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

Schottky-contact plasmonic dipole rectenna concept for biosensing

Open Access Open Access

Abstract

Nanoantennas are key optical components for several applications including photodetection and biosensing. Here we present an array of metal nano-dipoles supporting surface plasmon polaritons (SPPs) integrated into a silicon-based Schottky-contact photodetector. Incident photons coupled to the array excite SPPs on the Au nanowires of the antennas which decay by creating ”hot” carriers in the metal. The hot carriers may then be injected over the potential barrier at the Au-Si interface resulting in a photocurrent. High responsivities of 100 mA/W and practical minimum detectable powers of −12 dBm should be achievable in the infra-red (1310 nm). The device was then investigated for use as a biosensor by computing its bulk and surface sensitivities. Sensitivities of ∼ 250 nm/RIU (bulk) and ∼ 8 nm/nm (surface) in water are predicted. We identify the mode propagating and resonating along the nanowires of the antennas, we apply a transmission line model to describe the performance of the antennas, and we extract two useful formulas to predict their bulk and surface sensitivities. We prove that the sensitivities of dipoles are much greater than those of similar monopoles and we show that this difference comes from the gap in dipole antennas where electric fields are strongly enhanced.

© 2013 Optical Society of America

1. Introduction

Surface plasmon polaritons (SPPs) are of great interest in a wide range of fields from physics, chemistry, materials processing, to biology. SPPs are electromagnetic surface waves coupled to free electron oscillations propagating along a metal-dielectric interface at optical wavelengths [1]. SPPs have many interesting properties such as highly-enhanced fields [2], strong confinement (to sub-wavelength scale) [3, 4] and high bulk and surface sensitivities [57]. These characteristics are of interest to several applications such as waveguides [5, 8, 9], sensors [10], and nonlinear optics [11].

Advances in nanofabrication have made periodic metallic nanostructures and nanoantennas attractive for practical uses [2, 8, 12]. SPP biosensors have become key tools for the investigation of biomolecular interactions and for detection applications. Piliarik emphet al. measured the localized surface plasmon resonance shift of an arrays of chains of metallic nanorods in response to bulk refractive index changes as 140 nm/RIU [13]. Knight et al. reported an active optical monopole array which can be considered as a highly compact, wavelength-specific, and polarization-specific light detector appropriate for the sensing applications [14]. Rodriguez et al. reported a chemosensor fabricated from gold nano-crosses exhibiting a strongly localized SPP in the infrared and obtained a high bulk sensitivity of 740 nm/RIU [6]. Tsai emphet al. described an optical sensor consisting of square-lattice slab-like gold nano-rings, reporting a bulk sensitivity of 691 nm/RIU [7]. Star-shaped nanoparticles suitable for microscopic imaging exhibited a wavelength peak shift of 665 nm/RIU [12]. Rice-shaped nanoparticles are reported to have even larger sensitivity but with a broader spectral response [15]. The bulk sensitivity of a rod-like metal nanoparticle surface plasmon resonance is reported in [16]; the natural resonance frequency of this structure strongly depends on its aspect ratio and thickness: narrow rods with a high aspect ratio have lower resonant frequencies. Bulk sensitivities reported for cylindrical rods of aspect ratio ranging from 0.5 to 3 are from 150 to 450 nm/RIU. Mazzotta et al. measured photocurrent ratio sensitivity of a nanoplasmonic biosensor chip with integrated electrical detection (an array of gold nanodisks) using hole-mask colloidal lithography and they achieved a value around 133 nm/RIU as of it biosensing property [17]. Guyot et al. developed miniaturized silicon based nanoplasmonic biosensing platform as a symmetric nanohole array structure and got the sensitivity of 4×10−5 RIU which is greatly improved and promising for the application in portable nanoplasmonic multisensing and imaging [18]. The bulk sensitivity was also calculated by a quasistatic theory which serves as an upper bound to sensitivities of nanoparticles on dielectric substrates such as those associated with biomolecule sensing. Achieving a sharp resonant peak (which is easier to track) is a difficult challenge with most plasmonic nanos-tructures. A multipixel array of Fano-resonant asymmetric metamaterial is presented in [19], which is capable of confining light to nanoscale regions with a very sharp resonant peak. The narrower linewidth of this structure can also increase the electric field enhancement in regions of the metamaterial, which improves the sensitivity.

A Schottky barrier photodetector formed at the interface between a metal and a lightly doped semiconductor can be used to detect infrared radiation below the bandgap energy of the substrate via the internal photoelectric effect (IPE) [20]. Such detectors involving SPP excitations in grating [21], prism [22], waveguide [2325] and antenna [14, 26] geometries have been reported. The combination of a classical (microwave) antenna with a rectifier, termed a ”rectenna”, is characterized (in part) by its ability to convert microwave to dc power [27]. In this paper we propose a plasmonic rectenna array comprised of Au nanowires of rectangular cross-section operating as optical dipole antennas. The difference between dipole and monopole antennas is the existence of a gap in the former where highly localised fields are existed. The dipoles are inter-connected electrically (in an optically non-invasive manner), on a p-doped Si substrate forming a Schottky contact thereon. The antennas are assumed covered with H2O to represent use as a biochemical sensor. We investigate the spectral response of the rectenna as a function of the index of the superstrate and with various thicknesses of biochemical adlayer grown on the antennas. We combine numerical methods (FDTD, modal analysis) with an analytical model for IPE, and a transmission line model for the nano-dipole antennas to determine the performance characteristics of the proposed structure.

First the rectenna geometry and the methods applied in its analysis are discussed. Secondly, a particular geometry is considered and computed results are presented. Thirdly, we discuss the operation of the rectenna as a Schottky contact detector and we estimate its responsivity and detection limit. Finally, we assess the potential of the structure for bulk and surface (bio)chemical sensing.

2. Geometry

Figure 1(a) shows our rectenna concept, consisting of an array of Au nano-dipole antennas on Si, electrically interconnected via Au lines running perpendicularly to the dipole axes through the middle of each arm to a common circular contact pad; as the array is illuminated via x-polarized light, such interconnects do not affect the optical behavior of the array. The array is symmetric about the x and y axes. It is illuminated by a Gaussian beam focused onto the array through the substrate near λ0 = 1310 nm (Si is transparent at this wavelength) and its response monitored via the photocurrent generated by (IPE) using the Au contact pad (on top) and the Al Ohmic contacts (below). In a biosensing application, this arrangement is advantageous as it simplifies interrogation, because only the photocurrent need be monitored, and it separates the optics (bottom) from the micro-fluidics (top). A single dipole is depicted in Fig. 1(b) showing our definition of its dimensions and the coordinate system used for the analysis; a = 0 throughout except for the surface sensitivity computations.

 figure: Fig. 1

Fig. 1 (a) Array of Au dipoles on p-Si on p+-Si covered by H2O. Al Ohmic contacts are located at the bottom of the structure and an Au circular contact pad is connected to all dipole arms via optically non-invasive perpendicular Au interconnects. A plane wave source illuminates the array in the +z-direction from below. (b) Geometry of a unit cell of the array under study (interconnects are shown as well); the dipole is assumed covered by an adlayer of thickness a when the surface sensitivities are computed.

