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

Photoluminescence of two-dimensional plasmonic structures: enhancement, spectral and lifetime peculiarities below the lasing threshold

Open Access Open Access

Abstract

We study the interplay between optical properties and photoluminescence spectra of structures based on square lattices of silver nanodiscs or nanoholes in silver embedded in or adjoined to a dye-doped polymer waveguiding layer. The lattices and the waveguiding gain layer were designed to overlap the spectral features of such two-dimensional (2D) plasmonic structures. By varying the lattice constant, we demonstrate a condition for a significant increase of photoluminescence intensity and compare lifetime reduction for both nanodiscs- and nanoholes-based structures, when pumping them below lasing threshold. These enhancements are reached by matching the dark modes of the structure with the gain medium fluorescence spectrum.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Noble metal scatterers combined with a gain medium attracted attention because of the emergence of concepts of the spaser [1] and the plasmon nanolaser [24], and their further development–the plasmon-supported solid state light source [58]. Plasmon nanolasers are believed to be essential for advancing integrated optical circuits since plasmonic waves are spatially confined strong near fields that can afford the next level of device miniaturization due to reduction of the physical size, mode volume of the laser below the diffraction limit and ultra-fast modulation rates [9].

Due to capabilities of the modern fabrication technologies, a number of realizations of plasmon nanolasers are demonstrated, where a 2D plasmonic structure is based on either nanodisc or nanohole lattice embedded into or combined with a solid or liquid gain-material waveguide [24,10]. The features in transmission and reflection spectra of these 2D structures are mainly associated with resonant phenomena of two types [11,12]: (i) collective excitation of nanodiscs (electric dipoles) or nanoholes (magnetic dipoles) resulting in localized [1315] or surface [16] grating plasmon resonances (LGPR and SGPR) and (ii) coupling light scattered by a periodic nanoparticle or nanohole array to waveguide modes of a dielectric slab. These 2D structures (their optical spectra) can be complementary to each other and build a Babinet-like metasurface [17].

Much attention has been paid to studies on coupling between the diffraction orders and SGPR, when discussing extraordinary transmission and taking into account interplay between the SGPR and a quasi-cylindrical wave for designing nanolasers [1821]. It is reported that the best nanoholes-based realizations of nanolasers are either for diffraction orders of (1,1) [10,22,23] or when the gain layer is too thin to support the quasi-cylindrical waves, thus efficiently supporting only the SGPR [24]. However, there is a limited number of publications discussing nanolaser designs having a dye-doped polymer layer of a waveguiding thickness [4,10]. For the nanolasers based on square lattices of nanodiscs, the designs considered in literature are built to utilize the (1,0) diffraction orders [2528]. And, indeed, the question rises, which 2D plasmonic structure is more suitable for achieving lasing, the nanohole- or nanodiscs-based one? Another question is what if a waveguiding layer of a chosen gain medium is utilized to build photoluminescent 2D structures of both types? Last but not the least issue important for the nanolasers is the evolution of the fluorescence lifetime–its reduction due to the Purcell effect and photoluminiscence enhancement by structures [4,2933].

Despite extensive research on the plasmon nanolaser [34,35], there is no experimental comparative study where the interplay between spectral features and evolution of photoluminescence (PL) is discussed in detail for both nanodiscsс- and nanoholes-based plasmon nanolaser. Comparison between properties of such structures fabricated on the basis of the identical materials is worthy of consideration.

In our paper, we discuss the dependence of PL on optical properties of 2D structures. Under study are the silver nanodisc- and nanohole-based square lattices embedded into or adjoined to a waveguiding layer made from a polymer doped with fluorescent dye–negative photoresist SU-8 doped with Rhodamine 101 (SU-8–R101). For reference, we also study responses from the SU-8–R101 layers of the identical thickness spincoated onto a silica and silica/100 nm-thick Ag substrates. Resonances in spectra of the structures were designed to overlap the band of dye photoluminescence. All experiments were done for low pumping intensities to visualize correspondence between the spectral features due to a particular nanostructure and the photoluminescence spectrum together with the dye lifetime reduction.

2. Samples, experiment and calculations

Square lattices were lift-off deposited silver nanodiscs [Fig. (1a), (1c)] with nanometre thick chromium underlayer and nanostructured silver films fabricated on the UV-grade fused silica substrates using e-beam lithography and dry reactive-ion etching [Fig. (1b), (1d)]. The structures were nanodiscs- (samples ND hereafter) or nanoholes-based (perforated films, samples NH hereafter) square lattices with dimensions (chip size) of 500 µm × 500 µm; a 3 × 3 matrix of chips was located per the substrate. Samples ND had the following periods and parameters: D = 385, 395 and 405 nm, a disc height of 30 nm and diameter of 100 nm [ Fig. 1(a)]. For samples NH, periods were D = 545, 555 and 575 nm, a hole diameter–200 nm and a thickness–100 nm [Fig. 1(b)]. Note that the thickness of 100 nm and periods are choosen in accordance with our experience in fabrication of monocrystalline silver films, which are known to have atomically flat surfaces for thicknesses > 50 nm [36] and the related nanostructurs are demonstrated in sensing applications [37]. The holes were etched through the full thickness of the silver film; the etching rate was approximately 60 nm/min and the time –2 minutes.

 figure: Fig. 1.

