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

Plasmon hybridization for enhanced nonlinear optical response

Open Access Open Access

Abstract

We report the plasmon hybridization between silver nanoprisms and a thin gold film as a means to tune the plasmon resonance and to achieve enhanced optical second harmonic generation. The hybridization enhances the second harmonic counts by nearly three orders of magnitude when varying the spacer layer between the nanoprisms and the gold film. Finite-difference time-domain calculations agree within a factor of 2 with the experimental findings in terms of the predicted enhancement factor. This plasmon hybridization approach is promising for future applications, including multi-photon lithography and nonlinear sensing using metal nanoparticles.

©2012 Optical Society of America

1. Introduction

Plasmonics allows for enhanced local electromagnetic fields using metal nanostructures. These enhanced local fields are naturally appealing for nonlinear optical processes. Many works have studied the nonlinear optical response of metal nanostructures fabricated both by top-down [18] and bottom-up [912] approaches. Of particular interest is the nonlinear optical response of metal nanoparticles [13,14]. For example, these may be incorporated in other media for multi-photon lithography [15]. Symmetry is an important feature in second harmonic generation (SHG), and the structural properties of metal nanoparticles are commonly altered to achieve broken symmetry with greater SHG [16] and the related property of directional enhancement of the nonlinear response [17].

Past works have studied the nonlinear optical response of silver nanoparticles, even mapping down to the single nanoparticle level [18]. Silver nanoprisms are particularly interesting because of their single-crystal structure, the low loss of silver, the sharp tips of the nanoprism, and their lack of inversion symmetry. Silver nanoprisms of less than 100 nm in size, however, do not have a plasmonic resonance at the near-infrared wavelengths of a Ti:Sapphire laser, the most common source for ultra-fast measurements. To achieve the resonance at these wavelengths requires some tuning mechanism [1924]. This can be achieved by top-down fabrication of multi-resonant optical antenna structures. For silver nanoprisms, and other particles fabricated by bottom-up methods, we propose the plasmon hybridization approach to tune the resonance to that of the laser source.

Plasmon hybridization refers to coupling between metal nanoparticles [2527], or nanoparticles to other metal nanostructures (such as a metal film) [2835], to tune the optical response. For example, it has been shown that the plasmonic resonance of silver nanoparticles can be tuned by various amounts by spacing them off from a gold film with a spacer of various thicknesses [36]. The spacer layer thickness can also be tuned a posteriori by voltage controlled oxidation [37]. Here we are particularly interested in the hybridization between a metal nanoparticle and a thin metal film that supports short-range modes and gives precise tuning of the lowest order resonance [31]. A thin metal film is advantageous because it can transmit light; e.g., in applications where the film is deposited on top of a photoresist layer.

In this work, we use plasmon hybridization between colloidally synthesized silver nanoprisms and a 10 nm thick gold film to tune the plasmonic resonance around the peak wavelength of the fundamental laser source. At the peak wavelength, we obtain three orders of magnitude enhanced second harmonic generation, as compared to the far off-resonance condition of a large spacer layer, or having no metal film at all.

2. Silver nanoprisms in colloidal solution

2.1 Synthesis of silver nanoprisms

The silver nanoprisms were synthesized by the light-assisted conversion method [38,39] of silver nanospheres to nanoprisms. As a typical synthesis process, aqueous solution of AgNO3 (0.1 mM, 100 ml), (204390, Aldrich Chemicals), and trisodium citrate (0.3 mM), (S2990, ACP Chemicals Inc.), was prepared in presence of air with a moderate stirring rate (120 rpm). Next, NaBH4 solution (50 mM, 100 ml), (7420-1, Caledon Laboratories Ltd.), was injected to the system. Following this, Bis (p-sulfonatonphenyl) phenylphosphine dehydrate dipotassium salt (BSPP) (5 mM, 2 ml), (698539, Aldrich Chemicals), was dropped into the solution over 2 min. The system was irradiated with a 24 W halogen lamp (ser. 700, Sunnex Inc.) for 72 h, while the optical extinction of the samples from the solution was monitored as a function of time with a Cary 5 UV-VIS-NIR spectrometer.