Download Full Size | PDF

The finite difference time domain (FDTD) method [28] with a 0.5 × 0.5 × 0.5 nm3 discreti-sation in the region around the antenna was used for the modeling along with Paliks material data [29] for Au and Si, and Segelsteins data [30] for H2O which takes the water absorption into consideration. Transmittance and reflectance monitors were placed 2.5μm above and below the Si/H2O interface, respectively, parallel to the xy plane. We consider the case of an array that is small compared to the waist of the Gaussian beam, so the beam can be considered as a plane wave in the simulations. Thus, an x-polarised plane wave source located 2.8μm below the interface illuminates the array along the +z direction. We assumed doping levels for p+-Si and p-Si of 10+18 cm−3 and 10+15 cm−3; the attenuation of a plane wave at λ0 = 1310 nm propagating through a 500μm thick p+-Si substrate of this doping level, calculated using the Drude model [31], is 0.2037 dB, which is negligible in this work.

3. Optical response

From Fig. 1(b), we define the nanoantenna length (l), width (w), thickness (t), and the gap length (g), as well as the vertical and horizontal distance (p, q) between any two adjacent nanoantennas. Also the interconnect width is labeled as w′. Good dipole (infinitely periodic, ”good dipole” means a dipole that resonates near a desired wavelength and that performs well in terms of enhancement and absorptance) dimensions were determined via modeling to be: w = 30nm, t = 30nm, l = 210nm, g = 20nm and p = q = 300nm. Interconnect dimensions were taken as w′ = 20nm and t = 30nm. All dimensions except for g remain constant throughout the paper. Figure 2(a)–2(c) show the electric field distribution of the dipole array at resonance over xy cross-sections slightly inside the Au close to the Si/Au interface, through the middle of the Au dipoles, and slightly above the Au surface in H2O. As noted, the field intensity is very high in the dipoles near the Au/Si interface (Fig. 2(a)) and in the gap (Fig. 2(b)), compared to along the dipole arms and the ends. Localized fields in the gap make small-gap antennas highly sensitive to changes in this region (as discussed below), and strong fields in the Au near the Au/Si interface are desirable to enhance IPE; both attributes are useful for the envisaged biosensors. It is evident that the Au interconnects have a negligible influence on the field distribution, and that the periodicity is large enough for coupling between two neighboring dipoles to be negligible. Given the uniform illumination and their length, the dipoles resonate in their lowest order bonding mode [32], in the sab0 mode propagating along the nanowire waveguides forming the dipole arms [4, 5].

 figure: Fig. 2

Fig. 2 Distribution of |E|=|Ex|2+|Ey|2+|Ez|2 on xy cross-sectional planes for an Au dipole of dimensions w = 30 nm, t = 30 nm, l = 210 nm, g = 20 nm and p = q = 300 nm (a) 3 nm above the Au/Si interface, (b) 15 nm above the Au/Si interface, and (c) 3 nm above the Au surface in H2O. Computations performed at λ0 = 1353 nm (resonance wavelength).

Download Full Size | PDF

The transmittance T is calculated as a function of frequency (f) or wavelength via:

T(f)=SRe(Pm).dsSRe(Ps).ds
where Pm and Ps are Poynting vectors at the location of the monitor where T is calculated and at the location of the source, respectively. S is the surface of the reference plane where the transmittance is calculated [28]. Eq. (1) can also be used to compute the reflectance R of the system by replacing S with the appropriate reference plane.

Figure 3 plots the calculated transmittance (T), reflectance (R), absorptance (A), and the electric field enhancement Een of the array over wavelength, where A is given by:

A=1TR
and the electric field enhancement Een is calculated at the center of the gap, 15 nm above the Si/H2O interface, relative to the electric field at the same location in the absence of the antenna. From Fig.3 it is noted that Een peaks near resonance to a value of about 25. (Throughout this paper the resonant wavelength of the antennas (λ0r) refers to the free-space wavelength at which the absorptance curve reaches its maximum value.) The misalignment of extrema in the T, R, and A responses is due to absorption in Au. The absorptance of the system is due almost entirely to absorption in Au (losses are negligible in Si and H2O by comparison) [4].

 figure: Fig. 3

Fig. 3 Calculated transmittance (T), reflectance (R), absorptance (A) and electric field enhancement (Een) vs. free space wavelength (λ0).

Download Full Size | PDF

One can relate the optical performance and response of the dipoles to their geometry via the full parametric study reported in [4]. Decreasing the gap size changes the transmittance and reflectance as the proportion of Au covering the surface changes. Also, decreasing the gap size increases the capacitance causing a red-shift in the resonance of the system. Finally, decreasing the gap size increases the electric field therein, which as we show below, increases the bulk and surface sensitivities of the dipoles.

4. Quantum efficiency and responsivity

The Au dipoles are assumed to form a Schottky contact to Si thereby creating an infrared pho-todetector operating on the basis of IPE (Fig. 1(a)). Hot (energetic) conduction carriers are created in the Au dipoles by absorption of SPPs resonating on the antennas. If the photon energy is greater than the Schottky barrier energy, then hot carriers may be emitted over the barrier and collected in the Si as photocurrent. The internal quantum efficiency ηit is the number of hot carriers that contribute to the photocurrent per absorbed photon per second. We assumed p-Si because Schottky barriers are lower thereon than on n-Si, leading to increased quantum effi-ciency and detection at longer wavelengths. Also the effective Richardson coefficient for holes is lower than that for electrons which helps manage the dark current. We have used the thin-film model described in [33] to estimate the internal quantum efficiency, taking the attenuation length of hot holes in Au as 55 nm [34] and the Schottky barrier height as ϕB = 0.34 eV for Au on p-Si [35]. Figure 4(a) shows ηit computed over the wavelengths of interest.

 figure: Fig. 4

Fig. 4 (a) Internal quantum efficiency ( ηit) vs. λ0. (b) Responsivity (Resp) and minimum detectable power (Smin) vs. λ0.

