ArticlePDF Available

Near-Field Mediated Plexcitonic Coupling and Giant Rabi Splitting in Individual Metallic Dimers

Authors:

Abstract and Figures

Strong coupling between resonantly-matched localized surface plasmons and molecular excitons results in the formation of new hybridized energy states called plexcitons. Understanding the nature and tunability of these hybrid nanostructures is important for both fundamental studies and the development of new applications. We investigate the interactions between J-aggregate excitons and single plasmonic dimers and report for the first time a unique strong coupling regime in individual plexcitonic nanostructures. Dark-field scattering measurements and finite-difference time-domain simulations of the hybrid nanostructures show strong plexcitonic coupling mediated by the near-field inside each dimer gap, which can be actively controlled by rotating the polarization of the optical excitation. The plexciton dispersion curves, obtained from coupled harmonic oscillator models, show anticrossing behavior at the exciton transition energy and giant Rabi splitting ranging between 230-400 meV. These energies are, to the best of our knowledge, the largest obtained on individual hybrid nanostructures.
Content may be subject to copyright.
Near-Field Mediated Plexcitonic Coupling and Giant Rabi Splitting in
Individual Metallic Dimers
Andrea E. Schlather,
,
Nicolas Large,
,
Alexander S. Urban,
§,
Peter Nordlander,
,§,
and Naomi J. Halas*
,,,§,
Department of Chemistry,
Department of Electrical and Computer Engineering,
§
Department of Physics and Astronomy, and
Laboratory for Nanophotonics, Rice University, 6100 Main Street, Houston, Texas 77005, United States
*
SSupporting Information
ABSTRACT: Strong coupling between resonantly matched localized surface
plasmons and molecular excitons results in the formation of new hybridized energy
states called plexcitons. Understanding the nature and tunability of these hybrid
nanostructures is important for both fundamental studies and the development of
new applications. We investigate the interactions between J-aggregate excitons and
single plasmonic dimers and report for the rst time a unique strong coupling
regime in individual plexcitonic nanostructures. Dark-eld scattering measurements
and nite-dierence time-domain simulations of the hybrid nanostructures show
strong plexcitonic coupling mediated by the near-eld inside each dimer gap, which
can be actively controlled by rotating the polarization of the optical excitation. The
plexciton dispersion curves, obtained from coupled harmonic oscillator models, show anticrossing behavior at the exciton
transition energy and giant Rabi splitting ranging between 230 and 400 meV. These energies are, to the best of our knowledge,
the largest obtained on individual hybrid nanostructures.
KEYWORDS: Plexcitons, surface plasmons, molecular excitons, Rabi splitting, dark-eld spectroscopy, individual hybrid nanostructures
Future nanodevices will most likely include hybrid
nanostructures that combine the intrinsic properties of
dissimilar materials to forge new and interesting tunable
properties.
16
Although plexcitonics (i.e., the study of
plasmonexciton coupling) is still an experimentally and
theoretically challenging subject, several studies have reported
plexcitonic nanostructures composed of metallic nanostructures
interacting with molecular materials
613
and semiconduc-
tors.
1418
In a plexcitonic device, the localized surface plasmon
resonance (LSPR) of a metallic nanostructure couple with the
excitons of a complementary material. The resulting hybrid
nanostructures provide a uniquely adaptable platform for the
design and the implementation of functional optical devices at
the nanoscale. Important, already demonstrated applications
include chemical sensors,
19
pH meters,
20
light harvesting,
21
and
optically active devices.
22,23
The molecular complexes can also
be used to tune the optical properties of the metallic
nanostructure through a local modication of the dielectric
environment.
24
These are all examples of light-matter coupling
in the weak regime, where the LSPR modes are perturbed by
the presence of the molecule. Several concepts such as
hybridization,
25
Fano resonances,
26
and Rabi splitting
27
have
been successfully transferred from atomic and molecular
physics to describe analogue phenomena seen in plasmonic
systems. In the strong coupling regime,
713,2532
the coupling
between a molecular exciton and a plasmonic cavity results in
anticrossings of the hybrid plexciton dispersion curves and the
formation of two hybrid energy states separated by a Rabi
splitting energy. More recently, dynamic tuning
12,28
and
ultrafast manipulation
7,9,10
of the plexcitonic coupling have
been investigated in such hybrid nanostructures. However, all
of these studies have involved arrays of nanostruc-
tures,
7,9,1214,16,29
or ensembles of nanoparticles.
8,10,11,15,30
Previous studies have shown that coupling between
elementary excitations can be mediated by the near-eld and
in particular by the highly enhanced electric eld near the
surfaces of plasmonic nanostructures, known as hot
spots.
3135
The strong and localized eld in a plasmonic hot
spot enhances the interaction between LSPRs and the local
excitons, much as it enhances other molecular excitations.
Plasmonic dimers are the canonical geometry for the generation
of high-intensity hot spots, where the local eld is sucient to
give rise to surface enhanced Raman spectroscopy at the single
molecule level.
36
Plasmonic dimers have been widely studied
due to their simple geometry and the tunability of their far-eld
scattering properties in the visible/NIR range.