Fig. 1. (a) and (b) Sketches of 2D plasmonic structures under study. (c) and (d) SEM images of initial 2D lattices of nanodiscs and nanoholes.

Download Full Size | PDF

R101 powder was directly dissolved in SU-8 3050 (MicroChem Corp.) formulation based on cyclopentanone solvent, R101 mass concentration was 1% relative to SU-8 solid content. The solutions of R101 in SU-8 were spincoated on top of samples ND&NH, clean UV-grade silica and UV-grade silica/100 nm-thick Ag substrates and then cured at 95°C according to recommendation from Su-8 datasheet. Spectral and material parameters of the fabricated SU-8–R101 films are shown in Fig. 2. The thickness of these films measured with Zygo NewView 7300 optical profiler was hWG = 430 ± 10 nm. Note that the selected thickness was enough for the single TE mode regime in the studied spectral range of 500-900 nm, it was optimized so that to effectively excite the waveguide mode at λWG = λPL ≈ 600 nm–the maximum in the photoluminescence spectrum of SU-8–R101. Extinction/photoluminiscence was measured by a Shimadzu UV3600 spectrophotometer/an Ntegra Spectra spectrometer.

 figure: Fig. 2.

Fig. 2. (a) Normalized extinction (Ext) and photoluminescence (PL) of the SU-8–R101 films; a pumping wavelength of 473 nm is indicated by the arrow. (b) Measured refractive index of bulk SU-8–R101 (stars) and calculated ones of a SU-8–R101 planar waveguide on Ag (TM0–curve 2, TE0–3 and SPR–curve 6) and on UV-grade (TM0–curve 4 and TE0–curve 5). Fitted curve 1 was used for numerical calculation.

Download Full Size | PDF

In the photoluminescence experiments, for the pumping of R101 dye molecules, a pulsed laser operating at a wavelength of 473 nm with a repetition rate of 40 MHz and a pulse duration of 120 ps was utilized. Laser beam had a Gaussian beam profile and was focused in a focal spot with a diameter of 5 µm FWHM by an objective (10 × NA = 0.3). The pumping intensity in a single pulse in the focal spot was less than 10 kW/cm2. At these illumination conditions, photoluminescence intensity from pumped areas kept at a constant level for hours, which evidenced the absence of considerable bleaching of the dye. Photoluminescence spectra were measured in the reflection geometry where illumination and collection was done by the objective. To obtain full data, photoluminescence and transmission spectra were analyzed together with lifetime dynamics. The properties were studied by the Ntegra Spectra (NT-MDT, Russia) multichannel microscope&spectrometer having a fluorescence lifetime imaging microscopy (FLIM) channel. The measurements were carried out for the Ex-polarized light (see Fig. 1) at normal and oblique incidence (α = 0-2°). To obtain transmission spectra, a collimated beam from a white light source (a Nikon illuminator, 75 W) was prepared and reduced to a diameter of 300 µm by an aperture located in front of the chip under study; the beam divergence angle of the beam is less than 0.1°.

Numerical calculations were done by using Comsol Multiphysics. The calculation model of ND&NH structures was done for a single unit cell with periodic (Bloch) boundary conditions along the x and y axes and had the structural parameters shown in Fig. 1. To calculate the transmission spectra, the periodic ports were applied across the z axis to launch the incident wave. To calculate eigenmodes, perfectly matched layers were applied. Dispersion of the optical constants for Ag (ɛAg) was taken from Ref. [36,38], and that of SU-8–R101 were obtained by a VASE Woollam ellipsometer [see Fig. 2(b)]. Effective refractive indexes of waveguide modes in [Fig. 2(b)] were calculated by the T-matrix method [39,40].