2.2 Extinction evaluation of silver nanoprisms in aqueous solution

Figure 1(a) shows that after 17 h irradiation, the initial silver nanospheres with 400 nm plasmon resonance were converted to nanoparticles with plasmon resonance at 600 nm. Continuing irradiation for longer intervals, the extinction resonance shifted to longer wavelength, while the reaction was terminated after 72 h irradiation with final nanoprisms with 87 ± 13 nm edge length (uncertainty is standard deviation), 6 nm radius of curvature at the edges, and 12 nm thickness, and the plasmon resonance at 680 nm. The inset of Fig. 1(a) shows a scanning electron microscope (SEM) image of a typical silver nanoprism after 72 h irradiation. In this work, all hybrid structures and samples were prepared using the same solution of silver nanoprisms with a maximum extinction at 680 nm. Figures 1(b,c) show an atomic force microscope scan (Agilent 5500) in AC mode of the silver nanoprisms dispersed on a silicon substrate. The height was measured to be 12 nm (see Fig. 1(d)).

 figure: Fig. 1

Fig. 1 (a) Extinction of silver nanoparticles in aqueous solution as a function of irradiation time. Inset: Scanning electron microscope image of a typical silver nanoprism after 72 h irradiation. (b,c) Atomic force microscope (AFM) characterization of silver nanoprisms spin coated on a silicon substrate. (d) The height of a typical silver nanoprism was measured to be 12 nm with the AFM.

Download Full Size | PDF

3. Fabrication of hybrid systems

Figure 2(a) shows a schematic of the hybridized system containing the silver nanoprisms, spacer layer, thin gold film and glass substrate. Commercial substrates were used consisting of a 10 nm thick gold film with a 2 nm Ti adhesion layer on glass (AU.0100.CSS, Platypus Technologies). The commercial substrate had a flat and continuous film, even for 10 nm thickness, and we could achieve similar quality only by evaporating Au films with the substrate temperature elevated to 200C or more; however, the remainder of this work is based entirely on samples prepared with the commercial substrates. The spacer layer was created by spin coating poly methyl methacrylate (950 PMMA A2, MicroChem) at 3500 rpm for 90 sec with varying anisole (296295, Sigma-Aldrich) solvent concentration to achieve varying post-baking thicknesses between 5 nm and 20 nm [40]. The PMMA on substrate was baked for 5 min at 180C. Following this, the silver nanoprisms were spin coated onto the substrate at 700 rpm for 90 sec (1 drop of the undiluted solution). The surface coverage density was 120 nanoprisms per 60 µm2, assuring a good statistical average in the SHG measurements below.

 figure: Fig. 2

Fig. 2 (a) Schematic of silver nanoprisms with PMMA spacer layer over a 10 nm thick Au film adhered to a glass substrate with a 2 nm Ti layer. (b) Schematic of scattering measurement setup. WLS = white light source, obj = microscope objective lens. (c) Scattering measurement for three hybrid structures with different PMMA spacer layer thicknesses (shown in legend). Green dotted line: Ti:Sapphire spectrum. (d) Scattering simulation results for silver nanoprisms for the corresponding spacer thicknesses.

Download Full Size | PDF

4. Scattering of hybrid system

4.1 Scattering experimental setup

Figure 2(b) shows a schematic of the scattering measurement setup. Halogen white light (LS-1-LL, Ocean Optics Inc.) was focused into the glass side of the sample by a 20 × microscope objective (Mitutoyo Plan Apo, Edmund Optics Inc.), allowing for waveguiding in the glass substrate. The scattered light at the surface of the sample, where the nanoprisms and metal film were located, was collected normal to the surface by an optical fiber (0.22 NA) and the spectrum was recorded using a spectrometer (QE65000, Ocean Optics Inc.).

This measurement setup allowed for recording the scattered light from the nanoprisms. The setup was aligned to ensure that minimal light is coupled into the optical fiber when there are no nanoprisms. For a glass slide only, there is negligible scattering for the well-aligned setup; however, there is still some broadband scattering observed due to the roughness of the Au film for the case of the gold-over-glass structure.

