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

Magnetic dipole emission in resonant metal-dielectric-metal structures

Open Access Open Access

Abstract

We explore a possibility to control magnetic dipole emission with plasmonic cavities, placing Eu3+ emitters inside profile-modulated metal-dielectric-metal structures. Significant variations in the branching ratio of the magnetic and electric dipole transitions are observed as the function of the thickness of the intermediate layer. The experimental results are confirmed with numerical simulations which account for cavity and gap plasmon resonances and predict modifications in the spontaneous emission spectrum as the function of the gap size and a strong directionality of the emission for small thicknesses of the intermediate layer. The implications of having a competition between electric and magnetic dipole relaxation channels in Eu3+ are discussed.

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

1. Introduction

In the last two decades, plasmonic structures and metamaterials have attracted the attention due to their novel applications and unconventional physics. Strong modification of photonic modes achieved with these systems makes it possible to control light propagation in unprecedented ways as featured in experimental demonstrations such as optical cloaking, negative refraction, and subwavelength imaging below diffraction limit [17]. In addition, it has been predicted and experimentally proved that strongly modified local optical environment provides new possibilities to control quantum processes [5,6]. Among them, the most known phenomenon is an enhancement of the spontaneous emission due to an increased density of photonic modes (so-called Purcell effect, first shown by Purcell in 1946 using the resonant cavity [8]). Multiple demonstrations and applications include use of various resonant structures, such as cavities, plasmonic nanoparticles, nano-antennas, metasurfaces and metamaterials with enhanced densities of photonic modes [917], including hyperbolic metamaterials having a nominally infinite density of photonic modes [16,17].

Majority of the theoretical studies and experiments are primarily focused on electric dipole emitters. Similar arguments have been also applied to magnetic dipole transitions which are commonly weak in natural materials but can be enhanced and controlled via modified optical environment (magnetic Purcell effect [1829]), providing a new playground for fundamental physics and various applications. Various approaches have been taken in order to achieve this, using the structures which provide strong enhancement of the magnetic component of photonic modes, such as plasmonic nanostructures having magnetic resonance at the emission wavelength [20,21], films, metasurfaces, metamaterials [2227], and high-index dielectric nanostructures [28,29]. However, in the contrary to expectations, experimentally observed enhancements in the magnetic dipole emission are not significant. This can be due to a relatively low quality-factor of resonances in real nanostructures, high loss in metal, or other factors such as a competition between various channels of the relaxation, which are also affected by the environment.

In this work, we test another approach for controlling magnetic dipole emission. We use tri-layer sandwich-like structures consisting of two external metal layers and a dielectric layer between them, which can exhibit the resonances at the emission wavelength. The resonance behavior of such structures is determined by several parameters, including the distance between the metal layers, d, thickness of metal films, dielectric constants of the materials and surrounding dielectrics, with the dispersion relation split to several branches [30]. Here we consider the low frequency range and two types of the resonances, a cavity (Fabry-Perot-type) resonance, Fig. 1 (a), and a low-frequency branch of gap plasmon resonances, Fig. 1(b). In terms of waveguide applications, these modes can be also referred as asymmetric and symmetric modes (regarding to the orientation of fields on interfaces) [31,32]. At relatively large d, both types of resonances occur, while at small gap sizes, $d < \lambda /2,\; $ only the gap plasmon modes survive with strongly confined fields. Since that gap plasmons allow one to achieve strong field confinement, they are of great practical interest for various studies and applications in information transfer technologies, sensing and single-photon sources, nanoantennas, quantum computing, etc. [33,34].

 figure: Fig. 1.

Fig. 1. (a,b) Schematics of fields in (a) cavity resonance and (b) gap plasmon modes in metal-air-metal structures at λ = 600 nm, p-polarization. Electric fields are shown with arrows, magnetic fields are shown with color (in arb. u.). The gap size is 270 nm. The dielectric constant of the substrate is assumed to be 3.2.

Download Full Size | PDF

The cavity and gap plasmon resonances have different distributions of magnetic and electric fields as illustrated in Fig. 1. We expect the emitters located inside to be affected by the resonant modes. Due to different distributions of the fields, the effects on electric and magnetic emitters can be different.

