ArticlePDF Available

Designing dielectric resonators on substrates: Combining magnetic and electric resonances

Optica Publishing Group
Optics Express
Authors:

Abstract and Figures

High-performance integrated optics, solar cells, and sensors require nanoscale optical components at the surface of the device, in order to manipulate, redirect and concentrate light. High-index dielectric resonators provide the possibility to do this efficiently with low absorption losses. The resonances supported by dielectric resonators are both magnetic and electric in nature. Combined scattering from these two can be used for directional scattering. Most applications require strong coupling between the particles and the substrate in order to enhance the absorption in the substrate. However, the coupling with the substrate strongly influences the resonant behavior of the particles. Here, we systematically study the influence of particle geometry and dielectric environment on the resonant behavior of dielectric resonators in the visible to near-IR spectral range. We show the key role of retardation in the excitation of the magnetic dipole (MD) mode, as well as the limit where no MD mode is supported. Furthermore, we study the influence of particle diameter, shape and substrate index on the spectral position, width and overlap of the electric dipole (ED) and MD modes. Also, we show that the ED and MD mode can selectively be enhanced or suppressed using multi-layer substrates. And, by comparing dipole excitation and plane wave excitation, we study the influence of driving field on the scattering properties. Finally, we show that the directional radiation profiles of the ED and MD modes in resonators on a substrate are similar to those of point-dipoles close to a substrate. Altogether, this work is a guideline how to tune magnetic and electric resonances for specific applications.
This content is subject to copyright. Terms and conditions apply.
Designing dielectric resonators on
substrates: Combining magnetic and
electric resonances
J. van de Groepand A. Polman
Center for Nanophotonics, FOM Institute AMOLF, Science Park 104, 1098 XG, Amsterdam,
Netherlands
groep@amolf.nl
Abstract: High-performance integrated optics, solar cells, and sensors
require nanoscale optical components at the surface of the device, in order to
manipulate, redirect and concentrate light. High-index dielectric resonators
provide the possibility to do this efficiently with low absorption losses. The
resonances supported by dielectric resonators are both magnetic and electric
in nature. Combined scattering from these two can be used for directional
scattering. Most applications require strong coupling between the particles
and the substrate in order to enhance the absorption in the substrate.
However, the coupling with the substrate strongly influences the resonant
behavior of the particles. Here, we systematically study the influence of
particle geometry and dielectric environment on the resonant behavior of
dielectric resonators in the visible to near-IR spectral range. We show the
key role of retardation in the excitation of the magnetic dipole (MD) mode,
as well as the limit where no MD mode is supported. Furthermore, we study
the influence of particle diameter, shape and substrate index on the spectral
position, width and overlap of the electric dipole (ED) and MD modes.
Also, we show that the ED and MD mode can selectively be enhanced
or suppressed using multi-layer substrates. And, by comparing dipole
excitation and plane wave excitation, we study the influence of driving
field on the scattering properties. Finally, we show that the directional
radiation profiles of the ED and MD modes in resonators on a substrate are
similar to those of point-dipoles close to a substrate. Altogether, this work
is a guideline how to tune magnetic and electric resonances for specific
applications.
© 2013 Optical Society of America
OCIS codes: (140.4780) Optical resonators; (290.5850) Scattering, particles; (350.4238)
Nanophotonics and photonic crystals.
References and links
1. 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, 193–204 (2010).
2. C. H¨
oppener and L. Novotny, “Antenna-based optical imaging of single Ca2+transmembrane proteins in liquids,
Nano Lett. 8, 642–646 (2008).
3. H. A. Atwater and A. Polman, “Plasmonics for improved photovoltaic devices,” Nat. Mater. 9, 205–213 (2010).
4. V. E. Ferry, M. A. Verschuuren, H. B. T. Li, E. Verhagen, R. H. Walters, R. E. I. Schropp, H. A. Atwater, and A.
Polman, “Light trapping in ultrathin plasmonic solar cells,” Opt. Express 18, A237–A245 (2010).
5. M. K. Kwon, J. Y. Kim, B. H. Kim, I. K. Park, C. Y. Cho, C. C. Byeon, and S. J. Park, “Surface-plasmon-enhanced
light-emitting diodes,” Adv. Mater. 20, 1253–1257 (2008).
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26285
6. N. Yu, R. Blanchard, J. Fan, Q. J. Wang, C. P߬
ugl, L. Diehl, T. Edamura, M. Yamanishi, H. Kan, and F. Capasso,
“Quantum cascade lasers with integrated plasmonic antenna-array collimators,” Opt. Express 16, 19447–19461
(2008).
7. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (Wiley, 2008).
8. C. M. Soukoulis and M. Wegener, “Past achievements and future challenges in the development of three-
dimensional photonic metamaterials,” Nat. Photonics 5, 523–530 (2011).
9. Q. Zhao, J. Zhou, F. Zhang, and D. Lippens, “Mie resonance-based dielectric metamaterials, Mater. Today 12,
60–69 (2009).
10. L. Huang, Y. Yu, and L. Cao, “General modal properties of optical resonances in subwavelength nonspherical
dielectric structures,” Nano Lett. 13, 3559–3565 (2013).
11. Y. H. Fu, A. I. Kuznetsov, A. E. Miroschnichenko, Y. F. Yu, and B. Luk’yanchuk, “Directional visible light
scattering by silicon nanoparticles,” Nat. Commun. 4, 1527 (2013).
12. G. Mie, “Beitrge zur optik trber medien, speziell kolloidaler metallsungen,” Ann. Phys. 330, 377–445 (1908).
13. A. Garc´
ıa-Etxarri, R. G´
omez-Medina, L. S. Froufe-P´
erez, C. L´
opez, L. Chantada, F. Scheffold, J. Aizpurua, M.
Nieto-Vesperinas, and J. J. S´
aenz, “Strong magnetic response of submicron silicon particles in the infrared,” Opt.
Express 19, 4815–4826 (2011).
14. A. B. Evlyukhin, C. Reinhardt, A. Seidel, B. S. Luk’yanchuck, and B. N. Chichkov, “Optical response features
of Si-nanoparticle arrays,” Phys. Rev. B 82, 045404 (2010).
15. J. A. Schuller and M. L. Brongersma, “General properties of dielectric optical antennas,” Opt. Express 17, 24084–
24095 (2009).
16. A. I. Kuznetsov, A. E. Miroschnichenko, Y. H. Fu, J. B. Zhang, and B. Luk’yanchuk, “Magnetic light,” Sci. Rep.
2, 492 (2012).
17. A. B. Evlyukhin, C. Reinhardt, and B. N. Chichkov, “Multipole light scattering by nonspherical nanoparticles in
the discrete dipole approximation,” Phys. Rev. B 84, 235429 (2011).
18. A. B. Evlyukhin, S. M. Novikov, U. Zywietz, R. L. Eriksen, C. Reinhardt, S. I. Bozhevolnyi, and B. N. Chichkov,
“Demonstration of magnetic dipole resonances of dielectric nanospheres in the visible region,” Nano Lett. 12,
3749–3755 (2012).
19. M. K. Schmidt, R. Esteban, J. J. S´
aenz, I. Su´
arez-Lacelle, S. Mackowski, and J. Aizpurua, “Dielectric antennas -
a suitable platform for controlling magnetic dipolar emission,” Opt. Express 20, 13636–13650 (2012).
20. S. Person, M. Jain, Z. Lapin, J. J. S´
aenz, G. Wicks, and L. Novotny, “Demonstration of zero optical backscattering
from single nanoparticles,” Nano Lett. 13, 1806–1809 (2013).
21. I. Staude, A. E. Miroshnichenko, M. Decker, N. T. Fofang, S. Liu, E. Gonzales, J. Dominguez, T. S. Luk, D. N.
Neshev, I. Brener, and Y. Kivshar, “Tailoring directional scattering through magnetic and electric resonances in
subwavelength silicon nanodisks, ACS Nano 7, 7824–7832 (2013).
22. P. Fan, K. C. Y. Huang, L. Cao, and M. L. Brongersma, “Redesigning photodetector electrodes as an optical
antenna,” Nano Lett. 13, 392–396 (2013).
23. P. Spinelli, M. A. Verschuuren, and A. Polman, “Broadband omnidirectional antireflection coating based on
subwavelength surface Mie resonators, Nat. Commun. 3, 692 (2012).
24. P. Spinelli, B. Macco, M. A. Verschuuren, W. M. M. Kessels, and A. Polman, Al2O3/TiO2 nano-pattern antire-
flection coating with ultralow surface recombination, Appl. Phys. Lett. 102, 233902 (2013).
25. FDTD Solutions, Lumerical Solutions, Inc., http://www.lumerical.com.
26. E. D. Palik, Handbook of Optical Constants of Solids (Academic, 1985).
27. P. Spinelli, C. van Lare, E. Verhagen, and A. Polman, “Controlling Fano lineshapes in plasmon-mediated light
coupling into a substrate,” Opt. Express 19,A303–A311 (2011).
28. T. Coenen, J. van de Groep, and A. Polman, “Resonant Mie modes of single silicon nanocavities excited by
electron irradiation,” ACS Nano 7, 1689–1698 (2013).
29. F. J. Garc´
ıa de Abajo, “Optical excitations in electron microscopy, Rev. Mod. Phys. 82, 209–275 (2010).
30. R. Sapienza, T. Coenen, J. Renger, M. Kuttge, N. F. van Hulst, and A. Polman, “Deep-subwavelength imaging
of the modal dispersion of light,” Nat. Mater. 11, 781–787 (2012).