Download Full Size | PDF

Taking all Au/Si contacts into consideration (a 25×25 network of dipoles and their related interconnects, and a circular contact pad of diameter 10μm), a total contact area of ∼ 95.5×10−12 m2 is obtained. Consequently, a dark current of Id = 5.4μA is obtained [35] at 300 K. The rectenna and interconnect area is ∼ 0.17 × 10−12 m2 and the contact pad area is 95.45 × 10−12 m2, so the pad area dominates; the dark current associated with the latter could be eliminated by inserting a dielectric between the pad and the Si surface.

The responsivity Resp is defined as the ratio of the photocurrent to the incident optical power and can be expressed in terms of ηit as [36]:

Resp=κAηitqhν
where A is the absorptance, taken as that of the dipole array (Eq. (2)), and κ is the fraction of the absorptance that contributes to the photocurrent [36] (taken as κ = 1 herein). The minimum detectable power is given as the ratio of the dark current to the responsivity Smin = Id/Resp.

The detection performance of the rectenna was assessed over a range of wavelengths and is plotted in Fig. 4(b). Compared to other SPP Schottky detectors, the responsivity and minimum detectable power of this detector are quite reasonable, even with a large contact pad [23, 25]. This performance level is suitable for low-cost silicon based detection and for the intended biosensing, optical monitoring and interconnect applications. The wavelength response follows closely that of the absorptance of the dipole array (Fig. 4(b)), except that the former drops more rapidly at longer wavelengths due to the decrease in internal quantum efficiency ( ηit) as shown in Fig. 4(a).

5. Bulk sensitivity

Monitoring changes in a surface plasmon resonance is a widely used interrogation approach for measuring changes in the refractive index of bulk solutions and changes in the thickness or composition of thin bio(chemical) adlayers on the biosensor surface.

For bulk sensing, the index of the cover material nc was changed from the nominal one (H2O) over a large range, and absorptance spectra were computed using the FDTD method and the same mesh size as in the first section of this paper (0.5×0.5×0.5 nm3). Figure 5(a) shows A for all test cases considered. Increasing nc decreases A and increases the electric field enhancement (not shown). The bulk sensitivity ∂λ0r/∂nc of the rectenna and the peak responsivity (Resp,r) are plotted in Fig 5(b). The bulk sensitivity was computed by interpolating the spectral shifts using a cubic spline, then computing the derivative of the spline.

 figure: Fig. 5

Fig. 5 (a) Absorptance (A) vs. λ0 for several cover refractive indices nc ranging from 1 to 2.75). (b) Bulk sensitivity (∂λ0r/∂nc - blue) and peak responsivity (Resp,r - red) of the rectenna as a function of nc.

Download Full Size | PDF

The rectenna has a bulk sensitivity comparable to what is reported in [6, 7, 13]. Recent analytical studies show that the bulk sensitivity increases with the surface plasmon resonance wavelength and the upper limit of this sensitivity is totally independent of nanostructure shape [37]; the results of Fig. 5 are consistent with this trend. The peak responsivity drops linearly with nc following the peak absorptance, with the largest peak responsivity occurring for nc = 1. The drop in responsivity would translate to a commensurate drop in the monitored photocurrent as the bulk index changes. An incident optical power of 1 mW would yield a change in photocurrent of ∼ 40μA over the full index range, which is easily measureable.

5.1. Waveguide modal analysis

Modal analysis was carried out on the waveguide used as the arms of the dipoles using a mode solver based on the finite-difference method [26] in order to identify the mode propagating thereon. The wavelength for the modal analysis was set to λ0 = 1353 nm, which is the resonant wavelength of the rectenna covered by H2O. Figure 6(a) and (b) show the real part of the main transverse electric field (Ez) of the mode, plotted over the yz cross-section of one of the dipole arms for the extreme cases of bulk index used in the sensitivity study, i.e., nc = 1 and nc = 2.75. The distribution of Ez identifies the mode propagating along the nanowires as the sab0 mode [9, 32], which is excited on the dipoles given the polarization, orientation and uniformity of the source field (x-polarised plane wave). The field distribution is clearly perturbed by changes in the cover index (Fig. 6).

 figure: Fig. 6

Fig. 6 Real part of Ez of the sab0 mode plotted over the cross-section of a nanowire waveguide (λ0 = 1353 nm) computed using a mode solver. (a) nc = 1 (air) and (b) nc = 2.75.

Download Full Size | PDF

We compute the effective refractive index (neff) and the mode power attenuation (α) of the sab0 mode as a function of nc and plot the results in Fig. 7. Increasing nc increases the effective refractive index which causes, in part, the red shift observed in the absorptance curves of Fig. 5(a) (λ0rnc); the red shift is also due to changes in the gap capacitance Cg as nc changes. The mode power attenuation is also observed to increase with increasing nc explaining in part the broadening of the responses of Fig. 5(a) as nc increases.

 figure: Fig. 7

Fig. 7 Effective refractive index (neff blue) and mode power attenuation (α - red) of the sab0 mode resonating along the dipoles as a function of nc.

Download Full Size | PDF

5.2. Analytical expression for the bulk sensitivity of dipoles

An equivalent circuit was proposed in [4] to relate the resonant wavelength of dipole antennas to the effective index (neff) of the sab0 mode propagating along the nanowires forming the dipole, and to the characteristics of the dipole gap. We briefly summarize this circuit as it will be required in what follows. The circuit consists of two open-circuited transmission lines, one for each arm of the dipole, connected by a capacitor modeling the effects of the gap. The resonant frequencies of dipole antennas ω0r satisfy the following transcendental equation derived for the circuit:

tan(neffω0rε0μ0(d+δm))=2ω0rCgZ0
where
ω0r=2πc0λ0r

In Eq. (4), d + δm is the effective length of each transmission line section, where d = (lg)/2. The parameter δm is the distance from the end of each arm where the main field component decays to 1/e; in our dipole design δm ≈ 21 nm, taken as the decay of Ez (the main transverse electric field component of the sab0 mode - Fig. 6). Neglecting fringing fields, the capacitance of the gap Cg is:

Cg=εcAdg
where εc = ε0εr,c and εr,c are the absolute and relative permittivities of the material filling the gap at the wavelength of modal analysis, which is taken as a wavelength near the expected resonant wavelength of the dipole antenna. Ad = wt is the cross-sectional area of the dipole arms. In Eq. (4), Z0 is the characteristic impedance of the transmission lines calculated as:
Z0=f(y,z)Re[Zω(y,z)]dSf(y,z)dS
where Zω is the wave impedance of the sab0 mode:
Zω=k^(E×H*)(k^×H)(k^×H*)
k̂ = x̂ is the direction of modal propagation which is along the dipole axis, and E and H are the mode fields. In Eq. (7), f(y,z) is a weighting function defined from the sab0 mode fields as:
f(x,y)=|Ey(y,z)|2+|Ez(y,z)|2

Figure 8 plots the gap capacitance as a function of the bulk index. The red-shift in the absorptance curves are caused in part by the increase in Cg which is noted to increase from 0.4 to 3 aF. Figure 8 also plots Z0 as a function of nc, showing a non-negligible variation (compared to the surface sensing case as discussed below).

 figure: Fig. 8

Fig. 8 Gap capacitance Cg (blue) and characteristic impedance Z0 of the sab0 mode (red) as a function of nc.