3. Experimental results and discussion

Measured and calculated transmission spectra of ND&NH structures are demonstrated in Fig. 3. Firstly, let us discuss features in transmission spectra of samples ND. The most red wavelength-shifted band λLGPR is due to collective excitation of nanodiscs, and its spectral position at α = 0 is interpreted as the condition of first-order diffraction from grating [1315]:

$$ \lambda_{LGPR} = n_{EFF} D,$$
where λLGPR is a vacuum wavelength, D is a period, and nEFF is an effective refractive index of the medium, which can be estimated for 2D structures under study in Refs. [13,14] by the Maxwell Garnett approximation [41]:
$${n_{EFF\;MG}} = \sqrt {{\varepsilon _h}{{\left( {1 + 2f\frac{{Re \{{{\varepsilon_i}} \}- {\varepsilon_h}}}{{Re \{{{\varepsilon_i}} \}+ 2{\varepsilon_h}}}} \right)} / {\left( {1 - f\frac{{Re \{{{\varepsilon_i}} \}- {\varepsilon_h}}}{{Re \{{{\varepsilon_i}} \}+ 2{\varepsilon_h}}}} \right)}}}. $$
Here ɛh is a dielectric permittivity of the host medium (SU-8–R101), ɛi is that of a metal (treated as an inclusion), and f is a volume fraction of the metal inside the dielectric matrix medium with thickness of d. In accordance with our calculations, nEFF depends on the electric field distribution between the dielectric and vacuum and, therefore, tends to its maximum as d rises. Studies on the angle- and polarization-resolved spectra of analogous nanostructure can be found in Ref. [13], where discussion on spectral shifts of the λLGPR-band is given in detail.

 figure: Fig. 3.

Fig. 3. (а) Transmission spectra of samples ND with D = 385 (1), 395 (2) and 405 nm (3); line 4–a calculated spectrum for D = 405 nm. (b) Transmission spectra of samples NH with D = 545 (1), 555 (2) and 575 (3) nm, a hole diameter–dh = 200 nm; line 4–a calculated spectrum for D = 545 nm.

Download Full Size | PDF

The peaks at λWG = 570, 585 and 590 nm for samples ND with D = 385, 395 and 405 nm, correspondingly, were due to excitation of a TE0 waveguide mode. For this mode, the well-known expression [42] of

$$h_{W G}=\frac{\lambda_{W G}}{2 \pi \sqrt{n_{S U-8}^{2}-n_{U V-\text {grade}}^{2}}}\left(\pi \rho+\operatorname{arctg}\left(\frac{\mathrm{n}_{\mathrm{SU}-8}}{\mathrm{n}_{\mathrm{UV}-\mathrm{grade}}}\right)^{\chi} \cdot \sqrt{\frac{n_{U V-\text {grade}}^{2}-n_{\text {Air}}^{2}}{n_{\mathrm{SU}-8}^{2}-n_{U V-\text {grade}}^{2}}}\right)$$
$\chi = 0$ for TE waves, $\chi = 2$ for TM waves, ρ = 0, 1, 2… is the mode index TE0, TE1,… TM0, TM1,… etc. for the SU-8 waveguide on the UV-grade substrate, returns that minimal thickness of d = 430 nm corresponds to λWG = 590 nm. Indeed, if one calculates a product of D and nTE0, λWG = 590 nm is obtained for sample ND with D = 405 nm and d = 430 nm; Fig. 2(b) illustrates a dispersion of nTE0 calculated by the transfer matrix method [38,39]. Finally, the most blue-shifted band was addressed to the (1,0) order diffraction to the ТЕ1waveguide mode resulting in λLGPR(1,0)-band at 525 nm for the measured spectral range [Fig. 3(a)].

As for samples NH, a series had periods of D in a range of 545-575 nm to overlap λSGPR and λPL as discussed in Refs. [10,20], illustrating nanolasing for diffraction orders of {1,1}. Dispersions of excited SGPR follow the next formula:

$${\overrightarrow {{\; }k} _{spp}} = {\overrightarrow {{\; }k} _{x{\; }}} + l{\overrightarrow {{\; }G} _{x{\; }}} + {\; }m{\overrightarrow {{\; }G} _{y{\; }}}{\; }\, = {\; }\,\frac{{2\pi }}{\lambda }\sqrt {\frac{{Re({{\varepsilon_{Ag}}} )Re({{\varepsilon_d}} ){\; }}}{{Re({{\varepsilon_{Ag}}} )+ Re({{\varepsilon_d}} )}}{\; }}, $$
$${|{{{\overrightarrow {{\; }k} }_{spp}}} |^2} = {\left( {\frac{{2\pi }}{\lambda }sin\theta + l\frac{{2\pi }}{D}} \right)^2} + {\left( {m\frac{{2\pi }}{D}} \right)^2}, $$
where ${\overrightarrow {|\; G} _{x\; }}|= |{{{\overrightarrow {\; G} }_{y\; }}} |= 2\pi /D$ are the reciprocal lattice vectors, l is the diffraction order in the x-axis, m is the diffraction order in the y-axis, λ is the free space wavelength and nEFF is the effective index of the mode and $\theta $ is the angle of incidence of the light beam.