The cavity modes are decoupled to propagating modes at the condition

$$m\lambda = \; 2d\; n{\; }cos\theta $$
(neglecting penetration of the field in metal), where $\lambda $ is the wavelength in vacuum, $\theta $ is the angle of incidence, n is the refractive index, and m is an integer. The plasmonic mode is not expected to be decoupled without a special setup for the matching condition between the optical wavevector and gap plasmon wavevector, kGP, described with the relation [32],
$$\textrm{tanh}(d/2) = \frac{{\epsilon _d k_z^{(m )}}}{{\epsilon _d k_z^{(d )}}}\textrm{ } - \; ,\textrm{ with}\,k_z^{({m,d} )} = \; \,\sqrt {k_{GP}^2 - {\epsilon _{m,d}}k_0^2} ,$$
where $\epsilon _d $ and ɛm are the permittivities of dielectric and metal respectively, $k_z^{(m )}$ and $k_z^{(d )}$ are z-components of the wavevector for metal and dielectric respectively and k0 is the free space wavevector. At large gap sizes, kGP is comparable with the wavevector of the surface plasmon polariton (SPP) at a single metal-dielectric interface, kSPP and rapidly grows upon a decrease in the gap size. The matching condition can be achieved with the use of a high-index material, a grating or due to sharp surface features [35,36]. In our experiment we use grating-like structures where both external metal layers have one dimensional surface profile modulation with the subwavelength periodicity.

2. Experimental

Eu(NO3)3.Bpy2 is a convenient luminescent material for such an experiment, and has been used in the previous studies [17,20,2225,28] for probing effects of a modified optical environment on both electric and magnetic dipoles. Schematics of the energy levels and excitation and emission in this material are shown in Fig. 2(a). Eu3+ can be excited via ligand in the ultraviolet range and demonstrate bright emission in red (Fig. 2(b)) with several well-resolved peaks in the spectrum corresponding to the transitions from 5D0 to 7Fi sublevels. The predominantly magnetic dipole (MD) transition 5D0$\to $7F1 is observed around 590 nm, and the strongest 5D0$\to $7F2 transition associated primarily with the electric dipole (ED), is observed at 613 nm, Fig. 2(c).

 figure: Fig. 2.

Fig. 2. (a) Energy level diagram; (b) Photo of silver substrates covered with 25 nm-thick emitter layers under UV light illumination; (c) Spontaneous emission spectra of the film on the silver surface; (d) Schematics of the experimental profile-modulated structure.

Download Full Size | PDF

Eu(NO3)3.Bpy2 crystals are synthetized using the solution grown technique [22,37]. For fabrication of thin films, ethanol-water solution is prepared with the mixture of Eu(NO3)3.Bpy2 and polyvinylpyrrolidone (PVP) in 1:1 weight ratio.

As profile-modulated substrates we use polycarbonate substrates with the modulation period of 320 nm and height of 20-25 nm derived from digital storage BluRay-R (BR) discs (from Verbatium) following the procedure described in [38]. Silver is deposited with the thermal evaporation method. The first external silver layer is chosen to be thick enough (160 nm) to efficiently reflect the light, and the second silver layer is thin (30 nm) in order to provide efficient excitation of the emitters. Intermediate layers with Eu are fabricated using the spin coating technique. By regulating rotation speed, time and steps, we are able to produce the films with various thicknesses. The quality of the luminescent films is tested before the deposition of the second silver layer observing their luminescence under a UV lamp, Fig. 2(b). Planar structures on glass substrates are fabricated together with the profile-modulated structures; they are used as witness samples for the thickness measurement after each fabrication step. The thickness of the produced layer is measured in the flat structures with DekTekXT surface profilometer. Since particular layers in a profile-modulated structure and a witness sample are produced at the same thermal evaporation run or at the same spin coating parameters, we assume that they have the same thickness.

The schematics of fabricated three-layer systems is illustrated in Fig. 2(d). The silver layers have the same thicknesses for all systems in the series while the size of the emitter layer, d (in the range from 25 nm to 200 nm) is made different for each sample. Visual examination of the produced three-layer systems (under UV light) shows that in flat structures the emission is very weak, however it is significantly brighter in the profile-modulated systems, in particular, in the samples with small d and at large observation angles, where it is predominantly p-polarized.

