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

Cavity resonances of metal-dielectric-metal nanoantennas

Open Access Open Access

Abstract

We propose a new design of optical nanoantennas and numerically study their optical properties. The nanoantennas are composed of two cylindrical metal nanorods stacked vertically with a circular dielectric disk spacer. Simulation results show that when the dielectric disk is less than 5nm in thickness, such nanoantennas exhibit two types of resonances: one corresponding to antenna resonance, the other corresponding to cavity resonances. The antenna resonance generates a peak in scattering spectra, while the cavity resonances lead to multiple dips in the scattering spectra. The cavity resonant frequency can be tuned by varying the size of the dielectric disk. The local field enhancement inside the cavity is maximized when the diameter of the dielectric disk is roughly half that of the rod and when the cavity and antenna resonant frequencies coincide with each other. This new nanoantenna promises applications in single molecule surface enhanced Raman spectroscopy (SERS) owing to its high local field enhancements and large scale manufacturability.

©2008 Optical Society of America

1. Introduction

Focusing and confining electromagnetic waves into nanoscale dimensions by employing unique properties of surface plasmon polaritons are of great fundamental and practical interest [1–7]. As a result of nanofocusing/nanoconfinements, local electromagnetic field can be enhanced by as high as 3 orders of magnitude leading to a number of extraordinary effects on the materials and molecules at the focus. For example, nonlinear optical effects [3, 8, 9], fluorescence emission [10, 11], Raman signals [12, 13], and optical transmission through subwavelength holes can be enhanced significantly. Especially, recent surface enhanced Raman scattering (SERS) experiments show that the enhancements can reach up to 14 orders of magnitude, that not only makes SERS an attractive single molecule technique with promising applications but also posts fundamental questions regarding its physical mechanisms[5, 14–16].

Early experimental systems where single molecule SERS was observed were based on metallic nanoparticle aggregates [5, 14, 15]. In such colloidal systems, hotspots were normally found by chance from a large ensemble of samples which make the experiments not controllable. The theoretical illustration of hotspots with nanoparticle dimmers [4, 17] has stimulated new nanoantenna designs [9, 18–21], and the state-of-the-art techniques have been applied to fabricate these dimmers with nanogaps [22–24]. However, it is still challenging to build these nanoantenna dimmers with gaps below 5nm in a reproducible fashion and in large scale. In order to decipher the physical mechanisms of single molecule SERS and for practical industrial applications, a plasmonic system allowing for reproducible fabrication and well controlled EM enhancements is highly desired. Detailed characterization of the EM enhancements in such a model system will facilitate detailed comparison and validation of various theories. Recently, metal nanodisks separated with dielectric spacers have been proposed and demonstrated as tunable plasmonic resonators for SERS applications [25–27]. One great advantage of these metal-dielectric-metal nanodisks is that the dielectric gap may be controlled with sub-nm accuracy by using advanced deposition techniques. One issue of such plasmonic nanodisks is that Raman molecules can only be attached to the metal disk circumferences where the electrical field enhancement is still low in comparison to the field enhancement within the gap of nanoparticle dimmers.

In this paper, we propose a new design of optical nanoantennas and study their optical properties by numerical simulations. The nanoantennas are composed of two metal cylinders stacked vertically with a circular dielectric disk spacer. The diameter of the dielectric disk is designed smaller than that of the metal cylinder to accommodate SERS molecules. We show that when the dielectric disk is thinner than 5nm, such nanoantennas exhibit two types of resonances: one corresponding to antenna resonance, the other corresponding to cavity resonances. The antenna resonance generates a peak in scattering spectra, while the cavity resonances lead to multiple dips in the scattering spectra. The cavity resonances cause giant electrical field enhancements inside the nanocavity. The enhancement of electrical field inside the nanocavity is maximized when the diameter of the dielectric disk is roughly half that of the rod and when the cavity and antenna resonant frequencies coincide with each other. An advantage of such nanoantenna design is that the gap size can be accurately controlled below 1nm by using atomic layer deposition, making it a good candidate for single molecule SERS applications.