Download Full Size | PDF

We use now the equivalent circuit model to gain insight on the bulk sensing performance of the dipoles. Taking the derivative of both sides of Eq. (4) with respect to nc yields:

[(d+δm)ε0μ0ω0rneffnc+(d+δm)ε0μ0neffω0rnc]×[1+tan2(neffω0rε0μ0(d+δm)]=2ω0rCgZ0nc2ω0rZ0Cgnc2CgZ0ω0rnc
where we note that neff = neff (nc), ω0r = ω0r(nc), δm = δm(nc), Cg = Cg(nc) and Z0 = Z0(nc). In writing Eq. (10) we have supposed that ∂δm/∂nc ≈ 0, which is justified based on verifications carried out at our extreme values for nc. After some manipulations, the following equation for the bulk sensitivity is obtained from the above:
ζωω0rnc=ζnneffnc+ζCCgnc+ζZZ0nc
where
ζω=(d+δmc0)neff(1+4ω0r2Cg2Z02)+2CgZ0
ζn=(d+δmc0)neff(1+4ω0r2Cg2Z02)
ζC=2ω0rZ0
ζZ=2ω0rCg
and c0 is the speed of light in vacuum. The term ∂Cg/∂nc in Eq. (11) can be expressed in terms of dipole dimensions and gap material properties. Ignoring the imaginary part of the bulk material permittivity ( εr,cnc2) this term can be simplified to:
Cgnc=(wtg)εcnc=2ε0(wtg)nc
Substituting Eqs. (12) to (16) into Eq. (11), the expression for the bulk sensitivity becomes:
ω0rnc=ζω1[4(wtg)ω0rZ0]nc+ζω1(ζnneffnc+ζZZ0nc)

The second term in Eq. (17) comes from mode properties and the first term is directly related to the gap capacitance. In Eq. (12), 2CgZ0((d+δm)/c0)neff(1+4ω0r2Cg2Z02); neglecting this term and substituting in the expressions for ζω, ζn and ζZ from Eqs. (12) to (15) into Eq. (17), leads to the simplified expression for the bulk sensitivity:

ω0rnc[2wtc0ω0rg(d+δm)neff]2Z0nc+Z0ncnc21+(2wtg1ω0rZ0)2nc2ω0rneffneffnc

The bulk sensitivity can be written in terms of the resonant wavelength using:

λ0rnc=(2πc0ω0r2)ω0rnc
which yields
λ0rnc[2wtc0λ0rg(d+δm)neff]2Z0nc+Z0ncnc21+(4πwtg1c0λ0rZ0)2nc2+λ0rneffneffnc

Figure 9(a) compares the resonant wavelengths computed using the FDTD method (Fig. 5(a)) with those generated by the transmission line model (Eq. (4) using the corresponding modal parameters (neff, Z0) computed at the resonant wavelength of each FDTD test case. From Fig. 9(a) one observes that for small bulk indices, and perhaps for very large ones, the two curves diverge slightly from each other. However over most of the index range, the transmission line model agrees very well with the FDTD results. Based on these curves we compute ∂λ0r/∂nc by numerical differentiation of cubic spline interpolants as in Fig. 5(b) and plot as Fig. 9(b). The black curve plots the bulk sensitivity calculated analytically using Eq. (20). For all values of nc the analytical solution (Eq. (20)) agrees very well with the transmission line model (Eq. 4), and both agree very well with the FDTD results for large values of nc. The agreement deteriorates at lower values of nc which probably comes from neglecting ∂δm/∂nc and the fringing fields in the model for the gap capacitance (Eq. (6)). Still the transmission line model (Eq. (4)) and the analytical solution (Eq. (20)) can be used over a broad range of bulk indices to estimate performance rather than the more time-consuming FDTD computations.

 figure: Fig. 9

Fig. 9 (a) Resonant wavelengths computed using the transmission line model and the FDTD method as a function of bulk index nc. (b) Bulk sensitivity computed using the FDTD method (dashed blue), the transmission line model (Eq. (4) - red), and the analytical solution (Eq. (20) - black).

Download Full Size | PDF

One interesting and important property extracted from Fig. 9(b) is that we can fit a linear model to the black curve as ∂λ0r/∂nc ≈ 81nc + 199 (nm/RIU). Strictly, the slope and intercept of this linear model are functions of λ0r and the modal parameters neff and Z0 which change with nc; however, the model fits well over the range of nc investigated here. Comparing Eq. (20) with a linear model reveals that the slope term (81 nm/RIU2) comes from the gap capacitance and the intercept term (199 nm/RIU) mostly from perturbation of the mode fields.

6. Surface sensitivity

The surface sensitivity can be defined in a number of ways [5, 38, 39]. The definition adopted here is ∂λ0r/∂a where a is the thickness of the adlayer of refractive index na (Fig. 1(b)). The ad-layer is modeled as uniform dielectric layer with its thickness and index (a, na) being effective parameters [5]. The adlayer index is taken as na = 1.45, which is representative of biochemical material [5], and it is assumed to grow on all Au/H2O interfaces.

Previous studies [4, 40] indicate that the electric field enhancement is very sensitive to the gap size. Specifically, decreasing the gap increases the field enhancement in the gap, which in turn increases the overlap between the resonant mode and the adlayer, leading to increased sensitivity [41]. We therefore reduced the gap size from 20 to 10 nm and computed a series of absorptance curves as a increases from 0 to 5 nm using the FDTD method and a very small mesh size (0.2 × 0.2 × 0.2 nm3), as plotted in Fig. 10(a). For the case a = 5 nm the gap is fully loaded with adlayer and the shift in resonance relative to the no adlayer case (a = 0) is 30 nm.

 figure: Fig. 10

Fig. 10 (a) Absorptance A vs. λ0 for several adlayer thicknesses (a = 0 to 5 nm); the curves are offset vertically by −0.05 for clarity. (b) Surface sensitivity (∂λ0r/∂a - blue) and peak responsivity (Resp,r - red) as a function of a.