In the experiment, the profile-modulated structures with various gap sizes are studied using the standard spectrofluorometer (Fluorolog-QM). The samples are placed on the stage with the grooves oriented vertically. The emission spectra are collected in the front-face collection setup, where the angle between the incident light (excitation) and collected light (emission) is 22.5°. We use the incidence angle, θ = 20° for the incident light. The excitation wavelength is 330 nm, which corresponds to the maximum of the excitation band. The emission spectra are recorded in the direction close to normal (2.5°) in the range of 570 nm - 630 nm, which covers both magnetic dipole transition and the strongest electric dipole transition, Fig. 3(a). We use the same orientation (θ = 20°) for all the systems under study and estimate the ratio of the peak intensities of magnetic dipole and electric dipole transitions, IMD / IED as the function of d. Variation in the angle θ does not significantly affect the ratio for a particular system (but affects the intensity of the signal).

 figure: Fig. 3.

Fig. 3. (a) Emission spectrum in silver-emitter layer-silver structures with d = 150 nm (blue) and 30 nm (red). Branching ratio of MD and ED transitions in (b) three-layer structures, (c) in the presence of the original BR multilayer. (d) Emission spectra on top of silver with the multilayer. Insets show schematics of the corresponding structures. The materials are indicated.

Download Full Size | PDF

The results obtained in the samples with different emitter layer thicknesses are shown in Fig. 3(b). Here, the ratio IMD /IED is normalized to that observed in the emitter films deposited on glass (0.18). The experiment demonstrates the minimum at intermediate d (∼ 150 nm) and an increase in IMD / IED toward a decrease in d. The magnitude IMD / IED do not exceed the ratio observed on glass at any gap thickness.

In addition to these results, we would like to mention an interesting experimental finding. If in the fabrication, we do not remove the original multilayer from BR discs and deposit the luminescent material directly on it, the ratio IMD / IED in such structures is significantly higher than that in our main experimental three-layer structures. It also shows different dependence on the gap size, Fig. 3(c). The ratio is even higher in the structure without the external silver layer, Fig. 3(d). The MD line exceeds the strongest ED line, which was, according to our knowledge, never observed before in or in vicinity of any plasmonic structure. (Typical spectrum of Eu3+ on top of silver is shown in Fig, 2(c).)

3. Analysis and discussion

Let us first discuss resonance properties of profile-modulated three-layer structures. We use COMSOL Multiphysics for numerical simulations of the optical fields in three-layered structures with one dimensional (1-D) modulation of profile, corresponding to the geometry of our experimental structures, Fig. 2(d). Since we concentrate on the emission range between 580 nm - 620 nm, the simulations are performed at the wavelength λ = 600 nm which is close to both MD and ED transitions. As a metal we use silver with the dielectric permittivity, εM = - 16 + 0.44i at 600 nm. The permittivity of the dielectric layer εD = 2.56 corresponds to that of the organic material in our experimental systems.

Note that the resonance behavior of this system having a profile modulation and a relatively thin metal layer is more complex than that in the gap between two semi-infinite flat metals. However, both cavity resonance (Fig. 4(a)) and gap plasmon (Fig. 4(b)) resonances are clearly present. In our geometry (p < λ), the gap plasmons can be excited (or decoupled) at the condition,

$${k_{GP\; }} = \; \frac{{2\pi }}{p}\; - \; \frac{{2\pi }}{\lambda }\; sin\theta .$$

 figure: Fig. 4.

Fig. 4. (a,b) Distributions of electric (top) and magnetic(bottom) energy in (a) cavity resonance mode at d = 160 nm, and (b) plasmonic mode, d = 75nm. (c) Reflectivity vs incidence angle at d = 25 nm.

Download Full Size | PDF

At large gaps,

$${k_{GP\; }} \approx \; {k_{SPP\; }}\; = \frac{{2\pi }}{\lambda }\; \sqrt {\frac{{{\epsilon _M}{\epsilon _D}}}{{{\epsilon _M}\; + \; {\epsilon _D}}}} \; $$

The Eq. (4) predicts the excitation or decoupling of plasmons at the incidence angle of ∼ 56°. Simulations show that this estimation holds up to the smallest thicknesses of the intermediate layer (d = 25 nm) used in our experiments. In the reflectivity, Fig. 4(c), this is observed as a deep minimum, indicating that our geometry provides conditions for efficient plasmon excitation or decoupling.