2. Nanoantenna design and simulation methods

A schematic of the nanoantenna is shown in Fig. 1. The diameter of the dielectric disk is designed to be smaller than that of the metal cylinders to accommodate SERS molecules. Ag is chosen as an exemplary material for the metal cylinders to achieve wide tuning range of the antenna resonant frequency. SiO2 is chosen for the dielectric disk layer because atomic layer deposition process for SiO2 is readily accessible with sub-1nm accuracy in film thickness control [28, 29]. The optical properties of this new type of nanoantennas are studied by employing the commercial software package CST Microwave Studio (MWS). The software is based on finite-difference time domain (FDTD) method, and it has been shown that the simulation results of surface plasmon waves with this software agree well with analytical solutions [30].

 figure: Fig. 1.

Fig. 1. Schematic side view (left) and top view (right) of the metal-dielectric-metal nanoantenna. The wave vector k of the incident light is indicated by the arrow; the polarization is along the nanoparticle axis. The blue spot indicates the probe position where the local field spectra are calculated.

Download Full Size | PDF

The Drude model was utilized to describe the frequency dependence of silver permittivity: εAg(ω)=ε2 p/ω(ω-iδ), where these parameters used for simulations, ε=3.57, ωp=1.388×1016 rad/s, and δ=1.064×1014 Hz, are obtained by fitting the Drude model to the measured permittivity of bulk Ag material [31]. Bulk dielectric constant for SiO2, i.e. ε=2.13 is used for the SiO2 disks for all frequencies of simulations.

The nanoantenna is excited by a plane light wave with the electrical field along the nanoantenna axis and the wave vector perpendicular to the nanoantenna axis (Fig. 1). Four geometric parameters are needed for characterizing the nanoantenna: diameter and length of the metal cylinders, diameter and thickness of the dielectric disk. In our simulations, the diameter and length of the metal cylinders are fixed at 60nm and 80nm respectively. All local field spectra shown in this paper are for the electrical field at the probe position indicated in Fig. 1, and have been normalized by the strength of incident field to represent the local field enhancement. The far-field intensities are calculated for a position about 100 wavelengths away from the nanoparticle, and are rescaled to have unity maximum intensity for the sake of better comparison.

3. Cavity and Antenna Resonances

Local field and far field spectra for different SiO2 disk diameters and thicknesses are shown in Fig. 2; interesting resonance features can be observed. In the local field spectra, the resonant wavelength and the peak height vary considerably with the SiO2 disk sizes. When the SiO2 diameter is equal to that of the metal cylinder, no major resonant peaks exist with field enhancements larger than 50 even for 1nm thickness of SiO2 [Fig. 2(d)]. Reduction of the SiO2 disk diameter leads to dramatic changes in the local field spectra. When the SiO2 disk thickness is less than 5nm, the dominant resonance appears as a narrow peak. The bandwidth at the half field strength of this resonance is between 40nm to 100nm and increases with the SiO2 disk diameter [see Figs. 2(a)–2(c)]. When the gap is larger than 5nm the narrow peak shifts to a higher frequency. Meanwhile, the broad peak, which is less prominent at small gaps, becomes dominant.

 figure: Fig. 2.

Fig. 2. Local electrical field spectra (a–d) and far field scattering spectra (e–h) for the SiO2 disk diameters at 20nm (a, e), 30nm (b, f), 40nm (c, g) and 60nm (d, h).

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. (a). Simulated (symbols) and calculated (lines) cavity resonant wavelength as a function of the SiO2 disk thickness for different SiO2 disk diameters. (b) The electrical enhancements at the probe position as a function of the SiO2 disk thickness at different SiO2 disk diameters. (c) The local E-field enhancements as a function of SiO2 disk diameters for 1nm disk thickness. The solid lines are the guidance to the eye.