31. T. Coenen, E. J. R. Vesseur, and A. Polman, “Deep subwavelength spatial characterization of angular emission
from single-crystal Au plasmonic ridge nanoantennas,” ACS Nano 6, 1742–1750 (2012).
32. K. G. Lee, X. W. Chen, H. Eghlidi, P. Kukura, R. Lettow, A. Renn, V. Sandoghdar, and S. Gotzinger, “A planar
dielectric antenna for directional single-photon emission and near-unity collection efficiency,” Nat. Photonics 5,
166–169 (2011).
33. W. Lukosz and R. E. Kunz, “Light emission by magnetic and electric dipoles close to a plane interface. I. Total
radiated power, J. Opt. Soc. Am. 67, 1607–1614 (1977).
34. L. Novotny and B. Hecht, Principles of Nano-Optics (Cambridge University, 2006).
35. M. Kerker, D. Wang, and G. Giles, “Electromagnetic scattering by magnetic spheres,” J. Opt. Soc. Am. 73,
765–767 (1983).
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26286
1. Introduction
High-performance integrated optics, solar cells, sensors and many other optical devices
require the ability to concentrate, redirect and manipulate light at the surface of the device, in
order to optimize the coupling and trapping of light. Resonant nanostructures can have large
scattering cross sections at visible wavelengths, and therefore are suitable candidates for this
purpose. Plasmonic nanostructures have proven to be very successful in this field, and have
been used for light focusing and scattering in sensing and imaging [1, 2], solar cells [3, 4]
and light emitting devices [5, 6]. However, the performance of plasmonic nanostructures is
always limited by parasitic absorption in the metal. High-index dielectric or semiconductor
nanostructures support (Mie-like) geometrical resonances in the visible spectral range [7–10].
Unlike in plasmonic particles, these are driven by displacement currents rather than actual
currents [11]. Therefore, these resonances are characterized by (very) low losses. Scattering
from spherical dielectric particles in a homogeneous medium has already been described in
great detail using Mie theory [12–15], and particles weakly coupled to the substrate have also
been analyzed [16]. This work has shown that silicon nanoparticles support both electrical and
magnetic resonances in the visible and near-infrared spectral range [13,14, 16–19]. Recently,
it has also been shown that the interplay between electrical and magnetic resonances can be
used to control the angular radiation pattern of the resonant cavities [11,20,21], and that using
cylinders instead of spheres allows for tuning of the spectral spacing between the electric and
magnetic lowest order resonances [17,21]. For all practical applications particles are placed on
a substrate and in most cases strong coupling to the substrate is even desired (e.g. absorption
in photodiodes [22], solar cells [23, 24], etc.). This interaction with the substrate strongly
influences the resonant behavior, as well as the interaction between the magnetic and electric
resonances. So far, there has been no systematic study on the influence of particle geometry
and dielectric environment on the resonant properties of dielectric nanoparticles on a substrate.
Here, we use numerical simulations to systematically study the resonant properties of high-
index dielectric particles in the visible spectral range for different particle geometries and di-
electric environments. We show how electric and magnetic resonances are supported, and that
retardation plays a key role in the driving of the magnetic resonances. Also, we show how
varying the height and diameter of resonant nanocylinders placed on a dielectric substrate pro-
vides the possibility to spectrally tune the electric dipole (ED) and magnetic dipole (MD) reso-
nances. Furthermore, we study how the presence of a (high-index) substrate dramatically alters
the resonant properties, as well as the interplay between the different eigenmodes. We show
that by tuning the substrate geometry, the ED and MD resonances can be selectively enhanced
or suppressed. Also, we find that such resonators can be used as low-loss antennas that exhibit
directional emission into the substrate. Altogether, this work shows how, by engineering the
particle shape and dielectric environment, the scattering properties of single resonators placed
on a dielectric substrate can be tailored for specific applications.
2. Magnetic and electric modes in dielectric particles
The ability to excite magnetic (dipolar) resonances in silicon particles, as well as the driving
mechanism, has been described recently [13, 16, 18]. In short, the in-plane magnetic dipole
is driven by the electric field of light that couples to displacement current loops, vertically
oriented in the particle (see inset in Fig. 1(a)). This displacement current loop (orange) induces
a magnetic dipole moment (blue), oriented perpendicular to the electric field polarization.
Efficient driving of this displacement current loop requires significant retardation of the driving
field (red) throughout the particle, as the electric field should undergo a significant phase
shift in order to match the opposing electric field orientation in the top and bottom part of
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26287
the particle. The excitation of electric (dipolar) resonances on the other hand, only requires
collective polarization of the material inside the resonator by the electric field component of
the incident light.
Here, we systematically study the nature of the resonances supported by (mostly cylindrical)
nanoparticles while changing the particle height, diameter, shape, presence and refractive index
of a substrate, thickness of multi-layered substrates and driving field. To this end, we use finite-
difference time-domain simulations [25] to calculate the normalized scattering cross section
(scattering cross section normalized to geometrical cross section) Qscat =
σ
scat/
σ
geo and modal
field profiles inside the particles. A total-field scattered-field source (TFSF) source is used,
which launches a broad-band (
λ
=400 1000 nm) plane wave from the top under normal
incidence and filters out all the light that has not been scattered. Power transmission monitors
are positioned around the TFSF source to monitor the power of the scattered field, and field
monitors are positioned in and around the particle to monitor the local field intensity. Substrates
(when used) have a purely real refractive index nsub, to prevent absorption of the scattered light
before reaching the power monitors. Perfectly Matching Layers (PMLs) are used to prevent any
unphysical scattering from the simulation boundaries, and to mimic semi-infinite substrates.
Optical constants for Si and SiO2are taken from Palik [26]. Automatic non-uniform meshes
are used, and a 2.5 nm refinement mesh around the particle when monitoring the near-field
intensities. Identification of the modes (MD, ED) is done by studying the field profiles inside
the particle and identifying the corresponding current loops.
3. Retardation and particle height
We start by investigating the role of particle height on the excitation of resonant modes
in crystalline Si cylinders in a homogeneous air environment. We choose Si since this is
a high-index material with relatively low absorption losses in the visible and near-IR, and
has been used in most previous work on magnetic resonances [11, 13, 16, 18, 21, 23]. The
necessity of field retardation for the excitation of the MD mode suggests that if the particle is
shallow enough, there should not be enough retardation to drive a displacement current loop,
and therefore a MD resonance cannot be supported. The ED mode however, does not need
retardation inside the particle: since
μ
=1 both inside and outside the particle, the magnetic
current loop induced by the dipole moment can also extend outside the particle.
To study the effect of particle height on retardation, we calculate Qscat for a Si cylinder in
air, with diameter d=100 nm and height h=100 nm. Figure 1(a) shows Qscat as a function
of wavelength. Two clear peaks are observed which correspond to the ED mode at
λ
=440
nm and the MD mode at
λ
=503 nm. As mentioned above, identification of the modes (MD,
ED) is done by studying the field profiles inside the particle and identifying the corresponding
current loops. Figure 1(b) shows Qscat spectra for particle heights in the range h=50 250
nm. The vertical dashed line represents the crosscut shown in Fig. 1(a). The two maxima,
labeled ED and MD, show a clear red-shift and increase in Qscat for increasing h. Note that
two peaks can be observed for the entire range of h, even down to 50 nm. Thus, even a cylinder
only 50 nm high supports a vertically-oriented displacement current loop generating a MD
resonance. The strong increase in Qscat for larger height is due to the fact that the volume of
the particle increases and thus the polarizability and thereby
σ
scat, while the geometrical cross
section
σ
geo does not change. Unlike plasmonic particles, where an increase in hgives rise to
a small blue-shift for an in-plane dipole mode due to an increased restoring force, higher Si
cylinders give rise to a large red-shift that saturates above h220 nm. This can be understood
from Figs. 1(c)–1(e), which show vertical crosscuts parallel to the polarization of the driving
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26288
50 nm
e)
λ (nm)
50 100 150 200 250
400
450
500
550
600
650
2
6
10
14
18
Qscat
Si
100 nm
h
normalized |E|2
h=50 nm
height (nm)
0
1
h=100 nm h=175 nm
ED
MD
b)
E
H
k
normalized |E|2normalized |E|2
400 450 500 550 600 650 700
0
1
2
3
4
5
6
7
8
λ (nm)
Qscat
ED
MD
a)
m
E
p
H
0
1
0
1d)
50 nm
50 nm
c)
MD excitation
m
H
E
k
Fig. 1. (a) Qscat as a function of wavelength for a Si cylinder in air with h=100 nm and
d=100 nm for optical excitation under normal incidence. The peaks correspond to the ED
and MD mode, and the corresponding dipole moments and current loops are shown. The
polarization of the driving field is shown in the top right. The inset shows the excitation
mechanism of the MD mode. (b) Qscat (color) as a function of
λ
and hfor h=50250
nm. The particle geometry is shown as an inset. The vertical white line is the crosscut
corresponding to the spectrum shown in (a). The white dots correspond to the
λ
-hcom-
binations used for (c-e). (c-e) Normalized electric field intensity |E|2(color) and electric
field lines (gray) in vertical crosscuts through particles with h=50 nm (c), h=100 nm (d)
and h=175 nm (e), parallel to the electric driving field.
field of the normalized electric field intensity distribution |E|2(color) for h=50 nm (c),
h=100 nm (d) and h=175 nm (e). Electric field lines (gray) in this plane at resonance are
also plotted. The three combinations of hand
λ
from Figs. 1(c)–1(e) are indicated by white
dots in Fig. 1(b). In Figs. 1(d) and 1(e), both the intensity profiles and the field lines clearly
show the loop induced by the driving field, which in turn drives the out-of-plane MD mode. In
(c) however, this loop is not clearly visible. Detailed analysis of crosscuts through other planes
(not shown here) do show that there is a current loop. But, a large fraction of the current loop
is inside the evanescent near field in the air below the particle, where n=1 (hence the high
field intensity at the bottom corners). As the particle gets taller, Figs. 1(d) and 1(e), the current
loop fits better inside the particle and a larger fraction of the field is in the high index medium,
giving rise to a red-shift which saturates when the current loop easily fits inside. A similar
argument holds for the ED mode, where a larger fraction of the field is inside the cylinder for
larger h, giving rise to a red-shift.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26289
Concomitant with the strong blue-shift for smaller h, the index of Si also strongly increases
for smaller wavelengths (from n=4at
λ
=550 nm to n=5.1at
λ
=420 nm). This makes
the particles ”optically” higher, thus creating more space for a current loop and as a result, the
MD mode does not disappear. Note that for a non-dispersive medium this might not be the case.