Download Full Size | PDF

Fig. 10(b) gives ∂λ0r/∂a computed by interpolating the spectral shifts of Fig. 10(a) using a cubic spline, then computing the derivative of the spline. The largest surface sensitivity observed is ∂λ0r/∂a ∼ 8 nm/nm, occurring in the early stages of adlayer growth, and then dropping off significantly as the adlayer fills the gap region. The sensitivity is due mainly to the enhanced electric field in the gap and near the ends of each dipole where the enhanced fields overlap much more with the adlayer. Peak responsivity (Resp,r) values are also plotted in Fig. 10(b); an incident optical power of 1 mW would yield a change in photocurrent of ∼ 1μA over the full adlayer range, which should be readily measureable.

6.1. Waveguide modal analysis

Figure 11(a) and (b) show the real part of the main transverse electric field (Ez) of the mode, plotted over the yz cross-section of one of the dipole arms for the extreme cases of surface adlayer thickness used in the sensitivity study, i.e., a = 0 (no adlayer) and a = 5 nm. The distribution of Ez identifies the mode propagating along the nanowires as the sab0 mode [9, 30], which is excited on the dipoles given the polarization, orientation and uniformity of the source field (x-polarised plane wave). The field distribution is slightly perturbed by changes in the adlayer thickness (Fig. 11).

 figure: Fig. 11

Fig. 11 Real part of Ez of the sab0 mode plotted over the cross-section of a nanowire waveguide (λ0 = 1353 nm) computed using a mode solver. (a) a = 0 (no adlyaer) and (b) a = 5 nm.

Download Full Size | PDF

We have therefore computed the parameters of this mode (neff and α) as a function of a and we plot the results in Fig. 12. Increasing a decreases slightly neff and α of the mode.

 figure: Fig. 12

Fig. 12 Effective refractive index (neff blue) and mode power attenuation (α - red) of the sab0 mode resonating along the dipoles as a function of a.

Download Full Size | PDF

6.2. Analytical expression for the surface sensitivity of dipoles

The red shift in the absorptance curves is due to changes in the parameters of the sab0 mode resonating along the dipoles (Fig. 12) and to changes in the gap capacitance as the adlayer grows. The equivalent transmission line circuit described in the previous section can be used to take both effects into account using a new expression for Cg. Adding the adlayer produces three capacitors in series within the gap. Two capacitors (C1) are due to the adlayers growing on the ends of the right and left arms, and the third (C2) is formed by H2O filling the gap, as sketched in Fig. 13.

 figure: Fig. 13

Fig. 13 Schematic of a dipole gap showing three plate capacitances in series.

Download Full Size | PDF

From circuit theory, the total gap capacitance is calculated as:

Cg=(2C11+C21)1
where C1 and C2 are:
C1=εa(wta)
and
C2=εc(wtg2a)

In the above, εc = ε0εr,c and εa = ε0εr,a are the absolute permittivities of the bulk solution (H2O) and the adlayer, respectively. We ignore the imaginary parts of the permittivities and take εr,cnc21.3221.74 and εr,ana21.4522.11. After some manipulations, we write Cg as:

Cg=εaεcwt2(εaεc)a+εag

Figure 14 plots Cg (Eq. (24) and Z0 of the sab0 mode (Eqs. (7) to (9) as a function of the adlayer thickness a, showing that Cg increases and Z0 decreases slightly with a.

 figure: Fig. 14

Fig. 14 Gap capacitance Cg (blue) and characteristic impedance Z0 of the sab0 mode (red) as a function of a.

Download Full Size | PDF

Taking the derivative of Eq. (4) with respect to a yields:

ζωω0ra=ζnneffa+ζCCga+ζZZ0a
where ζω, ζn, ζC and ζZ are defined in Eqs. (12) to (15). The term ∂Cg/∂a in Eq. (25) can be evaluated using Eq. (24) and expressed in terms of dipole dimensions and the properties of the materials filling the gap (emphi.e., the adlayer and H2O) as:
Cga=2ε0εr,cεr,a(εr,aεr,c)wt[2(εr,aεr,c)a+εr,ag]2

The above reveals that ∂Cg/∂a is a saturating nonlinear function of the adlayer thickness a. Substituting Eq. (26) into Eq. (25), we obtain the following expression for the surface sensitivity as a function of a:

ω0ra=2ε0εr,cεr,a(εr,aεr,c)wt[2(εr,cεr,a)a+εr,ag]2ζω1ζCω0rZ0+ζω1(ζnneffa+ζZZ0a)

The surface sensitivity can be written in terms of the resonant wavelength using:

λ0ra=(2πc0ω0r2)ω0ra
which yields:
λ0ra=4ε0εr,cεr,a(εr,cεr,a)wt[2(εr,cεr,a)a+εr,ag]2ζω1ζCλ0rZ0ζω1λ0r22πc0(ζnneffa+ζZZ0a)

From Fig. 12, it is clear that ∂neff/∂a and ∂Z0/∂a are approximately constant (∼ −0.0014 nm−1 and 1Ω/nm) over the range of a considered. Thus from Eq. (29) we find that ∂λ0r/∂a is also a saturating nonlinear function of a (which follows from Eq. (24)).

Figure 15(a) compares the resonant wavelengths obtained via the FDTD (Fig. 10(a)) and with those obtained by solving Eq. (4) using the corresponding modal parameters (neff, Z0) computed at the resonant wavelength of every FDTD test case. Based on these results we compute ∂λ0r/∂a (by numerical differentiation of spline interpolants) and plot this as Fig. 15(b). The black curve plots the surface sensitivity calculated using the analytical solution, Eq. (29). For all adlayer thicknesses, the analytical solution (Eq. (29)) agrees very well with the transmission line model (Eq. (4)), and both are in reasonable agreement with the FDTD computations. The error probably comes from neglecting ∂δm/∂a and the fringing fields in the model for the gap capacitance (Eq. (24)). Still the transmission line model (Eq. (4)) and the analytical solution (Eq. (29)) can be used over a broad range of adlayer thicknesses to estimate the surface sensitivity of dipoles rather than the more time-consuming FDTD method. The results from the transmission line model differ from the FDTD results by a maximum of 25%.

 figure: Fig. 15

Fig. 15 (a) Resonant wavelengths computed using the transmission line model and the FDTD method as a function of adlayer thickness a. (b) Surface sensitivity computed using the FDTD method (dashed blue), the transmission line model (Eq.(4) - red), and the analytical solution (Eq.(29) - black).