At normal incidence we obtain:

$$\frac{{{l^2} + {m^2}}}{{n_{EFF}^2}} = \frac{{{D^2}}}{{{\lambda ^2}}}{\; }. $$
Note that, in our consideration, the condition of λSGPR(1,1) ≈ λWGλPL was satisfied. Figure 3(b) shows experimental and calculated transmission spectra of samples NH. The spectra illustrate bands with the minima corresponding to the resonant wavelengths λSGRP governed by D, l and m and, also, the dielectric background–nSU-8 and nUV-grade. For clarity, resonant wavelengths are noted for the next interfaces of sample NH with D = 545 nm: Ag/SU-8–λSGPR{1,1} = 570 nm and λSGPR(1,0) = 769 nm; UV-grade/Ag –λSGPR(1,1) = 602 nm and λSGPR(1,0) = 816 nm (not shown). If one calculates using Eqs. (1) and (5), the spectral positions of discussed resonances in the experimental spectra were in a good agreement with the calculated nEFF of the modes presented in Fig. 2(b). The minimum at 570 nm according to Eq. (5) returned a measured value of nEFF = 1.477 and a corresponding calculated value was 1.48 for the TE0 mode. The minimum at 602 nm attributed to a Wood anomaly also demonstrates agreement with Eq. (5), thus giving nEFF = 1.562. The minimum at 690 nm according to Eq. (1) returned a measured value of nEFF = 1.266 and a corresponding calculated value was 1.248 for the TM0 mode.

In PL spectra of samples ND (Fig. 4), we observed the following main features: (i) the PL intensity raised as λWG moved to λPL of the spectrum of SU-8–R101 layer [also, see Fig. 3(a)], (ii) for sample ND with D = 405 nm (lines 3), the PL band got spectrally split–one could resolve a peak at λWG and a spectrally narrow doublet, and (iii) maxima of the doublet corresponded to peaks in the transmission spectrum at oblique incidence (α = 2°) in a range of 610-620 nm [see the calculated spectrum in Fig. 4(b)]. Definitely, the observed increase in the PL intensity for the sample with D = 405 nm with respect to the other periods was due to the excitation of D1, D2 and D3 dark modes, which have higher Q-factors than the bright modes. Let us consider the bright and dark mode in more details.

 figure: Fig. 4.

Fig. 4. (а) Photoluminescence spectra of SU-8–R101 (dashed line) and samples ND with D = 385 (1), 395 (2) and 405 nm (3). (b) For the same samples ND, experimental transmission spectra at oblique incidence (α = 2°) and a calculated one for D = 405 nm (solid line). Eigen modes of sample ND with D = 405 nm are noted: λWG and λLGPR are bright modes and D1, D2 and D3 are dark modes.

Download Full Size | PDF

For system ND with D = 405 nm, distributions of the E- and H-fields in the unit cell were calculated (Fig. 5). Two bright modes are obtained at zero tangential wavenumber (which corresponds to the excitation at normal incidence) at λLGPR (panel b) and λWG (panel c). Both modes featured the electric dipole field configuration at nanodiscs. It is worth mentioning again that the mode observed at λLGPR [Fig. 5(b)] is the grating-associated localized plasmon hybridized with the TE mode. The mode observed at λWG [Fig. 5(c)] is a TE guided mode scattered by the array of nanodiscs [11]. Moreover, a set of dark modes exist in the system [43,44]. These modes can be excited at the oblique incidence only and are observed as the three (D1, D2, D3) narrow resonances in the PL and transmission spectra (Fig. 4). The dark modes are due to a hybridization between TE (or TM) modes of the SU-8 waveguide and different plasmonic multipolar modes of the nanodiscs. These modes are clearly identified basing on the symmetry of electric and magnetic fields with respect to the center of the nanoparticle. Correspondence of this symmetry to that of different electric and magnetic multipoles is illustrated in Fig. 5, which shows the calculated field distributions. We identify the modes as:

  • (i) D1 is a mode due to coupling between an electric quadrupole plasmon resonance of nanodiscs and the TE mode;
  • (ii) D2 is a mode due to coupling between the TE mode and a magnetic dipole plasmon mode oscillating along the z-axis;
  • (iii) D3 is a mode due to coupling between the TM mode and a magnetic dipole plasmon mode oscillating along the y-axis.

 figure: Fig. 5.

Fig. 5. (color online) Calculation model (a) and distribution of the |E|-field in the unit cell of sample ND: (b, c) bright and dark (d-f) eigenmodes. The instantaneous value is shown in pink and green vectors for the E- and H-fields.

Download Full Size | PDF

PL spectra of samples NH are demonstrated in Fig. 6. Whereas the transmission spectra of these samples had more complicated structure, the tendencies in them were similar to that of samples ND. Both the PL intensity and spectral positions of narrow resonant doublet resolved in the PL spectra were dependent on D. A significant difference between the PL spectra of samples ND and NH was in the intensities of the doublet peaks and their visualization conditions: ND–intensive doublet was seen only for D = 405 nm, NH–low-intensity doublet was resolved for every D.

 figure: Fig. 6.