Figure 2(a) shows Qscat as a function of wavelength for a d=250 nm cylinder in air, made
of a non-dispersive dielectric with n=2 for different heights in the range h=50 200 nm.
For h=50 100 nm, one very broad peak is observed, which corresponds to the ED mode.
When h>100 nm, there is enough retardation to support a displacement current loop and the
MD resonance is supported as well, which then red-shifts for larger h, as explained above.
Note that the Q-factor of the resonances is much lower than observed in Fig. 1(a) since the
lower index of the particle results in poor mode confinement and thus a large radiative loss
rates
γ
rad. Therefore, the ED and the MD spectrally overlap and one broad ED/MD peak is
observed. Also note that the identification of the modes is again done by studying the field
profiles. In this case however, the ED and MD modes spectrally overlap, such that in the near-
field profiles both dipolar field profiles and corresponding current loops can be distinguished.
The relatively sharp peak that appears in the blue spectral range for larger heights corresponds
to the magnetic quadrupole mode (MQ), as will be shown later. Thus, in the limit that there is
not enough retardation of the driving field inside the particle (low nand h), the MD mode is not
supported, whereas the ED mode is always supported.
λ
(
nm
)
Qscat
ED
ED/MD
λ
(
nm
)
300 400 500 600 700 800
Qscat
ED/MD
MQ
300 400 500 600 700 800
0
1
2
3
4
5
6
7
hn=2
250 nm
MQ hn=2
250 nm
0
1
2
3
4
5
6
7
n=2 substrate
h = 50 nm
h = 75 nm
h = 100 nm
h = 125 nm
h = 150 nm
h = 175 nm
h = 200 nm
a) b)
Fig. 2. (a) Qscat as function of wavelength for d=250 nm particles made of a non-
dispersive dielectric with n=2, plotted for heights in the range h=50 200 nm. The
geometry is shown as an inset. The peaks are labeled according to the resonance they cor-
respond to (MD, ED, MQ). For h<125 nm, no MD is supported. (b) Qscat for the same
particles as in (a), but now on a semi-infinite substrate wih n=2 (see inset for geometry).
It has been shown that when a dielectric resonator is positioned on a substrate with similar
index as the resonator, the displacement current loop can extend into the substrate [23] (also
see section 5). The substrate effectively provides additional ”optical” height to the resonator,
such that the MD mode can be supported for shallow particles that do not show this resonance
when positioned in air. Figure 2(b) shows Qscat for the same resonator as in (a), but now posi-
tioned on a semi-infinite substrate with nsub =2 (see inset). Comparison of (a) and (b) clearly
shows that all resonances have significantly broadened, due to the fact that part of the mode
profile extends into the substrate. This increases the radiative loss rate
γ
rad into the substrate,
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26290
resulting in a reduced resonance Q-factor. Although combining shallow particles (h<100 nm)
with a substrate provides enough retardation and space to support a current loop, the degree
of confinement is insufficient to support any resonance. However, as soon as resonant behavior
occurs by increasing h, the resonant peak corresponds to a combination of the ED and MD
eigenmodes.
4. Particle diameter and higher order modes
Increasing the particle diameter denables the excitation of resonant modes at larger wave-
lengths. A rule of thumb for the resonant wavelengths of the dipolar modes of resonant
nanocylinders in air is
λ
/nd. Furthermore, when the diameter is large enough, higher-order
modes will also be supported in the visible spectral range. In Fig. 3 we investigate the spectral
spacing of the ED and MD modes. Figure 3(a) shows Qscat (color) for h=100 nm Si cylinders
in air, with varying diameter in the range d=50250 nm.
For d<130 nm two clear resonances are present, which correspond to the ED and MD
modes, as was observed in Fig. 1 where the particle height was varied. As dincreases, both
modes strongly red-shift and start to overlap spectrally as d>160 nm, while the peak value of
Qscat stays roughly constant (8). The curvature of the MD branch is a result of the dispersive
nature of the Si. Note that the red-shift for the ED mode with increasing dis much stronger
than that of the MD mode, such that tuning the particle aspect ratio allows for control over the
spectral spacing between the ED and MD mode. In fact, a further increase in diameter will
cause the ED and MD branch in Fig. 3(a) to cross as shown in [21]. The vertical crosscut in Fig.
3(a) indicated by the white line at d=175 nm, is shown as the red spectrum in Fig. 3(c). The
spectrum shows how the ED mode appears as a shoulder on the dominant MD resonance. Thus,
in air, the ED and MD peaks are spectrally separated for small dand spectrally overlapping for
large d. A third peak is observed for around
λ
=466 nm, which also red-shifts with increasing
d. This peak is due to a magnetic quadrupole (MQ) as appears from Fig. 3(d), which shows a
vertical crosscut, parallel to the H-field of the driving field, of the normalized magnetic field
intensity |H|2(color) and the magnetic field lines (gray lines) in a d=175 nm particle at the
MQ resonance wavelength (
λ
=466 nm). The combination of dand
λ
used for this plot is
indicated with a white dot in Fig. 3(a). The crosscut shows four bright lobes in the corners
of the particle (the bottom two overlap, giving a bright spot in the center), corresponding to
the four poles of a quadrupole. The field lines show which pole is positive (field lines going
out of the particle) and negative (field lines going in), indicated by white +and signs. The
observed mode sequence MD-ED-MQ with decreasing resonance wavelengths in Figs. 3(a)
and 3(c) is in agreement with the well established mode hierarchy of magnetic and electric
modes in spheres [12, 16]. Note that a strong reduction in the cylinder aspect ratio can cause
the MD mode to be positioned at lower wavelength than the ED mode, thereby changing this
mode hierarchy [21].
Figure 3(b) shows Qscat for the same resonators as in (a), but now on a semi-infinite substrate
with nsub =3.5 (roughly equal to nSi, see inset for geometry). Comparing (a) and (b), four char-
acteristic differences are observed. First, the ED and MD modes in (b) are spectrally broader
and overlapping, now also for small d, which is a result of the increase of
γ
rad into the substrate.
Second, the peak values for Qscat are much higher for small d(11 instead of 8) and for
the MQ mode (7 instead of 5). Both the first effect (broadening) and second effect (in-
crease in Qscat) are attributed to the increased local density of optical states (LDOS) accessible
to the resonator when located on a substrate: there are more radiative states available to couple
to, thus increasing
γ
rad and therefore Qscat. Third, for d>160 nm the ED/MD peak becomes
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26291
50 nm
d)
+
+
λ (nm)
50 100 150 200 250
400
500
600
700
800
900
1000 0
2
4
6
8
10
diameter (nm)
λ (nm)
50 100 150 200 250
400
500
600
700
800
900
1000 0
2
4
6
8
10
diameter (nm)
Si
d
100 nm
Si
d
100 nm
n=3.5 substrate
400 500 600 700 800 900 1000
0
1
2
3
4
5
6
7
8
9
λ (nm)
Qscat
ED
MD
MQ
ED/MD
0
1
d=175 nm
ED
MD
MQ
ED/MD
MQ
a) b)
c)
normalized |H|2
Qscat Qscat
Fig. 3. (a) Qscat (color) as a function of wavelength for a h=100 nm Si cylinder in air, for
diameters in the range d=50250 nm. The geometry is shown as an inset. The peaks are
labeled according to the resonance they correspond to (ED,MD,MQ). (b) Qscat (color) for
the same resonator as in (a), but now on a semi-infinite substrate with n=3.5 (see inset).
(c) Spectra corresponding to crosscuts indicated by white certical dashed lines in (a) and
(b) showing the spectra for d=175 nm, in air (red) and on a n=3.5 substrate (blue).
The labels indicate the corresponding resonant modes. (d) Vertical crosscut through the
resonator in air (d=175 nm, see white dot in (a)) in the plane parallel to the H-field of the
driving field for
λ
=466 nm. Plotted is the normalized |H|2(color) and the magnetic field
lines (gray), showing the MQ mode profile. The corresponding poles are labeled with +
and signs (white).
very broad, scattering light over the entire visible spectral range, since a larger fraction of the
mode profile extends into the substrate for larger
λ
. Figure 3(c) (blue) shows the spectrum for
d=175 nm. Clearly, the ED/MD peak ranges all the way from
λ
=400 nm up to
λ
=1000
nm. These modes can serve to achieve efficient light coupling into a high-index substrate over
a broad spectral range, as shown in [23,24]. Fourth, the red-shift of the ED/MD peak is more
dramatic than in (a), which is a result of the fact that a large fraction of the near-field of the
resonance is inside the n=3.5 substrate instead of air (n=1).