Download Full Size | PDF

7. Comparison of dipole and monopole sensitivities

Using the transmission line model we derive the bulk and surface sensitivities of monopole antennas and compare them to those of dipole antennas. Replacing all dipoles with monopoles of comparable length (d = 2l and g = 0), we find the following equation for the resonant wavelengths of the equivalent transmission line circuit:

cot(neffω0rε0μ0(d+δm))=0
where the symbols retain the same meaning as in Eq. (11). We note, of course, that there is no lumped capacitance in this model because there is no gap. By taking the derivative of the above with respect to nc and a we find the following expressions for the bulk and surface sensitivities:
λ0rnc=λ0rneffneffnc
λ0ra=λ0rneffneffa

We note that a monopole corresponds to a dipole in the limit of g → 0, which leads to an infinite gap capacitance, i.e. Cg → ∞. In this limit, ζω, ζn and ζZ become very large (Eqs. (13), (14) and (15)) and Eqs. (20) and (29) simplify to Eqs. (31) and (32).

Comparing Eqs. (31) and (32) with the corresponding ones for dipoles (Eqs. (20) and (29)), we observe that the sensitivities of monopole antennas are smaller than those of dipole antennas. This is due to the absence of the gap in the former; thus, monopole resonances are altered only by changes in the properties of the sab0 mode resonating on the antenna, whereas in dipoles, changes in the cap capacitance lead to a larger shift in resonance. Also, the electric field is strongly enhanced in the gap of dipoles compared to other locations on the antenna. The bulk sensitivity of dipoles is at least 1.65× greater than that of comparable monopoles. Similarly the surface sensitivity of dipoles is at least 1.5× greater than that of monopoles. In both cases, the importance of the gap is manifest.

8. Conclusion

Arrays of Au nano-dipole antennas forming a Schottky contact to Si were proposed and investigated theoretically for use as a wavelength-selective sub-bandgap rectenna and biosensor. Photodetection occurs via the internal photoelectric effect for photon energies above the Schot-tky barrier height but below the bandgap of Si. Excitation by a normally-incident plane wave polarized along the length of the dipoles was considered. Under this excitation, non-invasive electrical contacts to each dipole arm can be introduced, facilitating collection of the photocur-rent. The plane-wave source excites a specific strongly-confined SPP mode ( sab0[9]) propagating along the arms of the dipole antennas, forming a bonding resonance as the main resonance of operation of the antennas. Utilizing a Schottky detector structure simplifies interrogation in that the transmittance or reflectance need not be measured; rather only the photocurrent collected at the contacts needs to be monitored. Illuminating the antennas through the substrate leaves the top side of the device available for integrating microfluidics.

As a rectenna, high responsivities (100 mA/W) and practical minimum detectable powers (−12 dBm) are predicted at wavelengths near 1310 nm. As a biosensor, the rectenna structure offers significant advantages as it simplifies the interrogation set-up: only the photocurrent generated by the device need be monitored, and the excitation optics (bottom) are physically separated from the micro-fluidics (top) by the device. Compelling sensitivities are predicted for the dipoles: 250 nm/RIU in bulk sensitivity and 8 nm/nm in surface sensitivity (biochemical adlayer in H2O). In terms of changes in photocurrent, an incident optical power of 1 mW would yield a change of ∼ 1μA as an adlayer grows to 5 nm in thickness, which is readily measureable.

Using a transmission line model to represent the dipole antennas, we derive analytical expressions for the bulk and surface sensitivities of the structure and validate them via comparisons with numerical results. We proved using these expressions that dipole antennas provide much greater bulk and surface sensitivities than monopole antennas because of the existence of the gap in the former, which introduces a capacitance (and enhanced fields) that strongly affects the resonance.

References and links

1. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007), 1st ed.

2. B. Hecht, H. Bielefeldt, L. Novotny, Y. Inouye, and D. W. Pohl, “Local excitation, scattering, and interference of surface plasmons,” Phys. Rev. Lett. 77, 1889–1892 (1996). [CrossRef]   [PubMed]  

3. W. L. Barnes, T. W. Preist, S. C. Kitson, and J. R. Sambles, “Physical origin of photonic energy gaps in the propagation of surface plasmons on gratings,” Phys. Rev. B 54, 6227–6244 (1996). [CrossRef]  

4. S. S. Mousavi, P. Berini, and D. McNamara, “Periodic plasmonic nanoantennas in a piecewise homogeneous background,” Opt. Express 20, 18044–18065 (2012). [CrossRef]   [PubMed]  

5. P. Berini, “Bulk and surface sensitivities of surface plasmon waveguides,” New Journal of Physics 10, 105010 (2008). [CrossRef]  

6. F. J. Rodriguez-Fortuno, M. Martinez-Marco, B. Tomas-Navarro, R. Ortuno, J. Marti, A. Martinez, and P. J. Rodriguez-Canto, “Highly-sensitive chemical detection in the infrared regime using plasmonic gold nanocrosses,” Applied Physics Letters 98, 133118 (2011). [CrossRef]  

7. C.-Y. Tsai, S.-P. Lu, J.-W. Lin, and P.-T. Lee, “High sensitivity plasmonic index sensor using slablike gold nanoring arrays,” Applied Physics Letters 98, 153108 (2011). [CrossRef]   [PubMed]  

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

9. P. Berini, “Plasmon-polariton waves guided by thin lossy metal films of finite width: Bound modes of symmetric structures,” Phys. Rev. B 61, 10484–10503 (2000). [CrossRef]  

10. E. M. Larsson, J. Alegret, M. Kll, and D. S. Sutherland, “Sensing characteristics of nir localized surface plasmon resonances in gold nanorings for application as ultrasensitive biosensors,” Nano Letters 7, 1256–1263 (2007). [CrossRef]   [PubMed]  

11. G. I. Stegeman, J. J. Burke, and D. G. Hall, “Nonlinear optics of long range surface plasmons,” Applied Physics Letters 41, 906–908 (1982). [CrossRef]  

12. C. L. Nehl, H. Liao, and J. H. Hafner, “Optical properties of star-shaped gold nanoparticles,” Nano Letters 6, 683–688 (2006). [CrossRef]   [PubMed]  

13. M. Piliarik, P. Kvasnička, N. Galler, J. R. Krenn, and J. Homola, “Local refractive index sensitivity of plasmonic nanoparticles,” Opt. Express 19, 9213–9220 (2011). [CrossRef]   [PubMed]  

14. M. W. Knight, H. Sobhani, P. Nordlander, and N. J. Halas, “Photodetection with active optical antennas,” Science 332, 702–704 (2011). [CrossRef]   [PubMed]  

15. H. Wang, D. W. Brandl, F. Le, P. Nordlander, and N. J. Halas, “Nanorice: a hybrid plasmonic nanostructure,” Nano Letters 6, 827–832 (2006). PMID: [PubMed]  . [CrossRef]  

16. M. M. Miller and A. A. Lazarides, “Sensitivity of metal nanoparticle surface plasmon resonance to the dielectric environment,” The Journal of Physical Chemistry B 109, 21556–21565 (2005). [CrossRef]  