Fig. 6. (а) Photoluminescence spectra of an SU-8–R101 layer on Ag (dashed line) and samples NH with D = 545 (line 1), 555 (2) and 575 nm (3). (b) For same sample NH with D = 545 nm, experimental and calculated spectra at normal and oblique incidence. Experimental spectra at oblique incidence did not reveal any dark mode resonances.

Download Full Size | PDF

In fact, maxima of the PL spectra for sample NH with D = 545 nm correlated with the transmission spectra and observed in between of plasmon resonances λSGPR(1,1) = 602 nm and λSGPR(1,1) = 641 nm (α = 2°) on UV-grade/Ag and Ag/SU-8 interface respectively. Low-intensity peaks in PL spectra coincided with the corresponding features in transmission spectra at λ = 617 and 625 nm that were likely manifestations of counter-propagating TE0 modes [Fig. 2(b), curve 3]. These modes can be resolved in a range of 610-630 nm; see features shown by arrows in the experimental and calculated spectra. Note that, for the studied series of samples NH, the resonance of λSGPR(1,1) excited on UV-grade/Ag interface influenced the intensity of PL spectrum, and the one on Ag/SU-8 interface did not.

The next important parameter under discussion for nanolasers is a reduction of the fluorescence lifetime; such data in Fig. 7 are summarized for the studied samples. The lifetimes at 619 nm were as follows: τ0 (UV-grade/SU-8–R101) = 3.49 ± 0.04 ns, τ (sample ND) = 3.04 ± 0.01, τ (sample NH) = 2.83 ± 0.01 and were obtained from analysis of the exponential decay profiles from plot (a). For a detection wavelength of 619 nm, plot (a) shows fluorescence kinetics of the SU-8–R101 layer on the UV-grade substrate and selected structures. It was observed that, in the vicinity of λWG in the case of sample Ag/SU-8–R101, ND and λSGPR(1,1)λWGλPL for sample NH, the fluorescence lifetime was reduced. The normalized lifetimes (a ratio of τi0) versus the detection wavelength are given in plot (b). For all studied structures, one can see a decrease of τi0 with a minimum corresponding well to the maximum of the photoluminescence spectrum of SU-8–R101 on the UV-grade substrate. The shorter lifetime (signature of larger Purcell factor [45]) was for sample NH, and a spectrally narrower feature in the τi0 curve was for sample ND. It is known that the fluorescence lifetime decreases as the density of states rises. For sample NH, this happened due to the surface plasmons–the highly absorbing channel. As for sample ND, the density of states rises at frequencies of the dark modes, which are the low absorbing channels.

 figure: Fig. 7.

Fig. 7. For the selected samples: (a) photoluminescence decay, (b) a normalized fluorescence lifetime versus wavelength.

Download Full Size | PDF

4. Conclusion

Optical properties of 2D plasmonic structures, where square lattices were fabricated from silver, and a waveguide layer was made of polymer doped by optical dye (SU-8–R101), were studied at low-intensive pumping in detail. Observed were a significant rise of photoluminescence intensity and shortening of the fluorescence lifetime for frequencies corresponding to high-Q optical resonances. Since in the studied structures, the resonances are accompanied by unavoidable absorption, we speculate that the best condition for demonstrating the nanolasing regime is when the low-damping dark modes are excited, i.e. in the case of the considered structures based on nanodiscs; samples ND should have a lower lasing threshold.

Acknowledgements

We thank the technology team at the BMSTU Nanofabrication Facility (FMN Laboratory, FMNS REC, ID 74300) for sample fabrication. This work was done during implementation of a research project under No. 7/004/2013-2018 funded by Advanced Research Foundation (Russia).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. D. J. Bergman and M. I. Stockman, “Surface plasmon amplification by stimulated emission of radiation: quantum generation of coherent surface plamons in nanosystems,” Phys. Rev. Lett. 90(2), 027402 (2003). [CrossRef]  

2. V. Noginov, G. Zhu, A. Belgrave, R. Bakker, V. Shalaev, S. Stout, E. Herz, T. Suteewong, and U. Wiesner, “Demonstration of a spaser-based nanolaser,” Nature 460(7259), 1110–1112 (2009). [CrossRef]  

3. M. P. van Exter, V. T. Tenner, F. van Beijnum, M. J. A. de Dood, P. J. van Veldhoven, E. J. Geluk, and G. W.’t Hooft, “Surface plasmon dispersion in metal hole array lasers,” Nature 492(7429), 411–414 (2012). [CrossRef]  

4. W. Zhou, M. Dridi, J. Y. Suh, C. H. Kim, D.T. Co, M. R. Wasielewski, G. C. Schatz, and T. W. Odom, “Lasing action in strongly coupled plasmonic nanocavity arrays,” Nat. Nanotechnol. 8(7), 506–511 (2013). [CrossRef]  