Download Full Size | PDF

The broad peaks observed at large gaps are fundamentally different from the narrow resonant peaks at small gaps. Close to the wavelengths of these broad peaks, far-field scattering spectra show resonant peaks as well. In contrast, at the wavelengths of these narrow peaks, the far-field spectra show dips. This indicates that the narrow peaks in local field spectra are resulted from cavity resonances, while the broad peaks are resulted from the plasmon resonances of the whole antenna.

Regardless these dips caused by the cavity resonances, the far-field scattering spectra consist of only single peaks [Figs. 2(e)–2(h)]. These far field resonances, denoted as antenna resonances, are similar to those observed previously in various nanoantenna dimmers [19].

It is important to note that the cavity and nanoantenna resonance frequencies are not always coincident with each other and can be widely tuned by varying the SiO2 disk sizes. Both the cavity and antenna resonant wavelengths decrease appreciably with the increase of the SiO2 disk thickness, while increase only slightly with the increase of SiO2 disk diameter [Fig. 3(a)]. This indicates that the accurate control of the nanogaps between two nanorods is critical to controlling cavity resonances and the local field enhancements, while the control of the SiO2 disk diameter is insignificant.

The local field enhancement at the dominant cavity mode frequencies increases when the SiO2 disk thickness decreases. Especially a sharp rise can be seen when the thickness is less than 5nm [Fig. 3(b)]. Meanwhile, the local field enhancements are affected by the diameter of the SiO2 disk. With the increase of SiO2 disk diameter, the local field enhancement goes up first and then declines. It can be found that the local field enhancement reaches its maximum when the SiO2 disk diameter is approximately half the diameter of the metal cylinder [Fig. 3(c)].

For these cavity resonance modes, we calculated the electrical field distributions at the plane through the middle of the SiO2 disk and the results for the SiO2 disk diameter at 30nm are presented in Fig. 4. It can be seen that when the SiO2 thickness is less than 5nm, the field distribution pattern for the major cavity mode does not vary with SiO2 thickness, while the field distribution patterns of other modes show significant change when the SiO2 thickness is changed. In addition, the electrical field distribution for the broad peak is very different from that of the major cavity mode at small gap, which can also be seen from the snapshots of local electrical field vectors. For the narrow band cavity mode, the electrical field direction inside the SiO2 disk is opposite to the field direction outside the disk (Fig. 4). While for the broadband mode, the electrical fields inside the gap between two Ag cylinders are parallel to the direction of incident field [Fig. 4(c)].

 figure: Fig. 4.

Fig. 4. (a). Local electrical field amplitude distributions in the plane through the middle of the SiO2 disk at three primary cavity resonant modes for SiO2 disk thickness at 1nm (first row), 2nm (second row), 5nm (third row) and 10nm (fourth row). The wavelengths indicated in each picture are the wavelengths of excitation. (b) and (c) Snapshots of local electrical field vectors for the major cavity mode for the SiO2 thickness at 2nm (b) and 5nm (c) respectively. For the sake of clarity, logarithmic color scales are used. Red color represents the maximal field strength of individual pictures, and thus does not indicate the same field strength for pictures of different modes.

Download Full Size | PDF

In order to understand these cavity resonances, it is necessary to recall the surface plasmon modes present in the dielectric gap of two metal films. For an interface between semi-infinite dielectric and metal materials, there exists only one surface plasmon mode whose wave vector is given by ksp=(ω/c)[εmεd/(εm+εd)]1/2 Here εm and εd are the permittivity of the metallic and dielectric materials, ω and c are the radial frequency and speed of light. When two metal-dielectric interfaces are brought to close proximity, the surface plasmons at two interfaces couple with each other, which results in separation of the degenerate mode into a symmetric mode and an anti-symmetric mode. The anti-symmetric mode only exists at high frequency ranges, while the symmetric mode can sustain at all frequencies for very small gaps. For the small dielectric thicknesses considered here, the anti-symmetric mode is out of range, and will not be considered. The dispersion relationship for the symmetric mode can be expressed as [32, 33]:

tanh(kdt2)=εdkmεmkd

where t is dielectric gap thickness, km=(k 2 sp-εm ω 2/c 2)1/2, kd=k 2 sp-εdω 2/c 2)1/2, and ksp is the wave vector for the gap surface plasmon. In addition, the permittivity of the dielectric disk affects the local field distribution. The electrical field decays with the distance from the metal-dielectric interface. For the field inside the dielectric gap, the decay length, also called the skin depth [32], can be given n as δd=1Re[(ksp2εdω2c2)12] . Given that an increase of εd leads to an increase of ksp, the skin depth δd decreases with the increase of the permittivity of the dielectric layer. When δd is larger than the dielectric thickness, the electrical field inside the gap will be homogeneous in the direction perpendicular to the gap plane.

In fact, the major cavity resonance observed above corresponds to formation of a standing wave of surface plasmons, as the symmetric distribution of electrical field shown in Fig. 4(b) is in agreement with the symmetric surface plasmon wave. Considering that the electrical field is maximal at the edge of the gap, the resonant condition can be written as:

kspd+ksp(Dd)=2π

where D and d are the diameter of the metal cylinder and dielectric disk respectively, ksp and ksp are the surface plasmon wave vectors for SiO2 and air portions of the cavity respectively. Simple numerical calculations yield that the cavity resonant wavelengths obtained from this condition agree well with the simulation results [Fig. 3(a)]. Some discrepancies at 10nm thickness of SiO2 are due to the fact that the simulated resonant wavelengths are assigned to these broad peaks. At this large gap size (10nm), the wave vectors for the symmetrical plasmon wave is decreased, thus Eq. (2) leads to the cavity resonant wavelength being much smaller than that of antenna resonances [small peaks for 10nm gap size in Figs. 2(a)–2(d)]. Moreover, at the antenna resonant frequency, kspd+ksp (D-d)≪π, so the electrical field inside the gap for the broad resonant peak [Fig. 4(c)] is always in the same direction.

Overall, the thickness and diameter of the dielectric disk determine the resonant wavelengths of the major cavity mode, providing a way of tuning the cavity resonance wavelength. There exist other cavity modes [Fig. 4(a)], which correspond to more complicated patterns of surface plasmon standing waves. From simulations, these high order cavity modes do not affect far field scattering much, as no dips can be seen in the far field spectra [Figs. 2(e)–2(h)].

These seemingly complicated cavity and antenna resonance behaviors can be qualitatively understood in the concept of optical circuit elements [34, 35]. It should be noted that the optical circuit element concept has been developed for very small plasmonic particles, extending this concept to large particles like ours implies that phase retardation has been omitted. Following the same analogy as the optical circuit elements, the metal cylinders can be considered as inductors (L1), while the nanocavity can be considered as two capacitors (C1 and C2) in parallel connections representing the cavity regions of SiO2 and air respectively. Cfringe represents the capacitance between the metal cylinders outside the cavity region. It is also important to note that the outside capacitor ring C2 and the inside capacitor C1 of SiO2 disk are connected through the Ag cylinder. Therefore, to take this into account, two inductors, L2 and L3, need to be added (Fig. 5). For simplicity, losses are ignored here. Such a circuit network contains two LC circuits and thus exhibit two resonant frequencies. L2, L3, and C1, C2 form a small LC circuit whose resonance corresponds to the cavity resonance. While all elements together form a big LC circuit whose resonance corresponds to the antenna resonance.

 figure: Fig. 5.

Fig. 5. Optical circuit element representation of the optical nanoantenna.

Download Full Size | PDF

At the cavity resonance, the current circulating inside the small loop should lead to opposite phases for C1 and C2, which explains the field directions for the cavity mode [Fig. 4(b)]. When the SiO2 thickness is increased or the capacitances of C1 and C2 are decreased, the resonant frequency of the small LC circuit increases. Especially, when the frequency of the input signals are close to the resonant frequency of the whole circuit while lower than that of the small LC circuit, the L2, L3, C1, C2 loop acts as more capacitive, and C1 and C2 should have the same phase. This agrees well with the above simulation results.