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26292
5. Substrate index and particle shape
Sections 3 and 4 have shown that when a cylindrical resonator is positioned on a substrate with
nsub npart, the ED and MD resonances broaden and spectrally overlap. Next, we investigate
the case of 1 <nsub <npart. We study how the spectral positions of the resonances depend on
substrate index and particle shape. Figure 4(a) shows Qscat for the same resonator as in Fig.
1(a), a Si cylinder with d=100 nm and h=100 nm, but now on a semi-infinite substrate with
an index in the range 1 nsub 3.5, varying in steps of 0.25.
Figure 4(a) shows three characteristic features. First, as was observed in Figs. 2(b) and 3(b),
increasing nsub from1to3.5 gradually broadens the resonance peaks (due to an increase in
γ
rad) and causes the ED and MD peaks to overlap spectrally, such that the ED mode becomes
a shoulder of the MD mode for high index. For nsub >2.25, the ED is no longer visible as a
separate peak. Second, a significant increase in Qscat is observed for increasing nsub, which is
attributed to the increase in the LDOS at the position of the particle. Third, a small red-shift is
observed with increasing nsub:
λ
MD =499 nm for nsub =1 and
λ
MD =515 nm for nsub =3.5.
This is caused by the index experienced by the near-field of the resonance. Note that this
small shift of only 16 nm is in sharp contrast with the dramatic 600 nm red-shift observed
for a plasmonic dipolar resonance of a similar cylinder made of Ag [27]. This contrast is
a direct result of the different nature of the resonance. While in plasmon resonances most
light is concentrated at the interface (in direct contact with the substrate), in case of dielectric
resonators most light is concentrated inside the resonator, making it less sensitive to substrate
index.
To investigate the influence of particle shape, Fig. 4(b) shows Qscat as a function of
wavelength for the same cylinder as in (a) (red), a d=114 nm sphere (green), and a cube with
sides of length l=92 nm (blue). The particles have the same volume, are all made of Si and the
geometries are shown below (a) and (b). The dashed lines in Fig. 4(b) correspond to particles in
air, the solid line to particles on a semi-infinite substrate with nsub =3.5. Note that the sphere
is sticking 7 nm into the substrate (center is at 50 nm height above the substrate) to prevent
unphysical hot spots at the infinitely sharp contact area (see sketched geometry). Figure 4(b)
shows that both in air and on a substrate, the shape of the particle has only a small influence
on the resonance wavelengths of the ED and MD modes. Both resonances are supported
for all shapes, and the cylinder proves to be the most efficient scatterer. Furthermore, for all
shapes Qscat increases significantly due to the enhanced LDOS, when put onto a substrate.
However, comparing the solid lines with the dashed lines shows that both the cylinder (red)
and the cube (blue) show the dramatic broadening of the ED and MD resonances when
compared to the case in air, causing the ED to appear as a shoulder on top of the MD mode. For
the sphere (green), the broadening is significantly less and the ED remains to be a separate peak.
To understand this difference, we plot vertical crosscuts (parallel to the E-field of the driving
field) through the center of all particles on the nsub =3.5 semi-infinite substrate, at the MD
resonance wavelength. Figures 4(c)–4(e) show the normalized |E|2(color) and theelectric field
lines (gray), while Figs. 4(f)–4(h) show the normalized |H|2(color). The magnetic field lines
are not plotted because H0 in this plane, which is perpendicular to the H-field orientation of
the driving field. Figures 4(c)–4(e) clearly show the displacement current loops that induce the
magnetic dipole mof the MD mode. Both (c) and (e) show that for the cylinder and the cube a
significant fraction of the displacement current loop extends into the substrate. For the sphere,
despite the fact that it sticks into the substrate, this effect is much smaller.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26293
400 450 500 550 600 650 700
0
2
4
6
8
10
12
λ
(nm)
Qscat
100 nm
Si
100 nm
nsub
n=1
n=3.5
MD
ED
a)
100 nm
100 nm
Si
114 nm
Si
92 nm
92 nm
92 nm
Si
λ (nm)
Qscat
b)
0
50 nm50 nm
50 nm
50 nm 50 nm 50 nm
normalized |E|2
on n=3.5 substrate
in air
cylinder
sphere
cube
c) d) e)
f) g) h)
400 450 500 550 600 650 700
1
3
5
7
9
11
1
0
normalized |E|2
1
0
normalized |E|2
1
0
normalized |H|2
1
0
normalized |H|2
1
0
normalized |H|2
1
n=3.5
air
n=3.5
air
n=3.5
air
n=3.5
air
n=3.5
air
n=3.5
air
Fig. 4. (a) Qscat as a function of wavelength for a Si cylinder with h=100 nm, d=100 nm,
on a semi-infinite substrate with 1 nsub 3.5, in index steps of 0.25. The geometry is
shown as an inset. (b) Qscat for the same cylinder (red), a d=114 nm sphere (green), and a
l=92 nm cube (blue) in air (dashed), all on a nsub =3.5 semi-infinite substrate (solid). All
particles consist of Si and have the same volume. The three geometries are sketched below
(a) and (b). Note that the sphere sticks 7 nm into the substrate to prevent an infinitely sharp
contact area. (c-e) Vertical crosscuts through the center of all three particles in the plane
parallel to the E-field of the source, showing the normalized |E|2(color) and electric field
lines (gray). The particle surroundings and air-substrate interface are indicated with white
dashed lines. The respective geometries are shown above the figures. The displacement
current loops are clearly visible. (f-h) The same crosscuts as in (c-e), now showing the
normalized |H|2(color) of the MD modes. The magnetic field lines are not plotted since
H0 in this plane (perpendicular to source
H)
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26294
As a result, the coupling to the substrate is much weaker for the sphere than for the cylinder
and cube. This is also clearly observable in Figs. 4(f)–4(h), which show that the magnetic field
profile of the MD mode inside the substrate is much brighter for the cylinder and the cube,
when compared to the sphere. The contact area of the resonator with the substrate is thus very
important to get strong coupling to the substrate and provides tunability of the radiation rate
γ
rad into the substrate. Note that in Figs. 4(c)–4(e) a second (reversed) current loop can be
observed in the substrate below the particles. We attribute this to image dipoles induced by the
dipole moments in the particles.
6. Selective mode enhancement with multi-layer substrates
Based on the results from sections 3-5 it is possible to design any resonator geometry
such that the particle is resonant at the desired wavelength. Furthermore, by tuning the
diameter and height separately it is possible to either have two spectrally separated ED/MD
resonances (using a small diameter, large height) or spectrally overlapping resonances (large
diameter, small height). However, for specific applications it may be useful to selectively
select one of the two modes to dominate the response of the resonator; the ED or the MD mode.
This may be particularly relevant if the radiation pattern of the resonator is important, such
as for example in light trapping geometries in photodiodes and solar cells, as will be shown
in section 8. Here we will show that this can be done using a multi-layer substrate. The inset
of Fig. 5(a) shows the proposed geometry: a Si cylinder with d=100 nm and h=100 nm is
positioned on a SiO2layer with thickness t, which is on top of a nsub =3.5 substrate (e.g. Si).
We vary the SiO2layer thickness from 0 t500 nm in steps of 20 nm and for each twe
calculate Qscat.
λ (nm)
t (nm)
0100 200 300 400 500
400
450
500
550
600 2
4
6
8
10
12
Si
100 nm
100 nm
n=3.5 substrate
SiO2t
MD
ED
12
14
10
8
6
8
10
4
2
Qscat MD
Qscat ED
λ=499 nm
λ=438 nm
t (nm)
0 100 200 300 400 500
Qscat
a) b)
c)
Fig. 5. (a) Qscat as a function of wavelength and oxide thickness t, for a d=100 nm,
h=100 nm Si cylinder, positioned on top of an oxide layer with thickness t, on top of a
semi-infinite substrate with nsub =3.5. The inset shows the geometry. The horizontal black
dashed lines indicate crosscuts at resonance wavelengths:
λ
=440 nm (ED) and
λ
=499
nm (MD). (b-c) Blue dots show the cross sections from (a), red lines show the fits from the
interference model. The free-space wavelengths found from the fit are shown on the bottom
left of the figures. The gray dashed lines show the peak value of Qscat for the same particle
on a semi-infinite SiO2substrate (limit of t).
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26295
The results are shown in Fig. 5(a), where Qscat is shown (color) as a function oftand
λ
.For
t=0, we observe one very broad peak: the resonator is strongly coupled to the substrate and
the ED and MD mode broaden and overlap as in Fig. 4. However, as tincreases Qscat of the MD
mode and ED mode are subsequently suppressed and enhanced. The horizontal dashed black
lines indicate crosscuts near the resonance wavelength for the MD mode (
λ
=499 nm) and the
ED mode (
λ
=440 nm), which are shown as blue dots in Figs. 5(b) and 5(c) respectively. Note
that the oscillation period is different for the ED and MD modes as they appear at different
wavelengths, such that any combination of MD/ED amplitude within the oscillation can be
chosen: at t400 nm both modes are in phase while at t60 nm they are out of phase. This
effect is caused by constructive and destructive interference of the incident light reflected from
the SiO2layer, resulting in strong variations in the driving field in the resonator, as follows
from a simple interference model. Indeed, the oscillation periods found in Figs. 5(b) and 5(c)
correspond to the half-wavelength of light in the oxide layer.