4.2 Scattering FDTD simulation

The proposed structure was simulated using a commercial finite-difference time-domain simulation package (FDTD ver. 7.5.7, Lumerical Solutions Inc.) to estimate the scattered power and local field enhancement. For accurate modeling of the structure, a mesh size of 1.5 nm was used. The simulation domain was terminated with a perfectly matched layer for minimal reflection. The complex permittivities of gold and silver were modeled using the experimental data of Johnson and Christy [41] and Palik [42], respectively. For the PMMA spacer layer, a refractive index of 1.5 was used. The silver nanoprisms were modeled with a rounded edge radius of 6 nm. The source was polarized along the axis of symmetry of the silver nanoprism in the x direction (see Figs. 5(a) and (c)). To calculate the scattered power, we employed the formalism of total field scattered field. A set of two-dimensional power monitors was used, enclosing the nanoprism. The total power exiting this closed surface was the scattered power. Also, we used a two-dimensional frequency domain field profile monitor to determine the near-field intensity distribution.

4.3 Scattering measurements

Figure 2(c) shows scattering experimental measurements for three different spacer layer thicknesses. The Ti:Sapphire laser spectrum is shown with a dotted line green line, to compare the peak position of light source and the hybridized resonances. Figure 2(d) shows the FDTD simulation results for silver nanoprisms for the corresponding spacer thicknesses. It is noted that the linewidth of the scattering spectra for the experimental spectra in Fig. 2(c) is broader than that of the numerical simulations in Fig. 2(d). This is typical due to differences in the nanoprism size variations (of the order of 10 nm), edge sharpness, and surface roughness of the Au film.

Figure 3 shows the experimental and simulation scattering peak wavelength for the hybrid plasmonic system (Ag-PMMA-Au-Ti-glass) as a function of PMMA thickness for additional thicknesses. Good agreement is seen between the measured scattering peak, and the one obtained by FDTD calculations. It is clear from Fig. 3 and Figs. 2(c,d) that we can tune the plasmonic resonance through the peak wavelength of the Ti:Sapphire laser (808 nm) by plasmon hybridization. Since the linear response gives the largest scattering at the peak wavelength of the Ti:Sapphire laser for a PMMA thickness of 10 nm, we expect that the local field enhancement, and hence the SHG, will be greatest for that thickness too.

 figure: Fig. 3

Fig. 3 Peak scattering wavelength as a function of PMMA spacer layer thickness from experiment and FDTD simulations.

Download Full Size | PDF

5. Second harmonic generation measurements

5.1 SHG experimental setup

Figure 4(a) shows a schematic of the SHG detection setup. A Ti:sapphire laser was used to produce ~25 fs pulses with a center wavelength of 808 nm and a repetition rate of 90 MHz. The applied average power was tuned via neutral density (ND) filters, while the maximum average power was limited to 30 mW. The laser light was focused onto the sample with off-axis illumination using a 20 × long working distance microscope objective (Mitutoyo Plan Apo, Edmund Optics Inc.) to produce a 60 µm2 spot area. The average power was measured at the same location as the sample. The reflected laser light was blocked with an off-axis metal stop, and the normal emission beam, containing the SHG signal and scattered fundamental beam, was guided to the streak camera (C5680, Hamamatsu) with a 45° cold mirror (FM04, Thorlabs Inc.), passing through a bandpass BG40 filter to remove the fundamental beam with the expense of 30% of second harmonic signal. The streak camera was operated in synchroscan (M5675, Hamamatsu) photon-counting mode with a microchannel plate gain of 6, in a manner similar to our previous work [1,3,5].

 figure: Fig. 4

Fig. 4 (a) Schematic of SHG measurement setup. ND = neutral density filter, obj = microscope objective lens, BG40 = blue-green band pass filter. (b) SHG measured using spectrum analyzer in front of streak camera (not shown in (a)). (c) SHG count dependence on PMMA spacer layer thicknesses. (d) Log-log plot of SHG count vs. incident laser power with linear-fit slope of 2.17 ± 0.20. For SHG, a slope of 2 is expected.

Download Full Size | PDF

5.2 SHG measurements

Figure 4(b) shows the measured second harmonic spectrum, taken from in front of the streak camera using the fiber-probe spectrum analyzer. Figure 4(c) shows the SHG counts as a function of PMMA thickness for 30 mW incident power. The maximum value obtained was 9000 ± 240 (error from standard deviations over 10 runs at different locations on the sample) for a 10 nm PMMA thickness. By comparison, for a sample of silver nanoprisms on a glass microslide alone, the SHG was 14 ± 4. This shows that the enhancement factor of the SHG was approximately three orders of magnitude. Figure 4(d) shows a log-log plot of the SHG versus incident laser power with a slope of 2.17 ± 0.20 for the 10 nm thickness sample. Considering that a slope of 2 is expected for SHG, this is a reasonable result. The PMMA does not show any SHG alone and is expected to give negligible contribution to the SHG signal as compared to the nanoprisms. Furthermore, negligible SHG was found for the gold film alone (counts between 0 and 5).