5. J. S. Biteena, N. S. Lewis, and H. A. Atwater, “Spectral tuning of plasmon-enhanced silicon quantum dot luminescence,” Appl. Phys. Lett. 88(13), 131109 (2006). [CrossRef]  

6. G. Lozano, D. J. Louwers, S. R. Rodrıguez, S. Mu-rai, O. T. Jansen, M. A. Verschuuren, and J. G. Rivas, “Plasmonics for solid-state lighting: Enhanced excitation and directional emission of highly efficient light sources,” Light: Sci. Appl. 2(5), e66 (2013). [CrossRef]  

7. J. Henson, E. Dimakis, J. DiMaria, R. Li, S. Minissale, L. D. Negro, T. D. Moustakas, and R. Paiella, “Enhanced near-green light emission from InGaN quantum wells by use of tunable plasmonic resonances in silver nanoparticle arrays,” Opt. Express 18(20), 21322 (2010). [CrossRef]  

8. J. Henson, J. DiMaria, E. Dimakis, T. D. Moustakas, and R. Paiella, “Plasmon-enhanced light emission based on lattice resonances of silver nanocylinder arrays,” Opt. Lett. 37(1), 79–81 (2012). [CrossRef]  

9. K. S. Daskalakis, A.I. Väkeväinen, J.-P. Martikainen, T. K. Hakala, and P. Törmä, “Ultrafast pulse generation in an organic nanoparticle-array laser,” Nano Lett. 18(4), 2658–2665 (2018). [CrossRef]  

10. X. Meng, J. Liu, A. V. Kildishev, and V. M. Shalaev, “Highly directional spaser array for the red wavelength region,” Laser Photonics Rev. 8(6), 896–903 (2014). [CrossRef]  

11. C. Tan, J. Simonen, and T. Niemi, “Hybrid waveguide-surface plasmon polariton modes in a guided-mode resonance grating,” Opt. Commun. 285(21-22), 4381–4386 (2012). [CrossRef]  

12. A.I. Väkeväinen, R. J. Moerland, H. T. Rekola, A.-P. Eskelinen, J.-P. Martikainen, D.-H. Kim, and P. Törmä, “Plasmonic surface lattice resonances at the strong coupling regime,” Nano Lett. 14(4), 1721–1727 (2014). [CrossRef]  

13. A. N. Shaimanov, K. M. Khabarov, A. M. Merzlikin, I.V. Bykov, and A. V. Baryshev, “Plasmon resonances in a two-dimensional lattice of metal particles in a dielectric layer: Structural and polarization properties,” J. Exp. Theor. Phys. 124(4), 584–591 (2017). [CrossRef]  

14. A. N. Shaimanov, K. M. Khabarov, and A. V. Baryshev, “Plasmonic magneto-optical nested 2D nanostructures: Tailoring responses through effective refractive index,” J. Magn. Magn. Mater. 485, 54–60 (2019). [CrossRef]  

15. B. Lamprecht, G. Schider, R. T. Lechner, H. Ditlbacher, J. R. Krenn, A. Leitner, and F. R. Aussenegg, “Metal nanoparticle gratings: Influence of dipolar particle interaction on the plasmon resonance,” Phys. Rev. Lett. 84(20), 4721–4724 (2000). [CrossRef]  

16. H. Raether, “Surface plasmons on smooth and rough surface and grating,” in Springer Tracts in Modern Physics, 111 (Springer, 1988).

17. V. V. Kornienko, A. N. Shaimanov, and A. V. Baryshev, “Overlapping the electric and magnetic dipole resonances of a silver 2D Babinet-type metasurface: Broadband high reflectance with local field enhancement,” J. Appl. Phys. 126(6), 063102 (2019). [CrossRef]  

18. F. J. Garcia-Vidal, L. Martin-Moreno, T. W. Ebbesen, and L. Kuipers, “Light passing through subwavelength apertures,” Rev. Mod. Phys. 82(1), 729–787 (2010). [CrossRef]  

19. A. Y. Nikitin, F. J. Garcia-Vidal, and L. Martin-Moreno, “Surface electromagnetic field radiated by a sub-wavelength hole in a metal film,” Phys. Rev. Lett. 105(7), 073902 (2010). [CrossRef]  

20. H. T. Liu and P. Lalanne, “Microscopic theory of the extraordinary optical transmission,” Nature 452(7188), 728–731 (2008). [CrossRef]  

21. F. van Beijnum, C. Retif, H. L. Chris, B. Smiet, and P. L. M. P. van Exter, “Quasi-cylindrical wave contribution in experiments on extraordinary optical transmission,” Nature 492(7429), 411–414 (2012). [CrossRef]  