37
Here, we
investigate the strong coupling between individual nano-
structured plasmonic dimers and J-aggregate complexes. By
tuning the dimensions of a plasmonic dimer we are able to
create spectral overlap with the J-aggregate exciton transition as
well as large near-eld enhancements. This allows us to report
for the rst time a unique strong coupling regime in individual
plexcitonic nanostructures.
Received: April 24, 2013
Revised: May 21, 2013
Letter
pubs.acs.org/NanoLett
© XXXX American Chemical Society Adx.doi.org/10.1021/nl4014887 |Nano Lett. XXXX, XXX, XXXXXX
Results. The plasmonic dimers used to probe the LSPR-
exciton interaction were created using planar fabrication on an
ITO-coated SiO2wafer. The J-aggregates used to form our
structures have been shown to exhibit strong coupling in a
number of other plexcitonic systems,
8,10,12,13,38
due to their
narrow exciton transition linewidths and high oscillator
strengths at room temperature. J-aggregates were formed
from monomers of a cyanine dye in a polyvinyl alcohol
(PVA) matrix (Supporting Information S1). The J-aggregates
were then spin-cast onto the patterned ITO substrate, resulting
in a J-aggregate/PVA lm with a thickness of 1525 nm.
Scattering measurements of single hybrid nanostructures
were taken using a hyperspectral transmission dark-eld
microscope.
Plasmonic dimers have two distinct dipole LSP modes in the
longitudinal and transverse directions. Light scattered by each
LSP mode can be collected independently, depending on the
angle of the linear polarizer in the collection path. The
geometries of the Au dimers were chosen so that the
longitudinal plasmon resonance overlaps with the J-aggregate
exciton peak at 693 nm (1.79 eV). Individual nanodisk
diameters of 60, 70, 85, 100, and 115 nm were used, with a
xed gap size of 15 nm to ensure the creation of an intense hot
spot in each dimer gap. Scanning electron micrographs and
longitudinal scattering spectra of the bare dimers conrm the
nanostructure dimensions and spectral overlap of each dimer
with the J-aggregate exciton peak (Figure 1). The longitudinal
LSPR is red-shifted with respect to the uncoupled nanodisk
plasmon resonance due to hybridization between the nano-
disks,
25,39
and is accompanied by a large near-eld enhance-
ment in the dimer gap.
40
The transverse LSP mode gives rise to
a near-eld enhancement weaker by an order of magnitude and
a weaker hybridization between the nanodisks resulting in a
higher resonance energy that, in the case of the smaller dimers,
does not overlap with the J-aggregate exciton.
When the J-aggregate/PVA layer is added, the longitudinal
LSP resonance of each dimer is strongly modied (cf. Figure
2a). The single scattering peak splits into two separate peaks,
separated by a peak minimum at the exciton wavelength (693
nm) revealing a strong coherent coupling between the
longitudinal LSP of the dimer and the exciton. In contrast,
when the transverse LSP of the dimer is excited, the resonance
peak for each structure is notably blue-shifted. While the
spectral position of the smaller dimers has been detuned from
the J-aggregate exciton (cf. green spectra in Figure 2e,f), the
larger dimers still have a large degree of spectral overlap (cf.
black spectra in Figure 2e,f). However, no peak splitting is
observed with this polarization for any of the dimers in the
series, indicating a much weaker coupling between the
transverse LSPR and the J-aggregate excitons.
We used the nite-dierence time-domain (FDTD) method
to calculate the optical properties of the hybrid plexcitonic
nanostructures. The scattering proles for all hybrid dimers
were calculated for both longitudinal and transverse polar-
izations. Calculated spectra (Figure 2b,e) are in very good
agreement with the experimental spectra (Figure 2a,d).
Specically, the calculated longitudinally polarized spectra
exhibit a strong dip at 693 nm and two distinct plexciton
resonances: a higher energy mode on the blue side of the
exciton which will be referred to as the upper branch (UB) and
a lower energy mode on the red side referred to as the lower
branch (LB). In contrast, the transverse polarized spectra show
a very weak dip at 693 nm (not resolved experimentally).
Figure 2c,d displays the calculated near-eld enhancement
distributions |E/Eo|2for the longitudinal and transverse
polarizations, respectively. The largest near-eld enhancement
for the longitudinally polarized dimer occurs inside the dimer
gap, where the maximum enhancement is 250 for the largest
dimer in the series (Figure 2c). The transverse near-eld maps,
magnied by a factor of 10 for clear visualization, illustrate the
absence of any near-eld enhancement inside the dimer gap
(Figure 2d). The maximum transverse near-eld enhancement
occurs outside the gap and is no larger than 25 for the largest
dimer, that is, an order of magnitude smaller than the largest
longitudinal mode gap enhancement. This suggests that strong
near-eld enhancements in the center of the dimers are
essential for the plexciton formation seen in the longitudinally
polarized spectra (Figure 2a,b).