5.3 Near-field enhancement simulation

To see if the experimentally observed 3-order of magnitude enhancement is reasonable from a theoretical point of view, we considered FDTD simulations. A two-dimensional frequency domain field profile monitor was used to determine the near-field intensity distribution. In Figs. 5(a,c) , the monitor was placed in xy plane, at z = 6 nm (at the center of the silver nanoprism). The other monitor was in the xz plane monitor at y = 0 (see Figs. 5(b,d)).

 figure: Fig. 5

Fig. 5 Near field map of the electric field intensity, (a,b) for a silver nanoprism on 10 nm PMMA, 10 nm Au, 2 nm Ti, glass substrate, at the source wavelength of 808 nm in (a) xy and (b) xz planes. (c,d) shows the same distribution for a silver nanoprism on glass substrate, at the source wavelength of 808 nm in (c) xy and (d) xz planes. The scale bar is logarithmic (base 10).

Download Full Size | PDF

Figures 5(a,b) show FDTD calculations of the near-field intensity (wavelength of 808 nm) for a silver nanoprism hybridized with a 10 nm Au film, with 10 nm PMMA spacer layer. This calculation gives a maximum enhancement of 1420 × the incident intensity. By comparison, we have made a similar calculation for a silver nanoprism on glass and found that the maximum enhancement is 82 × the incident intensity (Figs. 5(c,d)). Therefore, the hybridization process has a net enhancement of 17.3 × in the field intensity, and since the SHG scales as the square, this corresponds to 300 × in the SHG generation, which is half what was seen in the experiments.

While this is reasonable agreement considering the accuracy of numerical calculations, we note that the thin metal film may actually direct the SHG light into the detector, and thereby boost the signal in the experiments [4345]. We have not accounted for this effect in the calculations. In general, our calculations do not account for the electrodynamics of the system at the second harmonic wavelength. While there is some enhancement from gap plasmons in our structure [4648], these are not dominant, as can be seen from the field distribution in the calculations. For thinner gaps, significant field enhancements have been reported recently, even departing from local response theories [49].

We also performed calculations for 680 nm wavelength and found an even greater near-field enhancement for the Ag nanoprism on glass (i.e. without hybridization); therefore, the hybridization does not improve the near-field enhancement on resonance. In other words, if a source is available at 680 nm, it would be better to use that source with a non-hybridized system. The main function of the hybridization then was to tune the resonance to the wavelength of the common Ti:Sapphire source.

6. Conclusion

In conclusion, we have demonstrated that plasmon hybridization of metal nanoparticles to a thin metal film is an effective method to obtain SHG enhancement of close to three orders of magnitude. The enhancement comes from fine tuning of the plasmonic resonance to coincide with the excitation source. In the future, it would be interesting to apply these findings to enhanced multi-photon lithography [15], high-harmonic generation [50] and nonlinear sensing [51].

Acknowledgments

The authors acknowledge funding from the NRAS Research Team Program (BCIC, BCFRST, BC Government). GH acknowledges useful discussions with Dr. Hao Jiang.

References and links

1. A. Lesuffleur, L. K. S. Kumar, and R. Gordon, “Enhanced second harmonic generation from nanoscale double-hole arrays in a gold film,” Appl. Phys. Lett. 88(26), 261104 (2006). [CrossRef]  

2. W. Fan, S. Zhang, N.-C. Panoiu, A. Abdenour, S. Krishna, R. M. Osgood, K. J. Malloy, and S. R. J. Brueck, “Second harmonic generation from a nanopatterned isotropic nonlinear material,” Nano Lett. 6(5), 1027–1030 (2006). [CrossRef]  

3. A. Lesuffleur, L. K. S. Kumar, and R. Gordon, “Apex-enhanced second-harmonic generation by using double-hole arrays in a gold film,” Phys. Rev. B 75(4), 045423–045427 (2007). [CrossRef]  

