , 1 u 1 a c, 2 u 2 b c , 1 u 1 b c, 2 u 2 u 1 ρ 11 u 2 ρ 12 u 1 ρ 21 u 2 ρ 22 1 0 0 1 ] [ d 1 d 2 ]  +  [ a s b s β 1 β 2 0 0 ] ,
Φ SPP , i = | b i | 2 Σ S SPP , i n d s ,
Ψ ( r , ω ) = [ Ψ s,0 ( r , ω ) β 0 ( ω ) Ψ RGSP, +  ( r , ω ) ] + Ψ SPP ( r , ω ) b ( ω ) ,
det [ K ( ω ) ] = 0.
Ψ QNM , m ( r ) = Ψ SPP ( r , ω ~ m ) b m = Ψ SPP ( r , ω ~ m ) [ F ( ω ~ m ) d m ] .
d ( ω ) = K ( ω ) β ( ω ) det [ K ( ω ) ] ,
b ( ω ) = F ( ω ) K ( ω ) β ( ω ) det [ K ( ω ) ] + s ( ω ) .
Ψ ( r , ω ) = [ Ψ s,0 ( r , ω ) β 0 ( ω ) Ψ RGSP, +  ( r , ω ) ] + Ψ SPP ( r , ω ) { F ( ω ) K ( ω ) β ( ω ) det [ K ( ω ) ] + s ( ω ) } .
Ψ ( r , ω ) Ψ ( r , 0 ) + m = 1 M ω / ω ~ m ω ω ~ m p m ( r ) ,
p m ( r ) = lim ω ω ~ m ( ω ω ~ m ) Ψ SPP ( r , ω ) { F ( ω ) K ( ω ) β ( ω ) det [ K ( ω ) ] } = Ψ SPP ( r , ω ~ m ) F ( ω ~ m ) d m c m T β ( ω ~ m ) { det [ K ( ω ) ] / ω } ω = ω ~ m = c m T β ( ω ~ m ) { det [ K ( ω ) ] / ω } ω = ω ~ m Ψ QNM , m ( r ) .
Ψ ( r , ω ) m = 1 M α m ( ω ) Ψ QNM, m ( r ) ,
α m ( ω ) = ω / ω ~ m ω ω ~ m c m T β ( ω ~ m ) { det [ K ( ω ) ] / ω } ω = ω ~ m .
b ( ω ) m = 1 M α m ( ω ) b 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.