To prove that this effect allows for both suppression and enhancement of one of the modes,
the gray dashed lines in (b) and (c) show the peak value of Qscat of the same particle on a
semi-infinite substrate (t) with index nsub =1.5 (roughly equal to nSiO2), which is 8.8 for
the MD mode and 5.9 for the ED mode. As can be seen from both (b) and (c), the amplitude of
Qscat can both be lower and higher than these values. The modes are not fully suppressed since
the amplitudes of the incoming light and the light scattered from the Fabry-P´
erot cavity do not
completely cancel each other.
7. Plane wave versus dipole excitation
All the above describes the resonant behavior of cylinders, illuminated by plane-waves under
normal incidence. However, in many practical applications or experiments the driving field
may be light under large angles of incidence (e.g. illumination through a large NA objec-
tive [11,16]), or even a local source like an electron beam excitation [28], which corresponds
to a vertically oriented electric dipole [29,30]. Therefore, we investigate the resonant behavior
of cylinders excited by a point dipole. We study the scattering cross section of a d=100 nm,
h=100 nm Si cylinder in air and compare the data with those for plane wave illumination
from the top (like in sections 3-6). Efficient driving occurs when the fields of the electric dipole
source overlap those of the eigenmode. Thus, when the dipole is positioned in an anti-node if
the eigenmode field profile. Here, we average over all dipole positions to find all eigenmodes
that are efficiently excited by a vertical dipole. Note that this experimental configuration
corresponds to that of cathodoluminescence spectroscopy, as we have recently shown for
dielectric cylinders [28]. To simulate this, a vertically oriented electric dipole is positioned
in the center of the particle, and then scanned to the edge of the particle in steps of 5 nm.
For each step, a simulation is performed and the light scattered from the particle is collected
using transmission monitors. Next, because of cylindrical symmetry, the dataset is rotated
around the axis, and the total particle spectrum is obtained by adding all the individual spectra
(using the proper Jacobian). Since the dipole driving power and plane wave base amplitude can
be arbitrarily chosen, the spectra from both simulations are normalized for comparison. The
results are shown in Fig. 6(a), where the blue line corresponds to dipole excitation and the red
line to plane wave excitation. Both spectra show two peaks, corresponding to the MD mode at
long wavelengths and the ED mode at short wavelengths. Interestingly, the figure indicates that
both excitation mechanisms induce a similar MD mode, at the same wavelength, whereas the
ED energies are distinctively different.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26296
MD
ED
in-plane vertical
m
p
H
E
k
p
driving
field
400 450 500 550 600 650 700
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
λ (nm)
normalized radiated power
b)
a)
plane-wave
vertical dipole
MDhor
EDhor
EDvert
1
0
c) EDvert
MDhor
Fig. 6. (a) Normalized Qscat for a Si cylinder with d=100 nm, h=100 nm in air under
plane wave excitation (red), and excited by a vertically oriented electric dipole, calculated
by integrating spectra obtained from scanning the dipole position over a horizontal cross-
cut through the middle of the particle (blue). The modes are labeled in gray. (b) Excitation
mechanism for the in-plane polarized plane wave source (left column) and vertically ori-
ented electric dipole source (right column). The driving mechanism is shown for both the
MD mode (top row) and ED mode (bottom row). The driving field orientation is shown on
top. The orange circles indicate the displacement current loop inducing m, the red arrows
the E-field of the driving field (top row) and porientation (bottom row). The blue crosses
indicate the out-of-plane orientation of m.(c) Excitation efficiency maps of the EDvert (top)
and MDhor (bottom) modes driven by a vertically oriented electric dipole. The cross sec-
tions are taken at the particle half-height.
To further characterize this difference, we show in Fig. 6(b) the driving mechanisms of both
the MD mode (top row) and the ED mode (bottom row), for the in-plane excitation using a
plane wave (left column) and vertical dipole (right column). The driving field orientations are
shown on top. For the plane wave driving field (left column), the excitation of the in-plane
MD (labeled MDhor) is a result of the E-field of light (red) coupling to a vertically oriented
displacement current loop (orange), as was discussed in section 2. The projection of the driving
field on the current loop determines the driving efficiency, which occurs on the top and bottom
of the current loop in this case. Coupling of the plane wave to the in-plane ED (labeled EDhor)
mode (bottom left) occurs due to the in-plane polarization of the material by the driving field.
The vertically oriented dipole excitation (right column) has the E-component of the driving
field in the vertical direction. However, this driving field has good field overlap with the sides
of the vertical displacement current loop, on the left and right of the particle (see top right
in Fig. 6(a)). Therefore, the same in-plane m(MDhor) can be induced with both vertical and
in-plane E-fields, which explains why the peaks occur at the same resonance wavelength
(
λ
=503 nm).
The ED mode, however, is excited differently for the two cases. The ED mode excited by
the vertical dipole source is vertically oriented (EDvert), instead of horizontally. The different
alignment of the dipole moment pwithin the particle causes the eigenmode to occur at different
wavelengths: EDvert (417 nm) occurs at shorter wavelengths than EDhor (440 nm), as can be
seen in (a). Note that if a horizontal point dipole would be coupled to the particle, the EDhor
mode occur at the same resonance wavelength, while the MD mode will change. Under these
conditions, an in-plane current loop will be excited rather than a vertical one. So, the peak
corresponding to MDhor mode would occur in the spectrum, with a different resonance energy.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26297
To further corroborate the excitation mechanisms by the vertical point dipole, Fig. 6(c) shows
the excitation efficiency maps (corresponding to horizontal crosscuts of the mode profiles at
the particle half-height) for the EDvert (top) and MDhor mode (bottom). These maps show how
much light is scattered by the particle at resonance wavelength (20 nm bandwidth) as a function
of the excitation position (position of the dipole). This corresponds to the coupling efficiency,
which is determined by the field overlap of the eigenmodes and the E-fields of the dipole. Figure
6(c) shows that the EDvert (top) is excited most efficiently when the dipole is positioned in the
center (bright spot), which corresponds to the position of the vertically oriented p(see bottom
right in (a)). Note that the slightly brighter ring on the outside is caused by inefficient coupling
to the vertical component of the field lines of the EDhor mode [31]. The MDhor mode is excited
most efficiently when the dipole is positioned in the bright ring on the edge of the particle, in
agreement with the top right sketch in (a). Note that a ring is observed because the MDhor mode
is infinitely degenerate with respect to
φ
, which is a result of the cylindrical symmetry of the
particle.
8. Radiation patterns
One of the most important characteristics of dielectric resonators is the angular radiation
pattern. As specific eigenmodes are often characterized by distinct radiation profiles [28],
this allows for the design of angular or spectrally sensitive photodetectors, and solar cell
geometries with optimized light trapping. For resonators that are strongly coupled to a
high-index substrate, a large fraction of the scattered light is emitted into the substrate due to
the enhanced LDOS [32]. Here, we use the far-field transformation routine of Lumerical [25]
to calculate the radiation patterns of single silicon scatterers on a substrate under plane wave
illumination. We use a large (2.m×2.m×0.8 µm) simulation volume and apply the
far-field transformation to the light propagating downward into a semi-infinite substrate with
index nsub, collected by a monitor outside the TFSF source (150 nm below the air-substrate
interface). In this way, the radiation patterns only include the scattered light and not the
incident light.
Figure 7(a) shows Qscat of a Si cylinder with d=100 nm, h=100 nm on a semi-infinite
substrate with nsub =1.5 (red) and nsub =3.5 (blue). The geometry is shown as an inset, where
we define
θ
as the angle with respect to the surface normal pointing downward. For nsub =1.5,
we observe two distinct resonances due to the ED mode (labeled 1) and the MD mode (2).
For nsub =3.5, both peaks have broadened such that they spectrally overlap and the resonant
response is dominated by the MD mode (3).
Figure 7(b) shows the fraction of the scattered power that is scattered into the substrate as
a function of nsub, integrated over all wavelengths (red) and at the MD resonance wavelength
(blue). Both lines show an increase in downward scattering with increasing nsub due to the
increased LDOS [32]. For nsub =3.5 (close to nSi) more than 80% of the scattered light
goes into the substrate. Indeed, arrays of Si nanocylinders have shown to act as efficient
anti-reflection coatings [23]. Surprisingly, this scattering fraction is lower at resonance than
it is for the total power, which shows that the directionality is not optimum at the resonances
wavelength itself. However, the absolute power scattered into the substrate is determined by
Qscat and does peak at resonance (also see Fig. 8).
To show that the individual resonances can be used as directional antennas, we show in Fig.
7(c) the angular radiation profiles into the substrate for the Si cylinder (top row), as a function
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26298
400 450 500 550 600 650 700
0
2
4
6
8
10
12
λ (nm)
Qscat
n=3.5
MD
ED
1
2
3
1.5 22.5 33.5
index substrate
fraction into substrate
Si
100 nm
100 nm
nsub θ
n=1.5
a) b)
Si cylinder
point dipole
123
0.65
0.7
0.75
0.8
0.85
0.9
at MD resonance
integrated
0 (y)
45
90 (x)
135
180
225
270
315
10
20
30
40
50
MDy
EDx
c)
1
0.7
0
1
0
pxmymy
Fig. 7. (a) Qscat for a Si cylinder with d=100 nm, h=100 nm positioned on a semi-infinite
substrate with nsub =1.5 (red) and nsub =3.5 (blue). The geometry is shown as an inset.