Purcell effect [8] predicts the modification of the spontaneous emission rate as

$$\; \tilde{\tau }_r^{ - 1}\; = F\tau _r^{ - 1},$$
where $\tilde{\tau }_r^{ - 1}$ and $\tau _r^{ - 1}$ are the modified and free space transition rate correspondingly, and F is the Purcell factor describing the change in the local density of states at the frequency of the transition. In the plasmonic environment, it includes participation of all possible modes: propagating far-field, resonance and lossy high-k plasmonic modes [39,40]. The change in the transition rate leads to the change in the probability of the particular channel of the relaxation, and corresponding enhancement, G,
$$\; G\; = \frac{{\; F\tau _r^{ - 1}/\; ({\; F\tau_r^{ - 1} + \; \tau_n^{ - 1}} ){\; }}}{{\; \tau _r^{ - 1}/\; ({\; \tau_r^{ - 1} + \; \tau_n^{ - 1}} )}} = \; \; F\frac{{\; \tau _r^{ - 1} + \; \tau _n^{ - 1}}}{{\; F\tau _r^{ - 1} + \; \tau _n^{ - 1}}}.$$

Here $\; \tau _n^{ - 1}$ is the rate for other relaxation channels which are not affected, including non-radiative relaxation. If the affected transition is very weak, $\; \tau _r^{ - 1}$ and $F\; \tau _r^{ - 1}$, are much lower than $\; \tau _n^{ - 1},\; \,G \approx \; F$, and the emission can become F -fold brighter (neglecting losses in metal).

Note in organic systems with Eu3+, the luminescence of Eu3+ is already very efficient [22,37,41,42]. Depending on the excitation pathways, it can reach up to 50% of quantum yield [42] mostly in ED transitions. Thus, in the very best scenario and extremely high Purcell factor, visible enhancement of the ED emission can be only around two. On the other hand, MD transition is relatively weak (of about 8-18% of the total emission depending on a particular material), leaving a room for a possible enhancement. However, since both ED and MD transitions originate from the same energy level, ED channels of the relaxation may become the main competitors for MD channels.

$$\; {G_{ED/MD}}\; = \; {F_{ED/MD}}\frac{{\; \; \tau _{ED}^{ - 1} + \tau _{MD}^{ - 1} + \; \tau _n^{ - 1}}}{{\; {F_{ED}}\tau _{ED}^{ - 1} + \; {F_{MD}}\tau _{MD}^{ - 1} + \; \tau _n^{ - 1}}}, $$
where ${G_{ED/MD}},\,{F_{ED/MD}}\; $ and $\tau _{ED/MD}^{ - 1}$ are correspondingly the intensity enhancement, Purcell factor and the rate for the electric/magnetic dipole transition in free space. In plasmonic systems electric dipoles can be stronger coupled with propagating plasmons, particularly at short distances from metal [24,25]. In this case, the effect of the environment results primarily in the change in the branching ratio between MD and ED transitions, without any significant enhancement of the emission intensity. For the illustration purpose, let us assume the following branching ratio, $\; \tau _{ED}^{ - 1}\; \; :\; \tau _{ED}^{ - 1}\; \; :\; \tau _n^{ - 1}$ = 0.4 : 0.1 : 0.5, and the Purcell factors as $\; {F_{MD}} = \; 10,\; \,{F_{ED}} = \; 20.\; $ Substituting the numbers in Eq. (7), ${G_{ED}} = 2.1$ and ${G_{MD}} = 1.05$.

Let us estimate the change in the branching ratio in our systems following the approach used in [28], involving the Lorentz reciprocity theorem [40,43]. According to the theorem [44,45], the light out-coupling problem can be solved considering light in-coming to the emitter. In our estimations, we consider only probabilities of the emission to propagating waves. Since both ED and MD transitions originate at the same level, contributions of non-propagating modes affect the total relaxation rate and result only in reducing the total emission intensity.