In addition to the losses of metallic materials, the local field enhancements are affected by two design factors: radiation loss and the cavity/antenna coupling. When the gap size is reduced, it becomes difficult for the electrical fields to leak out of the cavity, or radiation loss is reduced for the cavity modes. When the resonant frequencies of cavity mode and antenna resonance coincide, the whole structure will be optimized to receive energy at the cavity mode frequency, and should leads to optimized local field enhancements. We simulated nanoantennas with increased Ag cylinder lengths. When the Ag cylinder is 120nm in length, the resonant frequency of the major cavity mode coincides well with the antenna resonance frequency at 384THz. The results show that the local field enhancements can increase by ~50% in comparison with these nanoantennas of 80nm length shown above. Therefore, by tuning the antenna resonance frequency and cavity resonance frequency to be coincident, the local field enhancements can be maximized.

It is worth pointing out that the major contributions to the large local field enhancements originate from the cavity resonances. Even when the cavity and antenna resonance frequencies do not match each other, reasonably high local field enhancements can be obtained. This may be relevant to previous experimental observations that the plasmon resonance wavelengths of hotspots do not correlated well with the laser wavelengths [36]. This also indicates that matching cavity resonant wavelengths with the excitation laser wavelength are critical to achieving large SERS signals.

4. Summary

In summary, we have proposed and numerically studied a new design of plasmonic nanoantennas which are composed of two metal cylinders stacked with a dielectric spacer. It is shown that by making the diameter of the dielectric disk smaller than that of the metal cylinder, cavity resonances can be generated inside the gap. This cavity resonance is different from the plasmon resonances of the whole antenna, and generates dips in stead of peaks in far field scattering spectra. Moreover, this cavity resonance is more critical to the local field enhancements within the nanocavity. Experimental testing of SERS using such nanoantennas may be helpful understand the potential roles of cavity resonances in SERS enhancements.

Acknowledgments

This work is supported by the Ohio Board of Regents Research Challenge, and a start-up fund from Kent State University. The authors like to acknowledge Cheng Sun, Nicholas X. Fang and Xiang Zhang for valuable discussions.

References and links

1. M. I. Stockman, “Nanofocusing of optical energy in tapered plasmonic waveguide,” Phys. Rev. Lett. 93, 137404 (2004). [CrossRef]   [PubMed]  

2. D. K. Gramotnev and K. C. Vernon, “Adiabatic nano-focusing of plasmons by sharp metallic wedges,” Appl. Phys. B-Lasers and Optics 86, 7–17 (2007). [CrossRef]  

3. E. Verhagen, A. Polman, and L. Kuipers, “Nanofocusing in laterally tapered plasmonic waveguides,” Opt. Express 16, 45–57 (2008). [CrossRef]   [PubMed]  

4. H. X. Xu, J. Aizpurua, M. Kall, and P. Apell, Electromagnetic contributions to single-molecule sensitivity in surface-enhanced Raman scattering. Phy. Rev. E62, 4318–4324 (2000). [CrossRef]  

5. H. X. Xu, E. J. Bjerneld, M. Kall, and L. Borjesson, “Spectroscopy of single hemoglobin molecules by surface enhanced Raman scattering”. Phys. Rev. Lett. 83, 4357–4360 (1999). [CrossRef]  

6. L. L. Yin, V. K. Vlasko-Vlasov, J. Pearson, J. M. Hiller, J. Hua, U. Welp, D. E. Brown, and C. W. Kimball, “Subwavelength focusing and guiding of surface plasmons,” Nano Lett. 5, 1399–1402 (2005). [CrossRef]   [PubMed]  

7. H. T. Miyazaki and Y. Kurokawa, “Squeezing visible light waves into a 3-nm-thick and 55-nm-long plasmon cavity,” Phys. Rev. Lett. 96, 097401 (2006). [CrossRef]   [PubMed]  

8. V. M. Shalaev, Nonlinear Optics of Random Media: Fractal Composites and Metal-Dielectric Films, (Springer, Berlin Heidelberg, 2000).