To further investigate the dependence of the coupling
behavior on the polarization-dependent near-eld enhancement
in the dimer gap, a spectrum was collected from each individual
nanostructure as the polarization angle was varied by 5°
increments. The linear polarizer in the collection path of the
microscope was rotated, and a spectrum was collected at each
angle from 0°to 90°, corresponding to the longitudinal
(horizontal arrow) and transverse (vertical arrow) polarizations,
respectively (Figure 3). The results for the 70 nm nanodisk
dimer,showninFigure3ain15°increments, show a
progressive emergence of the two plexcitonic peaks as the
polarization is rotated from transverse (90°) to longitudinal
(0°). Results from FDTD calculations show an identical trend
(Figure 3b). As the polarization angle decreases (polarization
changes from transverse to longitudinal), the UB plexciton
shifts to longer wavelengths, and the LB mode appears
gradually on the red side of the exciton transition (693 nm/
1.79 eV). The intensity of the LB peak increases as the
polarization angle approaches 0°, corresponding to the
maximum coupling. Each individual polarized spectrum was
Figure 1. Bare gold nanodisk dimers. (a) SEM images of ve gold
dimers with diameters increasing from top to bottom (diameters are
60, 70, 85, 100, and 115 nm). The interparticle gap is xed at 15 nm.
The scale bars correspond to 100 nm. (b) Longitudinally polarized
scattering spectra measured for the ve bare dimers. The exciton
absorption peak from the J-aggregate complex is shown as reference
(gray line).
Nano Letters Letter
dx.doi.org/10.1021/nl4014887 |Nano Lett. XXXX, XXX, XXXXXXB
Figure 2. Size dependence of hybrid nanodisk dimers. Scattering spectra of plexcitonic gold dimerJ-aggregate nanostructures with nanodisks
ranging from 60 to 115 nm in diameter recorded for (a) longitudinal and (e) transverse polarizations. Corresponding calculated spectra are shown
for (b) longitudinal and (f) transverse polarizations. The exciton transition (693 nm/1.79 eV) of the J-aggregate is indicated by the light blue vertical
line. Near-eld enhancement maps |E/E0|2calculated at the exciton energy are displayed in the center panels for both (c) longitudinal and (d)
transverse polarizations. The near-eld intensities calculated for the transverse excitation have been multiplied by a factor of 10.
Figure 3. Polarization dependence of plexcitonic properties on a single plasmonic dimer with individual disk diameters of 70 nm and a 15 nm gap.
(a) Polarized scattering spectra of plexcitonic dimer with detected polarization angles of 0°,15°,30°,45°,60°,75°, and 90°. (b) Scattering spectra
calculated for the same polarization angles. (c) Calculated near-eld enhancement |E/E0|2corresponding to increasing polarization angles.
Polarization directions are shown with the white arrows. (d) Shift of the UB mode as a function of the polarization angle (green dots). (e) Relative
scattering intensity of the LB mode as a function of the polarization angle (green dots). Both quantities are compared to the polarization dependence
of the near-eld intensity calculated in the dimer gap (blue line).
Nano Letters Letter
dx.doi.org/10.1021/nl4014887 |Nano Lett. XXXX, XXX, XXXXXXC
normalized to the maximum scattering amplitude of the UB, so
that the relative amplitudes of the LB could be compared (cf.
Figure 3a). Identical measurements were performed on dimers
with 85 and 100 nm disk diameters, which exhibit the same
polarization dependence (Supporting Information S2). The
polarization dependence of the UB shift (Figure 3d) and the
LB intensity (Figure 3e) are found to correlate very well with
the calculated near-eld enhancement. The polarization
dependence of the UB shift and LB intensity is most
pronounced around 4050°, where the near-eld intensity
varies most strongly and weakest around 0°and 90°, where the
intensity of the near-eld levels out. These similarities provide
strong evidence that the plexcitonic coupling is dependent on
the strength of the near-eld enhancement inside the dimer
gap.
Discussion. Hybridization diagrams are useful for visualiz-
ing the interaction between the modes in complex systems. For
longitudinal polarization, the coupling of the individual
nanodisk LSPRs leads to the formation of a bright bonding
and a dark antibonding dipolar dimer LSPR mode.
40
In Figure
4a, we show how the bright dipolar LSPR mode of the dimer
interacts with the exciton transition dipole of the J-aggregate
complex (Jex), leading to the formation of the two plexcitonic
UB and LB modes. For a symmetry broken dimer, the dark
LSPR mode could also couple to the J-aggregate
37,41
but due to
large energy detuning between the exciton and the dark LSPR
mode of the dimer, we rule out the possibility of the exciton
coupling to the higher order plasmon modes.
The interaction between the bonding dimer LSPR and the
excitons of the J-aggregate results in hybridized plexciton states
which exhibit typical anticrossing behavior. The energies of the
UB and LB plexciton states are calculated using a coupled
harmonic oscillator model,
42
ω
ωω
ωω
ℏ=
ℏ+
±ℏΩ+
E() 2
1
2()( )
plexiton
UB,LB p
p0
R2p0
2
where ω0and ωpare the uncoupled exciton and dimer LSPR
energies, respectively. The coupling energy, ΩR, also called
Rabi splitting, is given by the spatial overlap of the excitonic
transition dipole moment μ0(r) and the induced surface
plasmon electric eld Ep(r): ΩR=μ0(r)·Ep(r)dV. Previous
works have reported splitting energies ranging from few
millielectron volts to several hundreds of millielectron
volts.