In calculations, we assume the random orientation of the emitters and, average the results over 32 points which are distributed over the modulation period (Fig. 5(a)), since the fields strongly vary inside the gap. The probability of the emission under the angle θ is estimated as $P_\theta ^{ED} \propto \; \mathop \sum \limits_{,TE,TM} {|{E(\theta )} |^2}$ and $P_\theta ^{MD} \propto \; \mathop \sum \limits_{,TE,TM}{|{H\; (\theta )} |^2}$ for electric and magnetic dipole emitters respectively, where E and H are correspondingly electric and magnetic components of the optical modes. Taking into account symmetry of our system and assuming no transmission through the thick silver layer, we vary the incidence angle θ between 0 and 90° and calculate the transitions probabilities in the emitter layer for various sizes of the gap, d = 25 - 280 nm. At a small gap size, the emission is determined by gap plasmons. It is highly directional, showing significant enhancement (more than order of magnitude in comparison with emitters in free space) in the direction corresponding to the plasmon decoupling angle, Fig. 5(b). This is in line with the experimentally observed predominantly p-polarized emission at large observation angles. At larger d, the pattern is more complicated and much broader covering the whole angular range.

 figure: Fig. 5.

Fig. 5. (a) Positions of emitters (shown with red dots) in calculations of the emitting properties; (b) Angular dependence of the ED (blue) and MD (red) emission intensity in the structure with d = 25 nm; (c) Branching MD/ED ratio, calculations (circles) and experiment (crosses).

Download Full Size | PDF

The total MD or ED emission to free space is estimated with integration over θ . The resulting branching ratio, ${I_{MD}}/{I_{ED}}$ shows the variation with the gap size well corresponding to the experimental data, Fig. 5(c).

In our three-layer resonant structures both electric and magnetic dipoles are affected by plasmons or cavity modes; this restricts the opportunity to fully control or enhance magnetic dipole emission in Eu3+ ions. In principle, placing the emitters with nanoscale accuracy to certain places of standing resonance modes can provide for the enhancement, however, would require sophisticated fabrication approaches.

Commonly, in the research devoted to MD enhancement, the main attention is paid to the structures with a strong magnetic response. An alternative approach worth investigating can be in designing the structures primarily targeting decreasing the competitive ED transition rate. This would lead to a natural increase in the MD emission intensity and a better control of it. Our preliminary considerations show that this may be the cause for our surprising accidental finding of MD/ED brunching ratio > 1 in BR substrates with original recording multilayers, Fig. 3(d), and its behavior in complex cavities, Fig. 3(c). Since the composition of the BR discs is a proprietary information, for a proper study we plan to repeat these experiments with our own multilayers and publish in detail elsewhere.

4. Conclusions

In conclusion, profile-modulated three-layer systems consisting of two external silver layers and an intermediate layer with Eu3+ emitters are experimentally and theoretically studied, taking into account the competition of ED and MD channels of the relaxation. Strong directionality of the emission is predicted for small gap sizes; it is associated with decoupling of gap plasmons excited by emitters inside the cavity. Since both ED and MD emitters are affected, such structures do not provide for the efficient MD enhancement, and mainly cause the variation in the branching ratio of MD and ED transitions in the dependence of the emitter layer thickness.

Funding

National Nuclear Security Administration (NNSA) (DE-NA00040007); National Science Foundation (NSF) (#1830886, #2112595).

Disclosures

The authors declare no conflict of interest.

Data availability

Data supporting the results presented in this paper can be obtained from the authors upon reasonable request.

References

1. J. B. Pendry, “Negative refraction makes a perfect lens,” Phys. Rev. Lett. 85(18), 3966–3969 (2000). [CrossRef]  

2. V. M. Shalaev, “Optical negative-index metamaterials,” Nat. Photonics 1(1), 41–48 (2007). [CrossRef]  

3. W. Cai, U. K. Chettiar, A. V. Kildishev, and V. M. Shalaev, “Designs for optical cloaking with high-order transformations,” Opt. Express 16(8), 5444–5452 (2008). [CrossRef]  

4. Z. Jacob, L. V. Alekseyev, and E. Narimanov, “Optical Hyperlens: Far-field imaging beyond the diffraction limit,” Opt. Express 14(18), 8247–8256 (2006). [CrossRef]  

5. M. A. Noginov, “Metamaterials and Plasmonics with Optical Gain,” World Sci. Ser. Nanosci. Nanotechnol. 3153–191 (2017). [CrossRef]  

6. L. Novotny and B. Hecht, Principles of Nano-Optics (Cambridge University Press, 2006).

7. J. A. Schuller, E. S. Barnard, W. Cai, Y. C. Jun, J. S. White, and M. L. Brongersma, “Plasmonics for extreme light concentration and manipulation,” Nat. Mater. 9(3), 193–204 (2010). [CrossRef]  