9. P. J. Schuck, D. P. Fromm, A. Sundaramurthy, G. S. Kino, and W. E. Moerner, “Improving the mismatch between light and nanoscale objects with gold bowtie nanoantennas,” Phys. Rev. Lett. 94, 017402 (2005). [CrossRef]   [PubMed]  

10. J. R. Lakowicz, “Radiative decay engineering 5: metal-enhanced fluorescence and plasmon emission,” Anal. Biochem. 337, 171–194 (2005). [CrossRef]   [PubMed]  

11. E. Fort and S. Gresillon, “Surface enhanced fluorescence,” J. Phys. D-Appl. Phys. 41, 013001 (2008). [CrossRef]  

12. M. Fleischman, P. J. Hendra, and A. J. McQuillan, Chem. Phys. Lett.26, 123 (1974). [CrossRef]  

13. M. Moskovits, “Surface Enhanced Raman Spectroscopy,” Rev. of Mod. Phys. 57, 783–826 (1985). [CrossRef]  

14. K. Kneipp, Y. Wang, H. Kneipp, L. T. Perelman, I. Itzkan, R. Dasari, and M. S. Feld, “Single molecule detection using surface-enhanced Raman scattering (SERS),” Phys. Rev. Lett. 78, 1667–1670 (1997). [CrossRef]  

15. S. M. Nie and S. R. Emery, “Probing single molecules and single nanoparticles by surface-enhanced Raman scattering,” Science 275, 1102–1106 (1997). [CrossRef]   [PubMed]  

16. J. Jiang, K. Bosnick, M. Maillard, and L. Brus, “Single molecule Raman spectroscopy at the junctions of large Ag nanocrystals,” J. Phys. Chem. B 107, 9964–9972 (2003). [CrossRef]  

17. E. Hao and G. C. Schatz, “Electromagnetic fields around silver nanoparticles and dimmers,” J. Chem. Phys. 120, 357–366 (2004). [CrossRef]   [PubMed]  

18. A. Sundaramurthy, K. B. Crozier, G. S. Kino, D. P. Fromm, P. J. Schuck, and W. E. Moerner, “Field enhancement and gap-dependent resonance in a system of two opposing tip-to-tip Au nanotriangles,” Phys. Rev. B 72, 165409 (2005). [CrossRef]  

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

20. E. Cubukcu, E. A. Kort, K. B. Crozier, and F. Capasso, “Plasmonic laser antenna,” Appl. Phys. Lett. 89, 093102 (2006). [CrossRef]  

21. Y. Alaverdyan, B. Sepulveda, L. Eurenius, E. Olsson, and M. Kall, “Optical antennas based on coupled nanoholes in thin metal films,” Nat. Phys. 3, 884–889 (2007). [CrossRef]  

22. K. H. Su, Q. H. Wei, X. Zhang, J. J. Mock, D. R. Smith, and S. Schultz, “Interparticle coupling effects on plasmon resonances of nanogold particles,” Nano Lett. 3, 1087–1090 (2003). [CrossRef]  

23. T. Atay, J. H. Song, and A. V. Nurmikko, “Strongly interacting plasmon nanoparticle pairs: From dipole-dipole interaction to conductively coupled regime,” Nano Lett. 4, 1627–1631 (2004). [CrossRef]  

24. D. P. Fromm, A. Sundaramurthy, P. J. Schuck, G. Kino, and W. E. Moerner, “Gap-dependent optical coupling of single “Bowtie” nanoantennas resonant in the visible,” Nano Lett. 4, 957–961 (2004).