17. F. Mazzotta, G. Wang, C. Hgglund, F. Hk, and M. P. Jonsson, “Nanoplasmonic biosensing with on-chip electrical detection,” Biosensors and Bioelectronics 26, 1131 – 1136 (2010). [CrossRef]   [PubMed]  

18. L. Guyot, A.-P. Blanchard-Dionne, S. Patskovsky, and M. Meunier, “Integrated silicon-based nanoplasmonic sensor,” Opt. Express 19, 9962–9967 (2011). [CrossRef]   [PubMed]  

19. C. Wu, A. B. Khanikaev, R. Adato, N. Arju, A. A. Yanik, H. Altug, and G. Shvets, “Fano-resonant asymmetric metamaterials for ultrasensitive spectroscopy and identification of molecular monolayers,” Nat Mater 11, 69–75 (2012). [CrossRef]  

20. M. Casalino, G. Coppola, M. Iodice, I. Rendina, and L. Sirleto, “Critically coupled silicon fabry-perot photode-tectors based on the internal photoemission effect at 1550 nm,” Opt. Express 20, 12599–12609 (2012). [CrossRef]   [PubMed]  

21. S. R. J. Brueck, V. Diadiuk, T. Jones, and W. Lenth, “Enhanced quantum efficiency internal photoemission detectors by grating coupling to surface plasma waves,” Applied Physics Letters 46, 915–917 (1985). [CrossRef]  

22. C. Daboo, M. Baird, H. H. N. Apsley, and M. Emeny, “Improved surface plasmon enhanced photodetection at an augaas schottky junction using a novel molecular beam epitaxy grown otto coupling structure,” Thin Solid Films 201, 9 – 27 (1991). [CrossRef]  

23. A. Akbari, R. N. Tait, and P. Berini, “Surface plasmon waveguide schottky detector,” Opt. Express 18, 8505–8514 (2010). [CrossRef]   [PubMed]  

24. S. Zhu, G. Q. Lo, and D. L. Kwong, “Theoretical investigation of silicide schottky barrier detector integrated in horizontal metal-insulator-silicon-insulator-metal nanoplasmonic slot waveguide,” Opt. Express 19, 15843–15854 (2011). [CrossRef]   [PubMed]  

25. I. Goykhman, B. Desiatov, J. Khurgin, J. Shappir, and U. Levy, “Locally oxidized silicon surface-plasmon schot-tky detector for telecom regime,” Nano Letters 11, 2219–2224 (2011). [CrossRef]   [PubMed]  

26. E. S. Barnard, R. A. Pala, and M. L. Brongersma, “Photocurrent mapping of near-field optical antenna resonances,” Nat Nano 6, 588–593 (2011). [CrossRef]  

27. J. McSpadden, L. Fan, and K. Chang, “Design and experiments of a high-conversion-efficiency 5.8-ghz rectenna,” Microwave Theory and Techniques, IEEE Transactions on 46, 2053 –2060 (1998). [CrossRef]  

28. FDTD Solutions (Lumerical Solutions Inc.).

29. E. D. Palik, Handbook of Optical Constants of Solids (Academic Press, 1985).

30. D. J. Segelstein, “The complex refractive index of water,” Master’s thesis, University of Missouri, Kansas City, Missouri, USA (1981).

31. R. Soref and B. Bennett, “Electrooptical effects in silicon,” Quantum Electronics, IEEE Journal of 23, 123 – 129 (1987). [CrossRef]  

32. P. Nordlander, C. Oubre, E. Prodan, K. Li, and M. I. Stockman, “Plasmon hybridization in nanoparticle dimers,” Nano Letters 4, 899–903 (2004). [CrossRef]  

33. C. Scales and P. Berini, “Thin-film schottky barrier photodetector models,” Quantum Electronics, IEEE Journal of 46, 633 –643 (2010). [CrossRef]  

34. R. N. Stuart, F. Wooten, and W. E. Spicer, “Mean free path of hot electrons and holes in metals,” Phys. Rev. Lett. 10, 119–119 (1963). [CrossRef]  

35. S. M. Sze and K. K. Ng, Physics of Semiconductor Devices (Wiley, 2006), 3rd ed. [CrossRef]  

36. A. Akbari, A. Olivieri, and P. Berini, “Sub-bandgap asymmetric surface plasmon waveguide schottky detectors on silicon,” Accepted for publication in Sel. Top. Quantum Electronics, IEEE Journal of (2013).

37. S. J. Zalyubovskiy, M. Bogdanova, A. Deinega, Y. Lozovik, A. D. Pris, K. H. An, W. P. Hall, and R. A. Potyrailo, “Theoretical limit of localized surface plasmon resonance sensitivity to local refractive index change and its comparison to conventional surface plasmon resonance sensor,” J. Opt. Soc. Am. A 29, 994–1002 (2012). [CrossRef]  

38. J. Homola, “Surface plasmon resonance sensors for detection of chemical and biological species,” Chemical Reviews 108, 462–493 (2008). PMID: [PubMed]  . [CrossRef]  

39. V. Brioude and O. Parriaux, “Normalised analysis for the design of evanescent-wave sensors and its use for tolerance evaluation,” Optical and Quantum Electronics 32, 899–908 (2000). [CrossRef]  

40. L. Tang, S. E. Kocabas, S. Latif, A. K. Okyay, D.-S. Ly-Gagnon, K. C. Saraswat, and D. A. B. Miller, “Nanometre-scale germanium photodetector enhanced by a near-infrared dipole antenna,” Nat Photon 2, 226–229 (2008). [CrossRef]  

41. R. F. Harrington, Time-Harmonic Electromagnetic Fields (McGraw-Hill, 1961), 1st ed.

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 (15)