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

9. W. L. Barnes, “Topical Review. Fluorescence near interfaces. The role of photonic mode density,” J. Mod. Opt. 45(4), 661–699 (1998). [CrossRef]  

10. E. F. Schubert, A. M. Vredenberg, N. E. J. Hunt, Y. H. Wong, P. C. Becker, J. M. Poate, D. C. Jacobson, L. C. Feldman, and G. J. Zydzik, “Giant enhancement of luminescence intensity in Er-doped Si/SQ resonant cavities,” Appl. Phys. Lett. 61(12), 1381–1383 (1992). [CrossRef]  

11. A. M. Kern, D. Zhang, M. Brecht, A. I. Chizhik, A. V. Failla, F. Wackenhut, and A. J. Meixner, “Enhanced single-molecule spectroscopy in highly confined optical fields: from λ/2-Fabry–Pérot resonators to plasmonic nano-antennas,” Chem. Soc. Rev. 43(4), 1263–1286 (2014). [CrossRef]  

12. N. P. d. Leon, B. J. Shields, C. L. Yu, D. E. Englund, A. V. Akimov, M. D. Lukin, and H. Park, “Tailoring Light-Matter Interaction with a Nanoscale Plasmon Resonator,” Phys. Rev. Lett. 108(22), 226803 (2012). [CrossRef]  

13. J. E. Park, J. Kima, and J. M. Nam, “Emerging plasmonic nanostructures for controlling and enhancing photoluminescence,” Chem. Sci. 8(7), 4696–4704 (2017). [CrossRef]  

14. R. Collison, J. B. Pérez-Sánchez, M. Du, J. Trevino, J. Yuen-Zhou, S. O’Brien, and V. M. Menon, “Purcell Effect of Plasmonic Surface Lattice Resonances and Its Influence on Energy Transfer,” ACS Photonics 8(8), 2211–2219 (2021). [CrossRef]  

15. F. Yang, P. K. Jha, H. Akbari, H. C. Bauser, and H. A. Atwater, “A hybrid coupler for directing quantum light emission with high radiative Purcell enhancement to a dielectric metasurface lens,” J. Appl. Phys. 130(16), 163103 (2021). [CrossRef]  

16. L. Ferrari, D. Lu, D. Lepage, and Z. Liu, “Enhanced spontaneous emission inside hyperbolic metamaterials,” Opt. Express 22(4), 4301–4306 (2014). [CrossRef]  

17. M. A. Noginov, H. Li, Y. A. Barnakov, D. Dryden, G. Nataraj, G. Zhu, C. E. Bonner, M. Mayy, Z. Jacob, and E. E. Narimanov, “Controlling spontaneous emission with metamaterials,” Opt. Lett. 35(11), 1863–1865 (2010). [CrossRef]  

18. H. W. Wu, Z. Q. Sheng, H. Jing, R. H. Fan, and R. W. Peng, “Strong Purcell Effect for Terahertz Magnetic Dipole Emission with Spoof Plasmonic Structure,” ACS Appl. Nano Mater. 2(2), 1045–1052 (2019). [CrossRef]  

19. T. Feng, W. Zhang, Z. Liang, Y. Xu, and E. M. Andrey, “Isotropic Magnetic Purcell Effect,” ACS Photonics 5(3), 678–683 (2018). [CrossRef]  

20. R. Hussain, S. S. Kruk, C. E. Bonner, M. A. Noginov, I. Staude, Y. S. Kivshar, N. Noginova, and D. N. Neshev, “Enhancing Eu3 + magnetic dipole emission by resonant plasmonic nanostructures,” Opt. Lett. 40(8), 1659–1662 (2015). [CrossRef]  

21. H. W. Wu, J. Q. Quan, Y. Q. Yin, and Z. Q. Sheng, “Strong Purcell effect for magnetic dipole emission with spoof plasmonic spiral structure,” J. Opt. Soc. Am. B 37(1), 98–103 (2020). [CrossRef]  

22. N. Noginova, Y. Barnakov, H. Li, and M. A. Noginov, “Effect of metallic surface on electric dipole and magnetic dipole emission transitions in Eu3+ doped polymeric film,” Opt. Express 17(13), 10767–10772 (2009). [CrossRef]  