The peaks are labeled according to the corresponding resonance (ED / MD) and numbered
for reference in (c). (b) fraction of scattered power into the substrate as a function of nsub,at
resonance wavelength (blue) and integrated over the entire spectral range (400 700 nm).
A clear increase in downward scattered power is observed with increasing nsub, attributed
to the enhanced LDOS in the substrate. (c) Far-field radiation patterns for the numbered
peaks from (a) corresponding to the ED and MD modes of the Si cylinder (top row). 12
correspond to the ED and MD mode on a nsub =1.5 respectively, 3 to the MD mode on
nsub =3.5. The gray circles correspond to constant
θ
with equidistant spacing Δ
θ
=10.
The white dashed circles correspond to the critical angle
θ
cin the substrate. The orientation
of the ED and MD dipole moments are shown on the left. The bottom row shows the
calculated radiation patterns of the corresponding point-dipoles (pxand my), positioned 50
nm above the substrates with nsub =1.5 (left, center) and nsub =3.5 (right). Note that the
colorbar of the bottom row has been saturated at 0.7 for visibility.
of azimuthal angle and zenithal angle. From left to right, the patterns correspond to peaks
13 in Fig. 7(a), respectively. Incident light is polarized along ˆx, i.e. the
θ
=90direction.
The orientation of the ED and MD dipole moments is shown by the gray arrows on the left.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26299
Furthermore, the critical angle for light trapping in guided modes is plotted as a white dashed
line in each radiation pattern.
Radiation pattern 1 shows clear beaming from the ED mode into two specific angular
lobes along ˆy, close to the critical angle. The radiation pattern of the MD mode (pattern 2)
shows similar directional beaming close to the critical angle, but now along ˆx. The orientation
of the radiation patterns correspond to that of an electric dipole and magnetic dipole on a
substrate, oriented in the ˆx-direction (px) and ˆy-direction (my) respectively. To show this, we
also plot the radiation profiles into the substrate of point dipoles (px, bottom left, and my,
bottom center and bottom right) positioned 50 nm above a substrate with nsub =1.5 (bottom
left and bottom center) and nsub =3.5 (bottom right). The radiation patterns are obtained
from classical theory [33, 34]. Comparing radiation patterns 1 and 2 with those of pxand my
we find qualitative agreement, indicating that 100 nm high cylindrical particles at resonance
have radiation patterns similar to point dipoles. Finally, radiation pattern 3 in Fig. 7(c) shows
how the radiation pattern of the ˆy-oriented MD (MDy) resonance is influenced by an increase
in refractive index of the substrate to nsub =3.5. The same directional beaming is observed,
but now more pronounced (stronger beaming) and under smaller
θ
as the critical angle has
decreased to
θ
c=16.6. Again, good qualitative agreement is observed with the point dipole
model (bottom right).
The patterns observed in Fig. 7(c) show that selective excitation of eigenmodes (ED or
MD) allows for directional emission in the ˆxor ˆy-direction. Therefore, the combination of
spectrally separated ED/MD modes can be used for wavelength demultiplexing of polarized
light: short wavelengths are scattered along ˆyby the ED mode, long wavelengths along ˆxby the
MD mode. The height and diameter can be tuned to obtain the desired spectral response, the
index of the substrate defines the bandwidth and scattering angle (i.e. SiO2gives large angles).
Furthermore, the strong emission around the critical angle has direct implications for the use
of such resonators for light trapping in weakly absorbing substrates, such as solar cells. Note
that Fig. 7 only considers the radiation patterns of the lowest order dipolar modes. It has been
observed that higher order modes show strong beaming around the normal [28]. Furthermore,
the patterns observed in Fig. 7(c) change when the particle gets taller. In this case, the effective
dipole moments are positioned further away from the substrate, such that a larger fraction of the
emission is below the critical angle. Simulations show that increasing the particle height from
100 nm to 200 nm significantly changes the radiation patterns from dipolar like emission pro-
files as observed in Fig. 7(c), to strong beaming along the normal due to the increased distance
of the effective dipole moments to the substrate. Altogether, this indicates that tuning the geom-
etry of the resonator can give accurate control over the radiation properties of the scattered light.
Figure 8(a) shows a comparison of the forward (green) and backward (blue) scattered power
for a d=100 nm, h=100 nm Si cylinder on a substrate with nsub =1.5. Also shown are the
ratio Pdown/Pup (red), and the spectral range of the ED (shaded red) and MD (shaded blue)
resonances. Note that the oscillations in the red line for
λ
>580 nm are due to numerical
noise. The scattering ratio peaks at 14.8, which indicates strong directional scattering into the
substrate. This is a result of the interference of light scattered by the ED and MD modes, as
was theoretically predicted by Kerker et al. [35] and recently demonstrated experimentally for
spheres [11,20] and cylinders [21]. For Si spheres, the peak ratio observed was 8 [11], which
shows that particles that have strong coupling to a substrate can exceed this ratio significantly.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26300
However, in the latter case the directionality also originates from the high LDOS of the sub-
strate. To differentiate these two effects, Fig. 8(b) shows the ratio Pdown/Pup as a function of
wavelength for 1.25 nsub 3.5 in steps of 0.25. As was observed in Fig. 4(a), with increasing
nsub both the ED and MD resonance peaks significantly broaden and the total scattered power
increases. The peak scattering ratio is significantly reduced for larger nsub. Thus, the contrast
from the ED/MD interference is strongly suppressed, whereas the average value is similar (6.4
for nsub =3.5, 6.8 for nsub =1.5). Therefore, this graph clearly shows the trend from scattering
dominated by Kerker-type interference (high peaks and dips, low nsub) to scattering dominated
by LDOS (flat and high response, for high nsub). The reason for this transition can be understood
from the requirements for Kerker-type interference: as nsub increases, the overlap in the far-field
radiation patterns is reduced and the relative phase difference is less well defined. The combina-
tion of Kerker-type interference and LDOS dominated scattering can be tuned to get optimum
directionality for one specific wavelength, or reduced directionality over a broad spectral range.
Note that the optimum in forward scattered power always occurs at the MD resonance (see also
Fig. 7(b)) [21].
400 450 500 550 600 650 700
0
5
10
15
20
25
λ
(
nm
)
ratio
n=1.25
n=3.5
b)
ratio
400 450 500 550 600 650 700
0
1
2
3
4
5
6
7
Pup
Pdown
Pdown/Pup
ED MD
a)
scattered power (a.u.)
5
0
(
nm
)
λ
nsub=1.5
15
10
Fig. 8. (a) Back scattered power (blue), forward scattered power (green) and for-
ward/backward ratio (red) for a d=100 nm, h=100 nm Si cylinder on nsub =1.5. The
red and blue shaded regions indicate the spectral reange of the ED and MD resonances re-
spectively. Note that the oscillations in the red curve for
λ
>580 nm are due to numerical
noise. (b) Forward/backward scattering ratio as a function of wavelength for different nsub
ranging from 1.25 to 3.5 in steps of 0.25.
9. Conclusions
In conclusion, we have systematically studied the influence of particle geometry and dielectric
environment on the resonant behavior of single dielectric particles, both isolated and strongly
coupled to a substrate. We found that the lowest energy mode is an in-plane MD mode, driven
by a vertical displacement current loop inside the particle. In the limit of low nand low h,
there is not enough retardation to form a current loop, such that no MD mode is supported.
The excitation of the ED mode is not limited by such constraints. At larger diameters, the ED
and MD spectrally overlap, and the higher order MQ mode appears in the blue spectral range.
Furthermore, we observe that strong coupling to (dielectric) substrates requires a large contact
area between the resonator and the substrate. This allows the displacement current loop to
extend partially into substrate, yielding high radiative losses into the substrate. More complex
multi-layer substrate allow selective enhancement and/or suppression of ED and MD modes
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26301
through coupling with a Fabry-P´
erot mode in the multi-layer substrate. Also, by comparing
dipole and plane wave sources, we found that both vertical and horizontal E-fields can excite
the in-plane MD mode, whereas the ED modes have orthogonal dipole moments and different
resonance energies. Finally, by studying radiation patterns into the substrate we found that the
ED and MD modes have radiation patterns similar to point dipoles above substrate, showing
strong directional beaming around the critical angle. Altogether, this work is a guideline for the
engineering of dielectric resonators, which allows one to use the combined scattering from ED
and MD modes to taylor the scattering properties for specific applications.
Acknowledgments
We would like to thank Toon Coenen for useful discussions and careful reading of the
manuscript. Furthermore, we acknowledge Femius Koenderink for providing the magneto-
electric point-dipole code for calculating far-field radiation patterns. We thank SARA Com-
puting and Networking Services (www.sara.nl) for their support in using the Lisa Compute
Cluster. This work is part of the research program of the Stichting voor Fundamenteel On-
derzoek der Materie (FOM), which is financially supported by the Nederlandse Organisatie
voor Wetenschappelijk Onderzoek (NWO). This work was also funded by an ERC Advanced
Investigator Grant.
#195676 - $15.00 USD
Received 27 Aug 2013; revised 11 Oct 2013; accepted 16 Oct 2013; published 25 Oct 2013
(C) 2013 OSA
4 November 2013 | Vol. 21, No. 22 | DOI:10.1364/OE.21.026285 | OPTICS EXPRESS 26302
... As shown in Fig. 2(c), a periodic array of a-Si nanobeams with 320 nm height can achieve full 2 phase control with high transmission amplitude in the 100 nm to 290 nm width range. In this width range, Fig. 2(d) shows that the nanobeams support optical Mie-type resonances [32,33]. Each nanobeam of a different width is combined as a supercell, which acts as a phase pixel that collectively generates a desired wavefront for the transmitted light. ...