7,8,10,13,16,28,4346
By extracting the LSPR energy from
the scattering spectra of the bare nanodisk dimers and the
measured UB and LB modes as shown in Figure 4b, the Rabi
splitting is found to be approximately 230 meV which is in
good agreement with the highest values previously reported.
Moreover, for several hybrid dimer nanostructures (cf. Figures
3a and 4b) we were able to observe even larger coupling
leading to giant Rabi splitting energies of 400 meV
comparable to the largest values previously reported in
literature (Supporting Information S3). Energy transfer (i.e.,
Fano resonance and Rabi splitting) and plasmonic splitting
have been shown to occur in dierent coupling regimes dened
by the relative resonance linewidths (γLSPR and γ0) and by the
oscillator strength, f, of the excitonic resonance. Plasmonic
splitting is observed with large oscillator strengths ( f> 2) and
large molecular resonance linewidths (γ0> 200 meV).
47
In our
case f= 0.4 and γ0= 52 meV are indicative of an energy
transfer. Moreover, the transition from Fano resonance to Rabi
splitting has also been theoretically investigated.
27
It has been
shown that Rabi splitting occurs when ΩR>(γLSPR γ0)/2.
Our hybrid system satises this criterion, as (γLSPR γ0)/2
160 meV.
The large Rabi splitting energies in these hybrid nanostruc-
tures arise from alignment of the J-aggregate transition dipole
moment with the polarized near-eld inside the dimer
junction.
11
When the J-aggregate complex is moved away
from the center of the gap, the plexcitonic coupling becomes
strongly suppressed (Supporting Information S4). If the J-
aggregate transition dipole is moved only 30 nm away from the
gap center, the near-eld is substantially weaker, and no
measurable Rabi splitting can be detected.
The position of the J-aggregate complex relative to the dimer
junction is vital to obtain large plexcitonic coupling energies. In
order to obtain better control of the J-aggregate size and
location, an alternate hybrid dimer geometry was fabricated.
Instead of being cast from a PVA solution, the J-aggregates
were allowed to self-assemble on the surface of the metallic
disks from a solution of dye monomers.
8
In this case, J-
aggregates cover the entire surface of the metallic disk,
48
including the surface near the dimer gap. Adsorption of dye
molecules on the metal nanodisk surface occurs through a
combination of van der Waals forces and electrostatic attraction
between the Au and the positively charged nitrogen atoms on
the benzothiazole rings.
8
The UB and LB energies were
extracted from the measured longitudinal scattering spectra
(Supporting Information S5) and overlaid with the calculated
dispersion curve and the rst data set (Figure 4b). The Rabi
splitting energies of this second set (stars) are smaller than the
Figure 4. (a) Hybridization energy diagram of the plexcitonic dimer.
(b) Dispersion curves of the hybrid plexcitonic states extracted from
experimental data (stars and diamonds) and calculated from the
equation in the text (blue and red lines). Experimental points come
from two sets of samples: a thin J-aggregate layer formed at the surface
of the gold dimers (stars) (Supporting Information S5), and J-
aggregates formed in the PVA covering the dimers (diamonds). The
black and green dashed lines represent the uncoupled exciton and
surface plasmon energies, respectively.
Nano Letters Letter
dx.doi.org/10.1021/nl4014887 |Nano Lett. XXXX, XXX, XXXXXXD
rst set (diamonds) and more closely follow the theoretical
plexciton dispersion curves. This observation can be explained
by considering the position of the J-aggregates on the metal
surface. In the self-assembly method, only a small fraction of
the J-aggregates form in the nanodisk region that experiences
strong near-elds, that is, inside the dimer gap. Most of the J-
aggregate complexes form on the Au surfaces outside of the
dimer gap region and do not contribute to plexcitonic coupling.
While not all Au dimers showed strong coupling with J-
aggregates formed in solution (J-aggregates did not always
successfully align inside the dimer gaps), the coupling energies
measured for these structures (diamonds, Figure 4b) were
overall larger than that obtained when the J-aggregates self-
assembled directly at the nanodisk surface (stars, Figure 4b),
due to greater interaction with the intense near-eld.
Moreover, the high quality of the J-aggregates allows us to
perform a direct and rigorous comparison of the coupling
intensities for dierent gap sizes. For hybrid dimers with gap
sizes exceeding 25 nm, the plexcitonic coupling was found to be
too weak to induce signicant Rabi splitting. Dimers with gap
sizes of 30 and 60 nm showed no peak splitting with either J-
aggregates formed on the surface or in solution (Supporting
Information S6). In both cases, only weak coupling was evident,
as determined by the LSPR shift.
49
The reason for the weak
coupling for these larger gap dimers is the reduced near-eld in
the junction, the near-eld intensity being three times smaller
than that of the 15 nm gap dimer.