23. N. Noginova, R. Hussain, M. A. Noginov, J. Vella, and A. Urbas, “Modification of electric and magnetic dipole emission in anisotropic plasmonic systems,” Opt. Express 21(20), 23087–23096 (2013). [CrossRef]  

24. R. Hussain, D. Keene, N. Noginova, and M. Durach, “Spontaneous emission of electric and magnetic dipoles in the vicinity of thin and thick metal,” Opt. Express 22(7), 7744–7755 (2014). [CrossRef]  

25. S. Mashhadi, M. Durach, D. Keene, and N. Noginova, “Control of magnetic dipole emission with surface plasmon polaritons,” OSA Continuum 2(4), 1342–1349 (2019). [CrossRef]  

26. A. P. Slobozhanyuk, A. N. Poddubny, A. E. Krasnok, and P. A. Belov, “Magnetic Purcell factor in wire metamaterials,” Appl. Phys. Lett. 104(16), 161105 (2014). [CrossRef]  

27. D. G. Baranov, R. S. Savelev, S. V. Li, A. E. Krasnok, and A. Alù, “Modifying magnetic dipole spontaneous emission with nanophotonic structures,” Laser Photonics Rev. 11(3), 1600268 (2017). [CrossRef]  

28. A. Vaskin, S. Mashhadi, M. Steinert, K. E. Chong, D. Keene, S. Nanz, A. Abass, E. Rusak, D. Choi, I. Fernandez-Corbaton, T. Pertsch, C. Rockstuhl, M. A. Noginov, Y. S. Kivshar, D. N. Neshev, N. Noginova, and I. Staude, “Manipulation of Magnetic Dipole Emission from Eu3+ with Mie- Resonant Dielectric Metasurfaces,” Nano Lett. 19(2), 1015–1022 (2019). [CrossRef]  

29. Y. Yang, B. Zhu, H. Dai, and X. Sun, “Identical emission enhancement for arbitrary-orientation magnetic dipole emitters in silicon hollow nanocavity,” Opt. Express 27(18), 25931–25942 (2019). [CrossRef]  

30. E. N. Economu, “Surface plasmons in thin films,” Phys. Rev. 182(2), 539–554 (1969). [CrossRef]  

31. R. D. Kekatpure, A. C. Hryciw, E. S. Barnard, and M. L. Brongersma, “Solving dielectric and plasmonic waveguide dispersion relations on a pocket calculator,” Opt. Express 17(26), 24112–24129 (2009). [CrossRef]  

32. S. I. Bozhevolnyi, “Introduction to surface plasmon-polariton waveguides,” In Plasmonic Nanoguides and Circuits. Pan Stanford Publishing Pte. Ltd. (2009).

33. J. Jung and T. Sondergaard, “Gap plasmon-polariton nanoresonators: Scattering enhancement and launching of surface plasmon polaritons,” Phys. Rev. B 79(3), 035401 (2009). [CrossRef]  

34. F. Ding, Y. Yang, R. A. Deshpande, and S. I. Bozhevolnyi, “A review of gap-surface plasmon metasurfaces: fundamentals and applications,” Nanophotonics 7(6), 1129–1156 (2018). [CrossRef]  

35. H . Raether,“Surface plasmons on smooth and rough surfaces and on gratings,” Springer Tracts iModern Phys. 111, (Springer: Berlin), (1988). [CrossRef]  

36. S. de la Cruz, E. R. Mendez, D. Macıas, R. Salas-Montiel, and P. M. Adam, “Compact surface structures for the efficient excitation of surface plasmon-polaritons,” Phys. Status Solidi B 249(6), 1178–1187 (2012). [CrossRef]  

37. V. I. Tsaryuk, K. P. Zhuravlev, V. F. Zolin, V. A. Kudryashova, J. Legendziewicz, and R. Szostak, “Luminescence efficiency of aromatic carboxylates of europium and terbium when methylene bridges and nitro groups are present in the ligands,” J. Appl. Spectrosc. 74(1), 51–59 (2007). [CrossRef]  

38. D. Keene, T. Ronurpraful, and N. Nogniova, “Plasmon related electrical effects in strongly modulated metasurfaces,” Proc. SPIE 11080, 45 (2019). [CrossRef]  