Fig. 1
Fig. 1 (a) Array of Au dipoles on p-Si on p+-Si covered by H2O. Al Ohmic contacts are located at the bottom of the structure and an Au circular contact pad is connected to all dipole arms via optically non-invasive perpendicular Au interconnects. A plane wave source illuminates the array in the +z-direction from below. (b) Geometry of a unit cell of the array under study (interconnects are shown as well); the dipole is assumed covered by an adlayer of thickness a when the surface sensitivities are computed.
Fig. 2
Fig. 2 Distribution of | E | = | E x | 2 + | E y | 2 + | E z | 2 on xy cross-sectional planes for an Au dipole of dimensions w = 30 nm, t = 30 nm, l = 210 nm, g = 20 nm and p = q = 300 nm (a) 3 nm above the Au/Si interface, (b) 15 nm above the Au/Si interface, and (c) 3 nm above the Au surface in H2O. Computations performed at λ0 = 1353 nm (resonance wavelength).
Fig. 3
Fig. 3 Calculated transmittance (T), reflectance (R), absorptance (A) and electric field enhancement (Een) vs. free space wavelength (λ0).
Fig. 4
Fig. 4 (a) Internal quantum efficiency ( η i t) vs. λ0. (b) Responsivity (Resp) and minimum detectable power (Smin) vs. λ0.
Fig. 5
Fig. 5 (a) Absorptance (A) vs. λ0 for several cover refractive indices nc ranging from 1 to 2.75). (b) Bulk sensitivity (∂λ0r/∂nc - blue) and peak responsivity (Resp,r - red) of the rectenna as a function of nc.
Fig. 6
Fig. 6 Real part of Ez of the s a b 0 mode plotted over the cross-section of a nanowire waveguide (λ0 = 1353 nm) computed using a mode solver. (a) nc = 1 (air) and (b) nc = 2.75.
Fig. 7
Fig. 7 Effective refractive index (neff blue) and mode power attenuation (α - red) of the s a b 0 mode resonating along the dipoles as a function of nc.
Fig. 8
Fig. 8 Gap capacitance Cg (blue) and characteristic impedance Z0 of the s a b 0 mode (red) as a function of nc.
Fig. 9
Fig. 9 (a) Resonant wavelengths computed using the transmission line model and the FDTD method as a function of bulk index nc. (b) Bulk sensitivity computed using the FDTD method (dashed blue), the transmission line model (Eq. (4) - red), and the analytical solution (Eq. (20) - black).
Fig. 10
Fig. 10 (a) Absorptance A vs. λ0 for several adlayer thicknesses (a = 0 to 5 nm); the curves are offset vertically by −0.05 for clarity. (b) Surface sensitivity (∂λ0r/∂a - blue) and peak responsivity (Resp,r - red) as a function of a.
Fig. 11
Fig. 11 Real part of Ez of the s a b 0 mode plotted over the cross-section of a nanowire waveguide (λ0 = 1353 nm) computed using a mode solver. (a) a = 0 (no adlyaer) and (b) a = 5 nm.
Fig. 12
Fig. 12 Effective refractive index (neff blue) and mode power attenuation (α - red) of the s a b 0 mode resonating along the dipoles as a function of a.
Fig. 13
Fig. 13 Schematic of a dipole gap showing three plate capacitances in series.
Fig. 14
Fig. 14 Gap capacitance Cg (blue) and characteristic impedance Z0 of the s a b 0 mode (red) as a function of a.
Fig. 15
Fig. 15 (a) Resonant wavelengths computed using the transmission line model and the FDTD method as a function of adlayer thickness a. (b) Surface sensitivity computed using the FDTD method (dashed blue), the transmission line model (Eq.(4) - red), and the analytical solution (Eq.(29) - black).

Equations (32)

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

T ( f ) = S Re ( P m ) . ds S Re ( P s ) . ds
A = 1 T R
R esp = κ A η i t q h ν
tan ( n eff ω 0 r ε 0 μ 0 ( d + δ m ) ) = 2 ω 0 r C g Z 0
ω 0 r = 2 π c 0 λ 0 r
C g = ε c A d g
Z 0 = f ( y , z ) Re [ Z ω ( y , z ) ] d S f ( y , z ) d S
Z ω = k ^ ( E × H * ) ( k ^ × H ) ( k ^ × H * )
f ( x , y ) = | E y ( y , z ) | 2 + | E z ( y , z ) | 2
[ ( d + δ m ) ε 0 μ 0 ω 0 r n eff n c + ( d + δ m ) ε 0 μ 0 n eff ω 0 r n c ] × [ 1 + tan 2 ( n eff ω 0 r ε 0 μ 0 ( d + δ m ) ] = 2 ω 0 r C g Z 0 n c 2 ω 0 r Z 0 C g n c 2 C g Z 0 ω 0 r n c
ζ ω ω 0 r n c = ζ n n eff n c + ζ C C g n c + ζ Z Z 0 n c
ζ ω = ( d + δ m c 0 ) n eff ( 1 + 4 ω 0 r 2 C g 2 Z 0 2 ) + 2 C g Z 0
ζ n = ( d + δ m c 0 ) n eff ( 1 + 4 ω 0 r 2 C g 2 Z 0 2 )
ζ C = 2 ω 0 r Z 0
ζ Z = 2 ω 0 r C g
C g n c = ( w t g ) ε c n c = 2 ε 0 ( w t g ) n c
ω 0 r n c = ζ ω 1 [ 4 ( w t g ) ω 0 r Z 0 ] n c + ζ ω 1 ( ζ n n eff n c + ζ Z Z 0 n c )
ω 0 r n c [ 2 w t c 0 ω 0 r g ( d + δ m ) n eff ] 2 Z 0 n c + Z 0 n c n c 2 1 + ( 2 w t g 1 ω 0 r Z 0 ) 2 n c 2 ω 0 r n eff n eff n c
λ 0 r n c = ( 2 π c 0 ω 0 r 2 ) ω 0 r n c
λ 0 r n c [ 2 w t c 0 λ 0 r g ( d + δ m ) n eff ] 2 Z 0 n c + Z 0 n c n c 2 1 + ( 4 π wt g 1 c 0 λ 0 r Z 0 ) 2 n c 2 + λ 0 r n eff n eff n c
C g = ( 2 C 1 1 + C 2 1 ) 1
C 1 = ε a ( w t a )
C 2 = ε c ( w t g 2 a )
C g = ε a ε c w t 2 ( ε a ε c ) a + ε a g
ζ ω ω 0 r a = ζ n n eff a + ζ C C g a + ζ Z Z 0 a
C g a = 2 ε 0 ε r , c ε r , a ( ε r , a ε r , c ) w t [ 2 ( ε r , a ε r , c ) a + ε r , a g ] 2
ω 0 r a = 2 ε 0 ε r , c ε r , a ( ε r , a ε r , c ) w t [ 2 ( ε r , c ε r , a ) a + ε r , a g ] 2 ζ ω 1 ζ C ω 0 r Z 0 + ζ ω 1 ( ζ n n eff a + ζ Z Z 0 a )
λ 0 r a = ( 2 π c 0 ω 0 r 2 ) ω 0 r a
λ 0 r a = 4 ε 0 ε r , c ε r , a ( ε r , c ε r , a ) w t [ 2 ( ε r , c ε r , a ) a + ε r , a g ] 2 ζ ω 1 ζ C λ 0 r Z 0 ζ ω 1 λ 0 r 2 2 π c 0 ( ζ n n eff a + ζ Z Z 0 a )
cot ( n eff ω 0 r ε 0 μ 0 ( d + δ m ) ) = 0
λ 0 r n c = λ 0 r n eff n eff n c
λ 0 r a = λ 0 r n eff n eff a
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.