In conclusion, we have reported a rigorous investigation of
plexcitonic interactions between localized surface plasmon
resonances in individual hybrid metallic nanodisk dimers and J-
aggregate excitons. We have shown spatially resolved eld-
enhanced coherent coupling between two discrete optical
excitations, leading to a plexciton-induced transparency (i.e.,
almost complete suppression of far-eld scattering) in the
visible range. Careful geometrical design allowed a near-eld
enhanced plexcitonic coupling, giving rise to giant Rabi
splitting. Further engineering of these hybrid nanostructures,
by introducing symmetry breaking for instance, can lead to
comparably larger near-eld enhancements and introduce the
possibility of coupling excitons to higher order dark LSP
modes. The quantitative investigation of plexciton formation
we reported here with an unprecedented control in the single
particle regime opens up a new way to promising nanoscale
applications at optical frequencies.
Methods. Fabrication Process. The plasmonic dimers were
created using planar fabrication on an ITO-coated SiO2wafer.
Arrays of nanodisk dimers, spaced with a pitch of 10 μm, were
patterned by electron beam lithography using a 70 nm thick
poly(methyl methacrylate) positive resist (PMMA, 950 wt).
Electron beam evaporation was used to deposit a 2 nm Ti
adhesion layer followed by a 35 nm Au layer, where the layer
thicknesses were monitored via quartz crystal microbalance.
The excess metal and resist were removed by liftoin NMP (1-
methyl-2-pyrrolidone) at 65 °C for two hours to reveal the
nanostructures. Plasma cleaning (O2, 50 W, 100 mTorr) was
performed for one minute to remove traces of residual resist
around the nanostructures after lifto, as well as after scanning
electron microscope (SEM) imaging (FEI Quanta 650) to
remove carbon contamination from the electron beam.
J-aggregates were formed from monomers of a cyanine dye
(Supporting Information S1) by adding 5 μL aliquots of
monomers in ethanol (3 μg/mL) to 3 mL of a 5 mg/mL
polyvinyl alcohol (PVA) solution. UVvis spectroscopy was
used to monitor J-aggregate formation, which is indicated by a
narrowing and red shift of the monomer absorption peak from
599 to 693 nm. The J-aggregates were then spin-cast from
solution at 3000 rpm onto the patterned ITO substrate,
resulting in a J-aggregate/PVA lm with a thickness of 1525
nm, as determined by AFM measurements.
Hyperspectral Dark-Field Spectroscopy. Scattering meas-
urements were taken using a hyperspectral transmission dark-
eld microscope from Cytoviva HSI. In this conguration,
scattered light is collected over the entire sample surface area,
and each pixel of the image produced contains a complete
spectral prole. Unpolarized white light was focused onto the
sample through a high NA condenser lens at an angle varying
between 66°and 72°(corresponding to NA 1.21.4).
Scattered light from the nanostructures was collected by a
40×objective (NA 0.6) and passed through a 360°rotating
linear polarizer before entrance into a spectrograph and CCD
camera.
Finite-Dierence Time-Domain Simulations. We used the
nite-dierence time-domain (Lumerical Solutions) method to
calculate the optical properties of the hybrid plexcitonic
nanostructures. Geometrical parameters have been extracted
from SEM images of individual dimers in Figure 1. The dimers
were coated with a continuous layer of PVA and J-aggregates
and supported by an ITO-coated SiO2substrate. The bulk
dielectric function tabulated by Johnson and Christy was used
for Au,
50
and the dielectric function from Palik was used for
SiO2.
51
The J-aggregate complex was modeled as a dispersive
medium with dimensions set from approximated values
obtained through dynamic light scattering measurements (not
shown) of the J-aggregates in a PVA solution. In order to
account for the J-aggregate exciton, the dielectric permittivity of
the J-aggregate/PVA mixture has been described by the Lorentz
model as
ε
ωε
ω
ωω γω
=+
−−
f
i
() ()
J0
2
0
22
0
where ε= 2.5 is the high-frequency component of the J-
aggregate/PVA matrix dielectric function, f= 0.4 is the reduced
oscillator strength, ω0= 1.79 eV is the exciton transition
energy, and γ0= 52 meV is the exciton line width.
ASSOCIATED CONTENT
*
SSupporting Information
S1: J-aggregate formation and UVvis spectroscopy; S2:
polarization-dependent plexcitonic coupling for nanodisk
dimers (disk diameters D= 100 nm, D= 85 nm); S3: Rabi
splitting energy comparison with highest reported literature
values; S4: probing of near-eld mediated plexcitonic coupling
locality; S5: coupling between individual dimers and self-
assembled J-aggregate monolayers; S6: weak plexcitonic
coupling in large gap (g= 30 nm, g= 60 nm) nanodisk
dimers. This material is available free of charge via the Internet
at http://pubs.acs.org.
AUTHOR INFORMATION
Corresponding Author
*E-mail: halas@rice.edu.
Author Contributions
A.E.S. and N.L. contributed equally to this work. A.E.S. carried
out the fabrication of the nanostructures, optical measurements,
and data analysis. N.L. performed the theoretical study,
Nano Letters Letter
dx.doi.org/10.1021/nl4014887 |Nano Lett. XXXX, XXX, XXXXXXE
numerical simulations, and data analysis. A.S.U. optimized the
experimental data analysis process. N.J.H. and P.N. designed
the project. All of the authors discussed the results and wrote
the paper.
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
The authors wish to thank M. W. Knight, N. S. King, J. B.