4. T. Xu, X. Jiao, G.-P. Zhang, and S. Blair, “Second-harmonic emission from sub-wavelength apertures: Effects of aperture symmetry and lattice arrangement,” Opt. Express 15(21), 13894–13906 (2007). [CrossRef]   [PubMed]  

5. F. Eftekhari and R. Gordon, “Enhanced second harmonic generation form noncentrosymmetric nanohole arrayes in a gold film,” IEEE J. Sel. Top. Quantum Electron. 14(6), 1552–1558 (2008). [CrossRef]  

6. T. Xu, X. Jiao, and S. Blair, “Third-harmonic generation from arrays of sub-wavelength metal apertures,” Opt. Express 17(26), 23582–23588 (2009). [CrossRef]   [PubMed]  

7. G. X. Li, Z. L. Wang, S. M. Chen, and K. W. Cheah, “Narrowband plasmonic excitation on gold hole-array nanostructures observed using spectroscopic ellipsometer,” Opt. Express 19(7), 6348–6353 (2011). [CrossRef]   [PubMed]  

8. T. Utikal, T. Zentgraf, T. Paul, C. Rockstuhl, F. Lederer, M. Lippitz, and H. Giessen, “Towards the origin of the nonlinear response in hybrid plasmonic systems,” Phys. Rev. Lett. 106(13), 133901 (2011). [CrossRef]   [PubMed]  

9. M. Lippitz, M. A. van Dijk, and M. Orrit, “Third-harmonic generation from single gold nanoparticles,” Nano Lett. 5(4), 799–802 (2005). [CrossRef]   [PubMed]  

10. M. Pelton, M. Liu, S. Park, N. F. Scherer, and P. Guyot-Sionnest, “Ultrafast resonant optical scattering from single gold nanoroes: Large nonlinearities and plasmon saturation,” Phys. Rev. B 73(15), 155419 (2006). [CrossRef]  

11. A. K. Singh, D. Senapati, A. Neely, G. Kolawole, C. Hawker, and P. C. Ray, “Nonlinear optical properties of triangular silver nanomaterials,” Chem. Phys. Lett. 481(1-3), 94–98 (2009). [CrossRef]  

12. J. Butet, J. Duboisset, G. Bachelier, I. Russier-Antoine, E. Benichou, C. Jonin, and P.-F. Brevet, “Optical second harmonic generation of single metallic nanoparticles embedded in a homogeneous medium,” Nano Lett. 10(5), 1717–1721 (2010). [CrossRef]   [PubMed]  

13. K. Thyagarajan, S. Rivier, A. Lovera, and O. J. F. Martin, “Enhanced second-harmonic generation from double resonant plasmonic antennae,” Opt. Express 20(12), 12860–12865 (2012). [CrossRef]   [PubMed]  

14. H. Harutyunyan, G. Volpe, R. Quidant, and L. Novotny, “Enhancing the nonlinear optical response using multifrequency gold-nanowire antennas,” Phys. Rev. Lett. 108(21), 217403 (2012). [CrossRef]   [PubMed]  

15. G. Volpe, M. Noack, S. S. Aćimović, C. Reinhardt, and R. Quidant, “Near-field mapping of plasmonic antennas by multiphoton absorption in poly(methyl methacrylate),” Nano Lett. 12(9), 4864–4868 (2012). [CrossRef]   [PubMed]  

16. B. K. Canfield, H. Husu, J. Laukkanen, B. Bai, M. Kuittinen, J. Turunen, and M. Kauranen, “Local field asymmetry drives second-harmonic generation in non-centrosymmetric nanodimers,” Nano Lett. 7(5), 1251–1255 (2007). [CrossRef]   [PubMed]  

17. Y. Zhang, N. K. Grady, C. Ayala-Orozco, and N. J. Halas, “Three-dimensional nanostructures as highly efficient generators of second harmonic light,” Nano Lett. 11(12), 5519–5523 (2011). [CrossRef]   [PubMed]  

18. R. Jin, J. E. Jureller, H. Y. Kim, and N. F. Scherer, “Correlating second harmonic optical responses of single Ag nanoparticles with morphology,” J. Am. Chem. Soc. 127(36), 12482–12483 (2005). [CrossRef]   [PubMed]  

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

20. 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(1), 017402–017405 (2005). [CrossRef]   [PubMed]  