Article
Full-text available
A metasurface is an ultrathin-layered artificial material which consists of an array of nanoscaled optical element and is capable of manipulating light by adjusting the phase in subwavelength scale. Such structures are typically fabricated by cascaded CMOS (Complementary Metal-Oxide-Semiconductor) processes including deposition, etching, and lift-off techniques. Recently, as the demand of optical functionality increases, multiple metasurfaces are designed and integrated in cascaded fashion which, as a trade-off, causes more fabrication complexity and cost. In this article, we present the double-layered metasurface for phase and wavefront control using dielectric materials, which can be easily fabricated by a single lithographic step without the need for additional etching or lift-off process. The proposed metasurface consists of amorphous silicon (a-Si) nanobeams acting as phase retarders based on the Mie-type resonance and an additional polymer layer serving as nanostructured spacers between the constituent a-Si layers. We analyze the effect of actual fabrication techniques on the performance of our proposed metasurface and demonstrate that the double-layered phase gradient metasurface (PGM) deflects light approximately twice the angle compared to that of single-layered PGM. Furthermore, we extend our study to double-layered metalenses and achieve a high numerical aperture (NA) exceeding 0.9. The proposed design concept demonstrates the potential for enhancing light manipulation and paving the way for the development of advanced optical devices with improved performance and a simplified fabrication process.
... Figure 4(a) illustrates the impact of thickness (H) on absorbance, demonstrating a noticeable redshift in λ 1 and λ 2 as H increases. According to previous reports [44], as the particle height increases, the magnetic current loop becomes more compatible within the particle, resulting in a greater proportion of the electric field being confined to the silicon disk. This phenomenon may be the cause of the red shift in λ 1 . ...
Article
Full-text available
Due to its unique advantage of optical properties, nanophotonic metamaterials have gained extensive applications in perfect absorbers. However, achieving both dual-band and ultra-narrow linewidth in absorbers simultaneously remains a challenge for typical metal-dielectric based metamaterials. In this work, a dual-band ultra-narrow perfect absorber consisting of a double slotted silicon nanodisk array that located on a silver film with a silica spacer layer is proposed theoretically. By combining the hybrid mode excited by the coupling of diffraction wave mode and magnetic dipole mode with the anapole–anapole interaction, two absorption peaks can be induced in the near-infrared regime, achieving nearly perfect absorbance of 99.31% and 99.61%, with ultra-narrow linewidths of 1.92 nm and 1.25 nm respectively. In addition, the dual-band absorption characteristics can be regulated by changing the structural parameters of the as-proposed metamaterials. The as-designed metamaterials can be employed as efficient two-channel refractive index sensors, with sensitivity and figure of merit (FOM) of 288 nm RIU⁻¹ and 150 RIU⁻¹ for the first band, and sensitivity and FOM of up to 204 nm RIU⁻¹ and 163.2 RIU⁻¹ for the second band. This work not only opens up a new design idea for the realization of dual-band perfect absorber synchronously with ultra-narrow linewidth, but also provides potential attractive candidates for developing dual-frequency channel sensors.
... According to our resonance mode analysis based on CMD ( Supplementary Fig. 9), it is found that the micro-pyramid with w = 10 µm supports a strong electric dipole resonance, and weak magnetic dipole, electric quadrupole, and magnetic quadrupole resonances. Since the micro-pyramid is placed on the polydimethylsiloxane film, those resonance modes can couple to the polydimethylsiloxane film, which can cause strong forward scattering into the polydimethylsiloxane film 56,57 . Consequently, the reflection at the air-polydimethylsiloxane interface diminishes, more energy of the incident wave can be transmitted into the polydimethylsiloxane film, and the PMMM' absorption in MIR range is thus enhanced based on Beer's law. ...
Article
Full-text available
Transparent roofs and walls offer a compelling solution for harnessing natural light. However, traditional glass roofs and walls face challenges such as glare, privacy concerns, and overheating issues. In this study, we present a polymer-based micro-photonic multi-functional metamaterial. The metamaterial diffuses 73% of incident sunlight, creating a more comfortable and private indoor environment. The visible spectral transmittance of the metamaterial (95%) surpasses that of traditional glass (91%). Furthermore, the metamaterial is estimated to enhance photosynthesis efficiency by ~9% compared to glass roofs. With a high emissivity (~0.98) close to that of a mid-infrared black body, the metamaterial is estimated to have a cooling capacity of ~97 W/m² at ambient temperature. The metamaterial was about 6 °C cooler than the ambient temperature in humid Karlsruhe. The metamaterial exhibits superhydrophobic performance with a contact angle of 152°, significantly higher than that of glass (26°), thus potentially having excellent self-cleaning properties.
... For the NIR band, dielectric metasurfaces provide an exciting platform to drive diverse local optical responses efficiently within the subwavelength scale [7,8]. For instance, Mie resonances can be excited when light strongly interacts with dielectric nanostructures of characteristic size on the order of the wavelength of incident light [9,10]. Moreover, SLR formed by the period-arranged dielectric nanostructures can lead to remarkably narrow linewidths and high quality (Q-) factors. ...
Article
Full-text available
Low-cost nanocomposite metasurfaces have demonstrated attractive potential to replace the equivalent dielectric metasurfaces for light engineering. However, the resonance characteristics of embedded structures in nanocomposite metasurfaces have not been further analyzed beyond the effective refractive index. Herein, we have proposed customizable polarization-selective narrowband meta-filters using ultraviolet-curable (UV) nanocomposites. As an additional degree of freedom, near-field effects between highly concentrated doped nanoparticles can enhance the Mie resonance of the low aspect ratio (AR = 0.2) meta-units. The surface lattice resonances (SLRs) of meta-filters can be coupled with enhanced Mie resonances of individual meta-units to realize tunable narrowband (FWHM ∼0.007λ) reflections with intensities near unity. Meanwhile, the polarization-selective properties of the reflection peaks can be tuned by optimizing the asymmetric lattice. Such proposed new-generation customizable meta-filters will offer, to our knowledge, novel strategies for filtering specific near-infrared polarized fluorescence in the integrated imaging systems.
... The wavefront-shaping metasurface presented in this work is based on multipolar-resonant silicon nanoresonators [44,[48][49][50][51][52], with elliptical cross sections of short axis w, long axis l and height h. For initial characterization, the nanostructures are arranged in an equilateral rhombohedral lattice with side length L = 670 nm ( Fig. 1(a)) and periodic boundary conditions. ...
Article
Full-text available
Structured illumination microscopy (SIM) achieves superresolution in fluorescence imaging through patterned illumination and computational image reconstruction, yet current methods require bulky, costly modulation optics and high-precision optical alignment, thus hindering the widespread implementation of SIM. To address this challenge, this work demonstrates how nano-optical metasurfaces, rationally designed to tailor the far-field optical wavefront at sub-wavelength dimensions, hold great potential as ultrathin, single-surface, all-optical wavefront modulators for SIM. We computationally demonstrate this principle with a multipolar-resonant metasurface composed of silicon nanostructures that generate versatile optical wavefronts in the far field upon variation of the polarization or angle of incident light. Algorithmic optimization is performed to identify the seven most suitable illumination patterns for SIM generated by the metasurface based on three key criteria. We quantitatively demonstrate that multipolar-resonant metasurface SIM (mrm-SIM) achieves resolution gain that is comparable to conventional methods by applying the seven optimal metasurface-generated wavefronts to simulated fluorescent objects and reconstructing the objects using proximal gradient descent. Notably, we show that mrm-SIM achieves these resolution gains with a far-field illumination pattern that circumvents complex equipment and alignment requirements of comparable methodologies. The work presented here paves the way for a metasurface-enabled experimental simplification of structured illumination microscopy.
... At 1350 nm, the magnetic field is predominantly distributed in the Ti@a-C layer and at the interfaces of the Ti@a-C layer, owing to interband transitions in the Ti nanoparticles within the Ti@a-C cermet, the plasmon resonance of embedded Ti nanoparticles in the amorphous carbon matrix [28], and the absorption of PSPR at the upper and lower surfaces of the Ti@a-C layer [29]. Notably, when the resonant wavelengths of MD and ED modes are close to each other, the interaction of MD and ED modes results in a Huygens' metasurface, leading to the suppression of backward scattering [30,31]. This signifies a reduction in reflection within the 700-1500 nm range, a phenomenon referred to as the Kerker condition [32]. ...
Article
Full-text available
Efficient solar thermal conversion is crucial for renewable clean energy technologies such as solar thermal power generation, solar thermophotovoltaic and seawater desalination. To maximize solar energy conversion efficiency, a solar selective absorber with tailored absorption properties designed for solar applications is indispensable. In this study, we propose a broadband selective absorber based on amorphous carbon (a-C) metamaterials that achieves high absorption in the ultraviolet (UV), visible (Vis) and near-infrared (NIR) spectral ranges. Additionally, through metal doping, the optical properties of carbon matrix materials can be modulated. We introduce Ti@a-C thin film into the nanostructure to enhance light absorption across most of the solar spectrum, particularly in the NIR wavelength band, which is essential for improving energy utilization. The impressive solar absorptivity and photothermal conversion efficiency reach 97.8% and 95.6%, respectively. Notably, these superior performances are well-maintained even at large incident angles with different polarized states. These findings open new avenues for the application of a-C matrix materials, especially in fields related to solar energy harvesting.