Lassiter, Z. Fang, H. Sobhani, and N. T. Fofang for their insight
and input. This research was supported by the Robert A. Welch
Foundation under grants C-1220 and C-1222, the Oce of
Naval Research under grant N00014-10-1-0989, the DoD
NSSEFF (N00244-09-1-0067), and the U.S. Army Research
Laboratory and Oce under contract/grant number
WF911NF-12-1-0407.
REFERENCES
(1) Chen, H.; Shao, L.; Li, Q.; Wang, J. Chem. Soc. Rev. 2013,42 (7),
26792724.
(2) Tong, L.; Wei, H.; Zhang, S.; Li, Z.; Xu, H. Phys. Chem. Chem.
Phys. 2013,15 (12), 41004109.
(3) Sonnefraud, Y.; Leen Koh, A.; McComb, D. W.; Maier, S. A.
Laser Photonics Rev. 2012,6(3), 277295.
(4) Berkovitch, N.; Ginzburg, P.; Orenstein, M. J. Phys.: Condens.
Matter 2012,24 (7), 073202.
(5) Large, N.; Abb, M.; Aizpurua, J.; Muskens, O. L. Nano Lett. 2010,
10 (5), 17411746.
(6) Pérez-González, O.; Zabala, N.; Borisov, A. G.; Halas, N. J.;
Nordlander, P.; Aizpurua, J. Nano Lett. 2010,10 (8), 30903095.
(7) Vasa, P.; Pomraenke, R.; Cirmi, G.; De Re, E.; Wang, W.;
Schwieger, S.; Leipold, D.; Runge, E.; Cerullo, G.; Lienau, C. ACS
Nano 2010,4(12), 75597565.
(8) Fofang, N. T.; Park, T.-H.; Neumann, O.; Mirin, N. A.;
Nordlander, P.; Halas, N. J. Nano Lett. 2008,8(10), 34813487.
(9) Vasa, P.; Wang, W.; Pomraenke, R.; Lammers, M.; Maiuri, M.;
Manzoni, C.; Cerullo, G.; Lienau, C. Nat. Photonics 2013,7(2), 128
132.
(10) Fofang, N. T.; Grady, N. K.; Fan, Z.; Govorov, A. O.; Halas, N.
J. Nano Lett. 2011,11 (4), 15561560.
(11) Wiederrecht, G. P.; Wurtz, G. A.; Hranisavljevic, J. Nano Lett.
2004,4(11), 21212125.
(12) Zheng, Y. B.; Juluri, B. K.; Lin Jensen, L.; Ahmed, D.; Lu, M.;
Jensen, L.; Huang, T. J. Adv. Mater. 2010,22 (32), 36033607.
(13) Dintinger, J.; Klein, S.; Bustos, F.; Barnes, W. L.; Ebbesen, T. W.
Phys. Rev. B 2005,71 (3), 035424.
(14) Lawrie, B. J.; Kim, K. W.; Norton, D. P.; Haglund, R. F. Nano
Lett. 2012,12 (12), 61526157.
(15) Achermann, M. J. Phys. Chem. Lett. 2010,1(19), 28372843.
(16) Vasa, P.; Pomraenke, R.; Schwieger, S.; Mazur, Y. I.; Kunets, V.;
Srinivasan, P.; Johnson, E.; Kihm, J. E.; Kim, D. S.; Runge, E.; Salamo,
G.; Lienau, C. Phys. Rev. Lett. 2008,101 (11), 116801.
(17) Govorov, A. O.; Bryant, G. W.; Zhang, W.; Skeini, T.; Lee, J.;
Kotov, N. A.; Slocik, J. M.; Naik, R. R. Nano Lett. 2006,6(5), 984
994.
(18) Gómez, D. E.; Vernon, K. C.; Mulvaney, P.; Davis, T. J. Nano
Lett. 2009,10 (1), 274278.
(19) Murphy, C. J. Anal. Chem. 2002,74 (19), 520 A526 A.
(20) Bishnoi, S. W.; Rozell, C. J.; Levin, C. S.; Gheith, M. K.;
Johnson, B. R.; Johnson, D. H.; Halas, N. J. Nano Lett. 2006,6(8),
16871692.
(21) Govorov, A. O.; Carmeli, I. Nano Lett. 2007,7(3), 620625.
(22) Slocik, J. M.; Tam, F.; Halas, N. J.; Naik, R. R. Nano Lett. 2007,
7(4), 10541058.
(23) Artuso, R. D.; Bryant, G. W. Nano Lett. 2008,8(7), 2106
2111.
(24) Zhao, J.; Zhang, X.; Yonzon, C. R.; Haes, A. J.; Van Duyne, R. P.
Nanomedicine 2006,1(2), 219228.
(25) Prodan, E.; Radloff, C.; Halas, N. J.; Nordlander, P. Science
2003,302 (5644), 419422.
(26) Rahmani, M.; Lukyanchuk, B.; Hong, M. Laser Photonics Rev.
2013,7(3), 329349.
(27) Savasta, S.; Saija, R.; Ridolfo, A.; Di Stefano, O.; Denti, P.;
Borghese, F. ACS Nano 2010,4(11), 63696376.