21. H. Aouani, M. Navarro-Cia, M. Rahmani, T. P. H. Sidiropoulos, M. Hong, R. F. Oulton, and S. A. Maier, “Multiresonant broadband optical antennas as efficient tunable nanosources of second harmonic light,” Nano Lett. 12(9), 4997–5002 (2012). [CrossRef]   [PubMed]  

22. S. Linden, F. B. P. Niesler, J. Förstner, Y. Grynko, T. Meier, and M. Wegener, “Collective effects in second-harmonic generation from split-ring-resonator arrays,” Phys. Rev. Lett. 109(1), 015502–015506 (2012). [CrossRef]   [PubMed]  

23. M. Hentschel, T. Utikal, H. Giessen, and M. Lippitz, “Quantitative modeling of the third harmonic emission spectrum of plasmonic nanoantennas,” Nano Lett. 12(7), 3778–3782 (2012). [CrossRef]   [PubMed]  

24. T. Hanke, J. Cesar, V. Knittel, A. Trügler, U. Hohenester, A. Leitenstorfer, and R. Bratschitsch, “Tailoring spatiotemporal light confinement in single plasmonic nanoantennas,” Nano Lett. 12(2), 992–996 (2012). [CrossRef]   [PubMed]  

25. 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(8), 1087–1090 (2003). [CrossRef]  

26. P. Nordlander, C. Oubre, E. Prodan, K. Li, and M. I. Stockman, “Plasmon hybridization in nanoparticle dimers,” Nano Lett. 4(5), 899–903 (2004). [CrossRef]  

27. A. M. Funston, C. Novo, T. J. Davis, and P. Mulvaney, “Plasmon coupling of gold nanorods at short distances and in different geometries,” Nano Lett. 9(4), 1651–1658 (2009). [CrossRef]   [PubMed]  

28. 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]   [PubMed]  

29. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [CrossRef]   [PubMed]  

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

31. F. Le, N. Z. Lwin, N. J. Halas, and P. Nordlander, “Plasmonic interactions between a metallic nanoshell and a thin metallic film,” Phys. Rev. B 76(16), 165410 (2007). [CrossRef]  

32. 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]  

33. A. I. Maaroof, J. V. Nygaard, and D. S. Sutherland, “Plasmon hybridization in silver nanoislands as semishells arrays coupled to a thin metallic film,” Plasmonics 6(2), 419–425 (2011). [CrossRef]  

34. D. Y. Lei, A. I. Fernández-Domínguez, Y. Sonnefraud, K. Appavoo, R. F. Haglund Jr, J. B. Pendry, and S. A. Maier, “Revealing plasmonic gap modes in particle-on-film systems using dark-field spectroscopy,” ACS Nano 6(2), 1380–1386 (2012). [CrossRef]   [PubMed]  

35. J. J. Mock, R. T. Hill, Y.-J. Tsai, A. Chilkoti, and D. R. Smith, “Probing dynamically tunable localized surface plasmon resonances of film-coupled nanoparticles by evanescent wave excitation,” Nano Lett. 12(4), 1757–1764 (2012). [CrossRef]   [PubMed]  

36. M. Hu, A. Ghoshal, M. Marquez, and P. G. Kik, “Single particles spectroscopy study of metal-film-induced tuning of silver nanoparticle plasmon resonances,” J. Phys. Chem. C 114(16), 7509–7514 (2010). [CrossRef]  

37. C. Lumdee, S. Toroghi, and P. G. Kik, “Post-fabrication voltage controlled resonance tuning of nanoscale plasmonic antennas,” ACS Nano 6(7), 6301–6307 (2012). [CrossRef]   [PubMed]  

38. R. Jin, Y. Cao, C. A. Mirkin, K. L. Kelly, G. C. Schatz, and J. G. Zheng, “Photoinduced conversion of silver nanospheres to nanoprisms,” Science 294(5548), 1901–1903 (2001). [CrossRef]   [PubMed]  

39. J. E. Millstone, S. J. Hurst, G. S. Métraux, J. I. Cutler, and C. A. Mirkin, “Colloidal gold and silver triangular nanoprisms,” Small 5(6), 646–664 (2009). [CrossRef]   [PubMed]  