39. R. Salas-Montiel, M. Berthel, J. Beltran-Madrigal, S. Huant, A. Drezet, and S. Blaize, “Local density of electromagnetic states in plasmonic nanotapers: spatial resolution limits with nitrogen-vacancy centers in diamond nanospheres,” Nanotechnology 28(20), 205207 (2017). [CrossRef]  

40. A. A. Meza-Olivo, K. Garay-Palmett, S. Blaize, and R. Salas-Montiel, “Reciprocity and Babinet's principles applied to the enhancement of the electric and magnetic local density of states in integrated plasmonics on silicon photonics,” Appl. Opt. 57(30), 9155–9163 (2018). [CrossRef]  

41. Y. Kitagawa, M. Tsuruib, and Y. Hasegawa, “Bright red emission with high color purity from Eu(III) complexes with π-conjugated polycyclic aromatic ligands and their sensing applications,” RSC Adv. 12(2), 810–821 (2022). [CrossRef]  

42. C. Yang, L. Fu, Y. Wang, J. Zhang, W. Wong, X.. Ai, Y. Qiao, B. Zou, and L. Gui, “A Highly Luminescent Europium Complex Showing Visible-Light-Sensitized Red Emission: Direct Observation of the Singlet Pathway,” Angew. Chem. Int. Ed. 43(38), 5010–5013(2004). [CrossRef]  

43. S. Zhang, E. R. Martins, A. G. Diyaf, J. I. B. Wilson, G. A. Turnbull, and I. D. W. Samuel, “Calculation of the emission power distribution of microstructured OLEDs using the reciprocity theorem,” Synth. Met. 205, 127–133 (2015). [CrossRef]  

44. H. A. Lorentz, V. K. Akad, and W. Amsterdam, Collected Papers Vol III (Martinus Nijhoff, Hague), 4(176), 1–11(1896).

45. R. J. Potton, “Reciprocity in optics,” Rep. Prog. Phys. 67(5), 717–754 (2004). [CrossRef]  

Data availability

Data supporting the results presented in this paper can be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a,b) Schematics of fields in (a) cavity resonance and (b) gap plasmon modes in metal-air-metal structures at λ = 600 nm, p-polarization. Electric fields are shown with arrows, magnetic fields are shown with color (in arb. u.). The gap size is 270 nm. The dielectric constant of the substrate is assumed to be 3.2.
Fig. 2.
Fig. 2. (a) Energy level diagram; (b) Photo of silver substrates covered with 25 nm-thick emitter layers under UV light illumination; (c) Spontaneous emission spectra of the film on the silver surface; (d) Schematics of the experimental profile-modulated structure.
Fig. 3.
Fig. 3. (a) Emission spectrum in silver-emitter layer-silver structures with d = 150 nm (blue) and 30 nm (red). Branching ratio of MD and ED transitions in (b) three-layer structures, (c) in the presence of the original BR multilayer. (d) Emission spectra on top of silver with the multilayer. Insets show schematics of the corresponding structures. The materials are indicated.
Fig. 4.
Fig. 4. (a,b) Distributions of electric (top) and magnetic(bottom) energy in (a) cavity resonance mode at d = 160 nm, and (b) plasmonic mode, d = 75nm. (c) Reflectivity vs incidence angle at d = 25 nm.
Fig. 5.
Fig. 5. (a) Positions of emitters (shown with red dots) in calculations of the emitting properties; (b) Angular dependence of the ED (blue) and MD (red) emission intensity in the structure with d = 25 nm; (c) Branching MD/ED ratio, calculations (circles) and experiment (crosses).

Equations (7)

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

m λ = 2 d n c o s θ
tanh ( d / 2 ) = ϵ d k z ( m ) ϵ d k z ( d )   ,  with k z ( m , d ) = k G P 2 ϵ m , d k 0 2 ,
k G P = 2 π p 2 π λ s i n θ .
k G P k S P P = 2 π λ ϵ M ϵ D ϵ M + ϵ D
τ ~ r 1 = F τ r 1 ,
G = F τ r 1 / ( F τ r 1 + τ n 1 ) τ r 1 / ( τ r 1 + τ n 1 ) = F τ r 1 + τ n 1 F τ r 1 + τ n 1 .
G E D / M D = F E D / M D τ E D 1 + τ M D 1 + τ n 1 F E D τ E D 1 + F M D τ M D 1 + τ n 1 ,
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.