(28) Wurtz, G. A.; Evans, P. R.; Hendren, W.; Atkinson, R.; Dickson,
W.; Pollard, R. J.; Zayats, A. V.; Harrison, W.; Bower, C. Nano Lett.
2007,7(5), 12971303.
(29) Bellessa, J.; Symonds, C.; Vynck, K.; Beaur, L.; Brioude, A.;
Lemaitre, A. Superlattices Microstruct. 2011,49 (3), 209216.
(30) Lekeufack, D. D.; Brioude, A.; Coleman, A. W.; Miele, P.;
Bellessa, J.; De Zeng, L.; Stadelmann, P. Appl. Phys. Lett. 2010,96
(25), 2531073.
(31) Large, N.; Saviot, L.; Margueritat, J. R. M.; Gonzalo, J.; Afonso,
C. N.; Arbouet, A.; Langot, P.; Mlayah, A.; Aizpurua, J. Nano Lett.
2009,9(11), 37323738.
(32) Yin, H.; Zhang, H.; Cheng, X.-L. J. Appl. Phys. 2013,113 (11),
1131076.
(33) Tripathy, S.; Marty, R.; Lin, V. K.; Teo, S. L.; Ye, E.; Arbouet,
A.; Saviot, L.; Girard, C.; Han, M. Y.; Mlayah, A. Nano Lett. 2011,11
(2), 431437.
(34) Manjavacas, A.; Abajo, F. J. G. a. d.; Nordlander, P. Nano Lett.
2011,11 (6), 23182323.
(35) Jinna, H.; Chunzhen, F.; Junqiao, W.; Pei, D.; Genwang, C.;
Yongguang, C.; Shuangmei, Z.; Erjun, L. J. Opt. 2013,15 (2), 025007.
(36) Michaels, A. M.; Jiang, J.; Brus, L. J. Phys. Chem. B 2000,104
(50), 1196511971.
(37) Lassiter, J. B.; Aizpurua, J.; Hernandez, L. I.; Brandl, D. W.;
Romero, I.; Lal, S.; Hafner, J. H.; Nordlander, P.; Halas, N. J. Nano
Lett. 2008,8(4), 12121218.
(38) Juluri, B. K.; Lu, M.; Zheng, Y. B.; Huang, T. J.; Jensen, L. J.
Phys. Chem. C 2009,113 (43), 1849918503.
(39) Alonso-González, P.; Albella, P.; Golmar, F.; Arzubiaga, L.;
Casanova, F.; Hueso, L. E.; Aizpurua, J.; Hillenbrand, R. Opt. Express
2013,21 (1), 12701280.
(40) Thomas, R.; Swathi, R. S. J. Phys. Chem. C 2012,116 (41),
2198221991.
(41) Brown, L. V.; Sobhani, H.; Lassiter, J. B.; Nordlander, P.; Halas,
N. J. ACS Nano 2010,4(2), 819832.
(42) Rudin, S.; Reinecke, T. L. Phys. Rev. B 1999,59 (15), 10227
10233.
(43) Bellessa, J.; Bonnand, C.; Plenet, J. C.; Mugnier, J. Phys. Rev.
Lett. 2004,93 (3), 036404.
(44) Bellessa, J.; Symonds, C.; Vynck, K.; Lemaitre, A.; Brioude, A.;
Beaur, L.; Plenet, J. C.; Viste, P.; Felbacq, D.; Cambril, E.; Valvin, P.
Phys. Rev. B 2009,80 (3), 033303.
(45) Cade, N. I.; Ritman-Meer, T.; Richards, D. Phys. Rev. B 2009,79
(24), 241404.
(46) Bonnand, C.; Bellessa, J.; Plenet, J. C. Phys. Rev. B 2006,73
(24), 245330.
(47) Chen, H.; Shao, L.; Woo, K. C.; Wang, J.; Lin, H.-Q. J. Phys.
Chem. C 2012,116 (26), 1408814095.
(48) Vujačić, A.; Vasić, V.; Dramićanin, M.; Sovilj, S. P.; Bibić, N.;
Hranisavljevic, J.; Wiederrecht, G. P. J. Phys. Chem. C 2012,116 (7),
46554661.
(49) Zheng, Y. B.; Kiraly, B.; Cheunkar, S.; Huang, T. J.; Weiss, P. S.
Nano Lett. 2011,11 (5), 20612065.
(50) Johnson, P. B.; Christy, R. W. Phys. Rev. B 1972,6(12), 4370
4379.
(51) Palik, E. D. Handbook of Optical Constants of Solids; Academic
Press: New York, 1985; Vol. 1.
Nano Letters Letter
dx.doi.org/10.1021/nl4014887 |Nano Lett. XXXX, XXX, XXXXXXF
... The single MB molecule in CB [7], i.e., CB [7] @single MB, was modeled as a dispersive medium cylinder with a diameter of 1.5 nm and a length of 0.9 nm, which equals the diameter and thickness of the molecular spacer of CB [7], respectively. The dielectric permittivity of the cylinder was described using the Lorentz model 59 : ...