25. K. H. Su, Q. H. Wei, and X. Zhang, “Tunable and augmented plasmon resonances of Au/SiO2/Au nanodisks,” Appl. Phys. Lett. 88, 063118 (2006). [CrossRef]  

26. L. D. Qin, S. L. Zou, C. Xue, A. Atkinson, G. C. Schatz, and C. A. Mirkin, “Designing, fabricating, and imaging Raman hot spots”. Proc. of the Nat. Acad. Sci. USA 103, 13300–13303 (2006). [CrossRef]  

27. 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, 3964–3968 (2006). [CrossRef]   [PubMed]  

28. W. Gasser, Y. Uchida, and M. Matsumura, “Quasi-monolayer deposition of silicon dioxide,” Thin Solid Films 250, 213–218 (1994). [CrossRef]  

29. J. W. Klaus, O. Sneh, A. W. Ott, and S. M. George, “Atomic layer deposition of SiO2 using catalyzed and uncatalyzed self-limiting surface reactions,” Surf. Rev. Lett. 6, 435–448 (1999). [CrossRef]  

30. Z. W. Liu, Q. H. Wei, and X. Zhang, “Surface plasmon interference nanolithography,” Nano Lett. 5, 957–961 (2005). [CrossRef]   [PubMed]  

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

32. Y. Kurokawa and H. T. Miyazaki, “Metal-insulator-metal plasmon nanocavities: Analysis of optical properties,” Phys. Rev. B. 75, 035411 (2007). [CrossRef]  

33. S. I. Bozhevolnyi and T. Søndergaard, “General properties of slow-plasmon resonant nanostructures: nanoantennas and resonators,” Opt. Express 15, 10869–10877 (2007). [CrossRef]   [PubMed]  

34. N. Engheta, A. Salandrino, and A. Alu, “Circuit elements at optical frequencies: Nanoinductors, nanocapacitors, and nanoresistors,” Phys. Rev. Lett. 95, 095504 (2005). [CrossRef]   [PubMed]  

35. N. Engheta, “Circuits with light at nanoscales: Optical nanocircuits inspired by metamaterials,” Science 317, 1698–1702 (2007). [CrossRef]   [PubMed]  

36. E. C., Le Ru, C. Galloway, and P. G. Etchegoin, “On the connection between optical absorption/extinction and SERS enhancements,” Phys. Chem. Chem. Phys. 8, 3083–3087 (2006).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Schematic side view (left) and top view (right) of the metal-dielectric-metal nanoantenna. The wave vector k of the incident light is indicated by the arrow; the polarization is along the nanoparticle axis. The blue spot indicates the probe position where the local field spectra are calculated.
Fig. 2.
Fig. 2. Local electrical field spectra (a–d) and far field scattering spectra (e–h) for the SiO2 disk diameters at 20nm (a, e), 30nm (b, f), 40nm (c, g) and 60nm (d, h).
Fig. 3.
Fig. 3. (a). Simulated (symbols) and calculated (lines) cavity resonant wavelength as a function of the SiO2 disk thickness for different SiO2 disk diameters. (b) The electrical enhancements at the probe position as a function of the SiO2 disk thickness at different SiO2 disk diameters. (c) The local E-field enhancements as a function of SiO2 disk diameters for 1nm disk thickness. The solid lines are the guidance to the eye.
Fig. 4.
Fig. 4. (a). Local electrical field amplitude distributions in the plane through the middle of the SiO2 disk at three primary cavity resonant modes for SiO2 disk thickness at 1nm (first row), 2nm (second row), 5nm (third row) and 10nm (fourth row). The wavelengths indicated in each picture are the wavelengths of excitation. (b) and (c) Snapshots of local electrical field vectors for the major cavity mode for the SiO2 thickness at 2nm (b) and 5nm (c) respectively. For the sake of clarity, logarithmic color scales are used. Red color represents the maximal field strength of individual pictures, and thus does not indicate the same field strength for pictures of different modes.
Fig. 5.
Fig. 5. Optical circuit element representation of the optical nanoantenna.

Equations (2)

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

tanh ( k d t 2 ) = ε d k m ε m k d
k sp d + k sp ( D d ) = 2 π
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.