22. P. Melentyev, A. Kalmykov, A. Gritchenko, A. Afanasiev, V. Balykin, A. Baburin, E. Ryzhova, I. Filippov, I. Rodionov, I. Nechepurenko, A. Dorofeenko, I. Ryzhikov, A. Vinogradov, A. Zyablovsky, E. Andrianov, and A. Lisyansky, “Plasmonic nanolaser for intracavity spectroscopy and sensorics,” Appl. Phys. Lett. 111(21), 213104 (2017). [CrossRef]  

23. A. S. Baburin, A. I. Ivanov, I.V. Trofimov, A.A. Dobronosova, P.N. Melentiev, V.I. Balykin, D.O. Moskalev, A.A. Pishchimova, L.A. Ganieva, I.A. Ryzhikov, and I. A. Rodionov, “Highly directional plasmonic nanolaser based on high-performance noble metal film photonic crystal,” Proc. SPIE 10672, 106724D (2018). [CrossRef]  

24. Vasco T. Tenner, Michiel J. A. de Dood, and M. P. van Exter, “Measurement of the phase and intensity profile of Surface Plasmon Laser Emission,” ACS Photonics 3(6), 942–946 (2016). [CrossRef]  

25. A. Yang, Z. Li, M. P. Knudson, A. J. Hryn, W. Wang, K. Aydin, and T. W. Odom, “Unidirectional Lasing from template stripped Two-dimensional Plasmonic Crystals,” ACS Nano 9(12), 11582–11588 (2015). [CrossRef]  

26. A. Yang, T. B. Hoang, M. Dridi, C. Deeb, M. H. Mikkelsen, G. C. Schatz, and T. W. Odom, “Real-time tunable lasing from plasmonic nanocavity arrays,” Nat. Commun. 6(1), 6939 (2015). [CrossRef]  

27. A. H. Schokker and A. F. Koenderink, “Lasing at the band edges of plasmonic lattices,” Phys. Rev. 90(15), 155452 (2014). [CrossRef]  

28. L. Lin and Y. Yi, “Orthogonal and parallel lattice plasmon resonance in core-shell SiO2/Au nanocylinder arrays,” Opt. Express 23(1), 130–142 (2015). [CrossRef]  

29. E. M. Purcell, “Spontaneous emission probabilities at radio frequencies,” Phys. Rev. 69(1-2), 37–38 (1946). [CrossRef]  

30. Marjan Saboktakin, Xingchen Ye, Uday K. Chettiar, Nader Engheta, Christopher B. Murray, and Cherie R. Kagan, “Plasmonic Enhancement of Nanophosphor Upconversion Luminescence in Au Nanohole Arrays,” ACS Nano 7(8), 7186–7192 (2013). [CrossRef]  

31. V. Giannini, A. I. Fernandez-Dominguez, S. C. Heck, and S. A. Maier, “Plasmonic nanoantennas: fundamentals and their use in controlling the radiative properties of nanoemitters,” Chem. Rev. 111(6), 3888–3912 (2011). [CrossRef]  

32. S. Murai, M. A. Verschuuren, G. Lozano, G. Pirruccio, S. R. K. Rodriguez, and J. Gómez Rivas, “Hybrid plasmonic-photonic modes in diffractive arrays of nanoparticles coupled to light-emitting optical waveguides,” Opt. Express 21(4), 4250 (2013). [CrossRef]  

33. S. Kuhn, U. Hakanson, L. Rogobete, and V. Sandoghdar, “Enhancement of Single-Molecule Fluorescence Using a Gold Nanoparticle as an Optical Nanoantenna,” Phys. Rev. Lett. 97(1), 017402 (2006). [CrossRef]  

34. A. V. Dorofeenko, A. A. Zyablovsky, A. P. Vinogradov, E. S. Andrianov, A. A. Pukhov, and A. A. Lisyansky, “Steady state superradiance of a 2D-spaser array,” Opt. Express 21(12), 14539–14547 (2013). [CrossRef]  

35. Y. E. Lozovik, I. A. Nechepurenko, A. V. Dorofeenko, E. S. Andrianov, and A. A. Pukhov, “Highly sensitive spectroscopy based on a surface plasmon polariton quantum generator,” Laser Phys. Lett. 11(12), 125701 (2014). [CrossRef]  

36. A. S. Baburin, A. M. Merzlikin, A. V. Baryshev, I. A. Ryzhikov, Y. V. Panfilov, and I. L. A. Rodionov, “Silver-based plasmonics: golden material platform and application challenges,” Opt. Mater. Express 9(2), 611–642 (2019). [CrossRef]  

37. A.S. Baburin, A.S. Gritchenko, N.A. Orlikovsky, A.A. Dobronosova, I.A. Rodionov, V.I. Balykin, and P.N. Melentiev, “State-of-the-art plasmonic crystals for molecules fluorescence detection,” Opt. Mater. Express 9(3), 1173–1179 (2019). [CrossRef]  