40. MicroChem Data Sheet, “NanoTM PMMA and copolymer,” (MicroChem 2001). http://microchem.com/pdf/PMMA_Data_Sheet.pdf.

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

42. E. D. Palik, Handbook of Optical Constants of Solids (Academic Press, 1997).

43. D. Wang, W. Zhu, Y. Chu, and K. B. Crozier, “High directivity optical antenna substrates for surface enhanced Raman scattering,” Adv. Mater. (Deerfield Beach Fla.) 24(32), 4376–4380 (2012). [CrossRef]   [PubMed]  

44. A. Ahmed and R. Gordon, “Directivity enhanced Raman spectroscopy using nanoantennas,” Nano Lett. 11(4), 1800–1803 (2011). [PubMed]  

45. Q. Min, Y. Pang, D. J. Collins, N. A. Kuklev, K. Gottselig, D. W. Steuerman, and R. Gordon, “Substrate-based platform for boosting the surface-enhanced Raman of plasmonic nanoparticles,” Opt. Express 19(2), 1648–1655 (2011). [CrossRef]   [PubMed]  

46. 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(9), 097401 (2006). [CrossRef]   [PubMed]  

47. T. Søndergaard and S. I. Bozhevolnyi, “Strip and gap plasmon polariton optical resonators,” Phys. Status Solidi, B Basic Res. 245(1), 9–19 (2008). [CrossRef]  

48. J. Jung, T. Søndergaard, and S. I. Bozhevolnyi, “Gap plasmon-polariton nanoresonators: Scattering enhancement and launching of surface plasmon polaritons,” Phys. Rev. B 79(3), 035401–035409 (2009). [CrossRef]  

49. C. Ciracì, R. T. Hill, J. J. Mock, Y. Urzhumov, A. I. Fernández-Domínguez, S. A. Maier, J. B. Pendry, A. Chilkoti, and D. R. Smith, “Probing the ultimate limits of plasmonic enhancement,” Science 337(6098), 1072–1074 (2012). [CrossRef]   [PubMed]  

50. S. Kim, J. Jin, Y.-J. Kim, I.-Y. Park, Y. Kim, and S.-W. Kim, “High-harmonic generation by resonant plasmon field enhancement,” Nature 453(7196), 757–760 (2008). [CrossRef]   [PubMed]  

51. J. Butet, I. Russier-Antoine, C. Jonin, N. Lascoux, E. Benichou, and P.-F. Brevet, “Sensing with multipolar second harmonic generation from spherical metallic nanoparticles,” Nano Lett. 12(3), 1697–1701 (2012). [CrossRef]   [PubMed]  

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 (a) Extinction of silver nanoparticles in aqueous solution as a function of irradiation time. Inset: Scanning electron microscope image of a typical silver nanoprism after 72 h irradiation. (b,c) Atomic force microscope (AFM) characterization of silver nanoprisms spin coated on a silicon substrate. (d) The height of a typical silver nanoprism was measured to be 12 nm with the AFM.
Fig. 2
Fig. 2 (a) Schematic of silver nanoprisms with PMMA spacer layer over a 10 nm thick Au film adhered to a glass substrate with a 2 nm Ti layer. (b) Schematic of scattering measurement setup. WLS = white light source, obj = microscope objective lens. (c) Scattering measurement for three hybrid structures with different PMMA spacer layer thicknesses (shown in legend). Green dotted line: Ti:Sapphire spectrum. (d) Scattering simulation results for silver nanoprisms for the corresponding spacer thicknesses.
Fig. 3
Fig. 3 Peak scattering wavelength as a function of PMMA spacer layer thickness from experiment and FDTD simulations.
Fig. 4
Fig. 4 (a) Schematic of SHG measurement setup. ND = neutral density filter, obj = microscope objective lens, BG40 = blue-green band pass filter. (b) SHG measured using spectrum analyzer in front of streak camera (not shown in (a)). (c) SHG count dependence on PMMA spacer layer thicknesses. (d) Log-log plot of SHG count vs. incident laser power with linear-fit slope of 2.17 ± 0.20. For SHG, a slope of 2 is expected.
Fig. 5
Fig. 5 Near field map of the electric field intensity, (a,b) for a silver nanoprism on 10 nm PMMA, 10 nm Au, 2 nm Ti, glass substrate, at the source wavelength of 808 nm in (a) xy and (b) xz planes. (c,d) shows the same distribution for a silver nanoprism on glass substrate, at the source wavelength of 808 nm in (c) xy and (d) xz planes. The scale bar is logarithmic (base 10).
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.