Article
Optically resonant particles are key building blocks of many nanophotonic devices such as optical antennas and metasurfaces. Because the functionalities of such devices are largely determined by the optical properties of individual resonators, extending the attainable responses from a given particle is highly desirable. Practically, this is usually achieved by introducing an asymmetric dielectric environment. However, commonly used simple substrates have limited influences on the optical properties of the particles atop. Here, we show that the multipolar scattering of silicon microspheres can be effectively modified by placing the particles on a dielectric-covered mirror, which tunes the coupling between the Mie resonances of microspheres and the standing waves and waveguide modes in the dielectric spacer. This tunability allows selective excitation, enhancement, suppression, and even elimination of the multipolar resonances and enables scattering at extended wavelengths, providing transformative opportunities in controlling light–matter interactions for various applications. We further demonstrate with experiments the detection of molecular fingerprints by single-particle mid-infrared spectroscopy and with simulations strong optical repulsive forces that could elevate the particles from a substrate.
Article
Full-text available
Nanomechanical resonators provide a versatile platform for nanoscale mass sensing and force microscopy, as well as for enhancing light‐matter interaction offering unique functionality for optomechanical applications. In this way, discovering new approaches for coupling light with the mechanical degrees of freedom opens the strong desire paths for further developing of nanomechanical technology. Here, the parametric optothermal modulation of hybrid nanomechanical systems consisting of carbon nanowire with a silicon nanoparticle on its top, is reported. The mechanism of the modulation is based on the periodic optical heating of the nanowire and further modulation of the elasticity parameters. Utilizing the silicon nanoparticle provides additional functionality owing to optical absorption enhanced with Mie resonance and the unique feature of optical Raman thermometry enabling optical monitoring of local temperature. It is shown that the parametric mechanism of modulation allows for a significant increase of the optomechanical coupling strength.
Article
Full-text available
This article explores the design and optimization of nanodisk metasurfaces for achieving high reflectivity at a defined wavelength. The telecom wavelength of 1550 nm is particularly focused, selected for its potential applications in next‐generation gravitational wave detectors. At this wavelength, the research goes toward the development of thin, low‐loss, high‐reflective coatings, where the metasurface can be chosen as an alternative. An optimization process for the dimensional parameters of nanodisks is proposed based on a systematic tuning approach, which facilitates the realization of various configurations of high‐reflective metasurfaces. The concept of the “magnetic mirror effect” is examined in detail, where the magnetic dipole resonance aligns with the anapole state. Additionally, high reflectivity at the electric dipole resonance (“electric mirror effect”) and at the excitation of several multipole moments is explored, including high‐order modes. This variety of configurations affords more flexibility in the phase manipulation of the reflected beam. Furthermore, the potential experimental realization of mirror effects is discussed by exploring the structure in the surrounding medium with a refractive index of nd = 1.4. This research platform provides a promising tool for the fabrication of high‐reflective nanodisk metasurfaces and demonstrates its applicability across various fields.
Article
Full-text available
In the framework of the discrete dipole approximation we develop a theoretical approach that allows the analysis of the role of multipole modes in the extinction and scattering spectra of arbitrary shaped nanoparticles. The main attention is given to the first multipoles including magnetic dipole and electric quadrupole moments. The role of magnetic quadrupole and electric octupole modes is also discussed. The method is applied to nonspherical Si nanoparticles with resonant multipole responses in the visible optical range, allowing a decomposition of single extinction (scattering) peaks into their constituent multipole contributions. It is shown by numerical simulations that it is possible to design silicon particles for which the electric dipole and magnetic dipole resonances are located at the same wavelength under certain propagation directions of incident light, providing new possibilities in metamaterial developments.
Article
Full-text available
Interference of optically-induced electric and magnetic modes in high-index all-dielectric nanoparticles offers unique opportunities for tailoring directional scattering and engineering the flow of light. In this article we demonstrate theoretically and experimentally that the interference of electric and magnetic optically-induced modes in individual subwavelength silicon nanodisks can lead to the suppression of resonant backscattering and to enhanced resonant forward scattering of light. To this end we spectrally tune the nanodisk's fundamental electric and magnetic resonances with respect to each other by a variation of the nanodisk aspect ratio. This ability to tune two modes of different character within the same nanoparticle provides direct control over their interference, and, in consequence, allows for engineering the particle's resonant and off-resonant scattering patterns. Most importantly, measured and numerically calculated transmittance spectra reveal that backward scattering can be suppressed and forward scattering can be enhanced at resonance for the particular case of overlapping electric and magnetic resonances. Our experimental results are in good agreement with calculations based on the discrete dipole approach as well as finite-integral frequency-domain simulations. Furthermore, we show useful applications of silicon nanodisks with tailored resonances as optical nanoantennas with strong unidirectional emission from a dipole source.
Article
Full-text available
Subwavelength dielectric structures offer an attractive low-loss alternative to plasmonic materials for the development of resonant optics functionalities such as metamaterials and optical antennas. Non-spherical like rectangular dielectric structures are of the most interest from the standpoint of device development due to fabrication convenience. However, no intuitive fundamental understanding of the optical resonance in non-spherical dielectric structures is available, which has substantially delayed the development of dielectric resonant optics devices. Here we elucidate the general fundamentals of the optical resonance in non-spherical subwavelength dielectric structures with different shapes (rectangular or triangular) and dimensionalities (1D nanowires or 0D nanoparticles). We demonstrate that the optical properties of non-spherical dielectric structures are dictated by the eigenvalue of the structure's leaky modes. Leaky modes are defined as optical modes with propagating waves outside the structure. We also elucidate the dependence of the modal eigenvalue on physical features of the structure. The eigenvalue shows scale invariance with respect to the size of the structure, weak dependence on the refractive index, but linear dependence on the size ratio of different sides of the structure. We propose a modified Fabry-Perot model to account for the linear dependence. The knowledge of leaky modes, including the role in optical responses and the dependence on physical features, can serve as a powerful guide for the rational design of devices with desired optical resonances. It may open up a pathway to design devices with functionality that has not been explored due to the lack of intuitive understanding, for instance, imaging devices able to sense incident angle, or superabsorbing photodetectors.
Article
Full-text available
Directional light scattering by spherical silicon nanoparticles in the visible spectral range is experimentally demonstrated for the first time. These unique optical properties arise because of simultaneous excitation and mutual interference of magnetic and electric dipole resonances inside a single nanosphere. Such behaviour is similar to Kerker's-type scattering by hypothetic magneto-dielectric particles predicted theoretically three decades ago. Here we show that directivity of the far-field radiation pattern of single silicon spheres can be strongly dependent on the light wavelength and the nanoparticle size. For nanoparticles with sizes ranging from 100 to 200 nm, forward-to-backward scattering ratio above six can be experimentally obtained, making them similar to 'Huygens' sources. Unique optical properties of silicon nanoparticles make them promising for design of novel low-loss visible- and telecom-range metamaterials and nanoantenna devices.
Book
Nano-optics is the study of optical phenomena and techniques on the nanometer scale, that is, near or beyond the diffraction limit of light. It is an emerging field of study, motivated by the rapid advance of nanoscience and nanotechnology which require adequate tools and strategies for fabrication, manipulation and characterization at this scale. In this 2006 text the authors provide a comprehensive overview of the theoretical and experimental concepts necessary to understand and work in nano-optics. With a very broad perspective, they cover optical phenomena relevant to the nanoscale across diverse areas ranging from quantum optics to biophysics, introducing and extensively describing all of the significant methods. Written for graduate students who want to enter the field, the text includes problem sets to reinforce and extend the discussion. It is also a valuable reference for researchers and course teachers.
Article
We present a nano-patterned dielectric coating for crystalline Si solar cells that combines excellent anti-reflection and passivation properties. The nano-patterned coating comprises an array of TiO2 nanocylinders placed on top of an ultra-thin Al2O3 layer on a flat Si(100) wafer. The antireflection effect stems from the preferential forward scattering of light through leaky Mie resonances in the TiO2 nanocylinders. The Al2O3 layer provides excellent passivation of the Si surface. We experimentally demonstrate ultralow surface recombination with carrier lifetimes above 4 ms, combined with a reflectivity of 2.8% averaged over a broad spectral range.
Article
Expressions for the total power radiated by magnetic and electric dipoles of arbitrary orientation located in a medium 1 at distance z0 from the interface to a homogeneous or planar stratified medium 2 are derived. A relation between the normalized powers radiated by magnetic and electric dipoles is established. For a homogeneous loss-free medium 2, curves of the normalized powers L(z0)/L ∞ radiated by magnetic and electric dipoles versus the normalized distance z0/λ1, are presented for different values of the relative refractive index n = n2/n1, as the only parameter. The computer calculations are compared with analytical expressions derived for small and large distances. For n > 1, the contribution of the evanescent waves to the radiated power is calculated separately. We show that the classical results for the normalized radiated power yield the correct normalized spontaneous emission rates from an excited atomic state for electric and magnetic dipole transitions, respectively. We point out that the results for the electric dipole also give the change of the total power scattered by a small dielectric scattering particle when it is placed close to an interface.
Article
We present the first experimental demonstration of zero backscattering from nanoparticles at op- tical frequencies as originally discussed by Kerker et. al. [M. Kerker, D. Wang, and C. Giles, J. Opt. Soc. A 73, 765 (1983)]. GaAs pillars were fabricated on a fused silica substrate and the spectrum of the backscattered radiation was measured in the wavelength range 600-1000 nm. Suppression of backscattering occurred at ~725 nm, agreeing with calculations based on the discrete dipole approximation. Particles with zero backscattering provide new functionality for metamaterials and optical antennas.