38. G.M. Yankovskii, A.V. Komarov, R.S. Puz’sko, A.V. Baryshev, K.N. Afanas’ev, I.A. Boginskaya, I.V. Bykov, A.M. Merzlikin, I.A. Rodionov, and I.A. Ryzhikov, “Structural and Optical Properties of Single and Bilayer Silver and Gold Films,” Phys. Solid State 58(12), 2503–2510 (2016). [CrossRef]  

39. M. Born and E. Wolf, Principles of Optics, (Pergamon Press, 1970).

40. S. Zouhdi, A. V. Dorofeenko, A. M. Merzlikin, and A. P. Vinogradov, “Theory of zero-width band gap effect in photonic crystals made of metamaterials,” Phys. Rev. B 75(3), 035125 (2007). [CrossRef]  

41. A. H. Sihvola and J. A. Kong, “Effective permittivity of dielectric mixtures,” IEEE Trans. Geosci. Remote Sensing 26(4), 420–429 (1988). [CrossRef]  

42. H. Unger and E. Conwell, Planar Optical Waveguides and Fibres, (Oxford University Press, 1977).

43. W. Zhou and T. Odom, “Tunable subradiant lattice plasmons by out-of-plane dipolar interactions,” Nat. Nanotechnol. 6(7), 423–427 (2011). [CrossRef]  

44. T. K. Hakala, H.T. Rekola, A.I. Väkeväinen, J.-P. Martikainen, M. Nečada, A.J. Moilanen, and P. Törmä, “Lasing in dark and bright modes of a finite-sized plasmonic lattice,” Nat. Commun. 8(1), 13687 (2017). [CrossRef]  

45. A. Krasnok and A. Alu, “Active Nanophotonics,” Proc. IEEE 108(5), 628–654 (2020). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. (a) and (b) Sketches of 2D plasmonic structures under study. (c) and (d) SEM images of initial 2D lattices of nanodiscs and nanoholes.
Fig. 2.
Fig. 2. (a) Normalized extinction (Ext) and photoluminescence (PL) of the SU-8–R101 films; a pumping wavelength of 473 nm is indicated by the arrow. (b) Measured refractive index of bulk SU-8–R101 (stars) and calculated ones of a SU-8–R101 planar waveguide on Ag (TM0–curve 2, TE0–3 and SPR–curve 6) and on UV-grade (TM0–curve 4 and TE0–curve 5). Fitted curve 1 was used for numerical calculation.
Fig. 3.
Fig. 3. (а) Transmission spectra of samples ND with D = 385 (1), 395 (2) and 405 nm (3); line 4–a calculated spectrum for D = 405 nm. (b) Transmission spectra of samples NH with D = 545 (1), 555 (2) and 575 (3) nm, a hole diameter–dh = 200 nm; line 4–a calculated spectrum for D = 545 nm.
Fig. 4.
Fig. 4. (а) Photoluminescence spectra of SU-8–R101 (dashed line) and samples ND with D = 385 (1), 395 (2) and 405 nm (3). (b) For the same samples ND, experimental transmission spectra at oblique incidence (α = 2°) and a calculated one for D = 405 nm (solid line). Eigen modes of sample ND with D = 405 nm are noted: λWG and λLGPR are bright modes and D1, D2 and D3 are dark modes.
Fig. 5.
Fig. 5. (color online) Calculation model (a) and distribution of the |E|-field in the unit cell of sample ND: (b, c) bright and dark (d-f) eigenmodes. The instantaneous value is shown in pink and green vectors for the E- and H-fields.
Fig. 6.
Fig. 6. (а) Photoluminescence spectra of an SU-8–R101 layer on Ag (dashed line) and samples NH with D = 545 (line 1), 555 (2) and 575 nm (3). (b) For same sample NH with D = 545 nm, experimental and calculated spectra at normal and oblique incidence. Experimental spectra at oblique incidence did not reveal any dark mode resonances.
Fig. 7.
Fig. 7. For the selected samples: (a) photoluminescence decay, (b) a normalized fluorescence lifetime versus wavelength.

Equations (6)

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

λ L G P R = n E F F D ,
n E F F M G = ε h ( 1 + 2 f R e { ε i } ε h R e { ε i } + 2 ε h ) / ( 1 f R e { ε i } ε h R e { ε i } + 2 ε h ) .
h W G = λ W G 2 π n S U 8 2 n U V grade 2 ( π ρ + arctg ( n S U 8 n U V g r a d e ) χ n U V grade 2 n Air 2 n S U 8 2 n U V grade 2 )
k s p p = k x + l G x + m G y = 2 π λ R e ( ε A g ) R e ( ε d ) R e ( ε A g ) + R e ( ε d ) ,
| k s p p | 2 = ( 2 π λ s i n θ + l 2 π D ) 2 + ( m 2 π D ) 2 ,
l 2 + m 2 n E F F 2 = D 2 λ 2 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.