Article
Full-text available
Experimental realization of strong coupling between a single exciton and plasmons remains challenging as it requires deterministic positioning of the single exciton and alignment of its dipole moment with the plasmonic fields. This study aims to combine the host–guest chemistry approach with the cucurbit[7]uril-mediated active self-assembly to precisely integrate a single methylene blue molecule in an Au nanodimer at the deterministic position (gap center of the nanodimer) with the maximum electric field (EFmax) and perfectly align its transition dipole moment with the EFmax, yielding a large spectral Rabi splitting of 116 meV for a single-molecule exciton—matching the analytical model and numerical simulations. Statistical analysis of vibrational spectroscopy and dark-field scattering spectra confirm the realization of the single exciton strong coupling at room temperature. Our work may suggest an approach for achieving the strong coupling between a deterministic single exciton and plasmons, contributing to the development of room-temperature single-qubit quantum devices.
... One particular research direction in quantum plasmonics has been the occurence of strong coupling between excitons and plasmonic excitations [11]. Strong coupling has already been reported for the coupling of surface plasmon polaritons with different types of emitters, such as J-aggregates [12][13][14][15][16][17][18][19][20], dye molecules [21][22][23] and quantum dots [24,25]. The new hybrid mode that arises from this strong coupling, the plexciton, is expected to be interesting for a variety of applications, such as light harvesting, light emitting devices and optical communications [26]. ...
Article
Full-text available
Strong coupling of quantum states with electromagnetic modes of topological matter offer an interesting platform for the exploration of new physics and applications. In this work, we report a novel hybrid mode, a surface topological plexciton, arising from strong coupling between the surface topological plasmon mode of a Bi2Se3 topological insulator nanoparticle and the exciton of a two-level quantum emitter. We study the power absorption spectrum of the system by working within the dipole and rotating-wave approximations, using a density matrix approach for the emitter, and a classical dielectric-function approach for the topological-insulator nanoparticle. We show that a Rabi-type splitting can appear in the spectrum suggesting the presence of strong coupling. Furthermore, we study the dependence of the splitting on the separation of the two nanoparticles as well as the dipole moment of the quantum emitter. These results can be useful for exploring exotic phases of matter, furthering research in topological insulator plasmonics, as well as for applications in the far-infrared and quantum computing.
Article
In this work, we use optical reflectance spectroscopy to study the plexciton formation between Ag nanoclusters and C60 molecules. The Ag clusters are fabricated on high-quality graphene under ultrahigh vacuum (UHV) and exhibit a strong absorption band at 3.4–3.6 eV due to the localized surface plasmons (LSPs). The plexciton formation is studied by depositing C60 on the Ag clusters in the same chamber under UHV. The deposition of C60 molecules leads to a splitting of the LSP band into multiple peaks with a systematic peak energy shift as a function of C60 coverage θC60. Notably, the details of the energy shifts and intensity variation sensitively depend on the LSP energy. Model calculations in which the plexciton coupling is approximated by a point-dipole interaction predict the evolution of the plexciton eigenenergy and oscillator strength with θC60 that reasonably explains the experiments. We demonstrate that spectral changes under the deposition of C60 cannot be explained by considering only one electronic transition in C60, as is commonly assumed. Instead, it is necessary to take into account three electronic transitions of C60 and their intermixing with the LSP to fully comprehend the spectral evolution. The plexciton coupling energy is estimated to exceed 0.6 eV, indicating that the system reaches a strong coupling regime. Our findings suggest the importance of simultaneously considering multiple electronic excitations of molecules to understand strongly coupled plexcitons.
Article
Full-text available
Plasmonic modes confined to metallic nanostructures at the atomic and molecular scale push the boundaries of light–matter interactions. Within these extreme plasmonic structures of ultrathin nanogaps, coupled nanoparticles, and tunnelling junctions, new physical phenomena arise when plasmon resonances couple to electronic, exitonic, or vibrational excitations, as well as the efficient generation of non-radiative hot carriers. This review surveys the latest experimental and theoretical advances in the regime of extreme nano-plasmonics, with an emphasis on plasmon-induced hot carriers, strong coupling effects, and electrically driven processes at the molecular scale. We will also highlight related nanophotonic and optoelectronic applications including plasmon-enhanced molecular light sources, photocatalysis, photodetection, and strong coupling with low dimensional materials.
Article
The room temperature strong coupling between the photonic modes of micro/nanocavities and quantum emitters (QEs) can bring about promising advantages for fundamental and applied physics. Improving the electric fields (EFs) by using plasmonic modes and reducing their losses by applying dielectric nanocavities are widely employed approaches to achieve room temperature strong coupling. However, ideal photonic modes with both large EFs and low loss have been lacking. Herein, we propose the abnormal anapole mode (AAM), showing both a strong EF enhancement of ∼70-fold (comparable to plasmonic modes) and a low loss of 34 meV, which is much smaller than previous records of isolated all-dielectric nanocavities. Besides realizing strong coupling, we further show that by replacing the normal anapole mode with the AAM, the lasing threshold of the AAM-coupled QEs can be reduced by one order of magnitude, implying a vital step toward on-chip integration of nanophotonic devices.
Article
Full-text available
The properties of organic semiconductor excitons in strong interaction with surface plasmons are described. For this purpose, a J -aggregated cyanine dye layer is deposited on a silver film and studied with reflectometry experiments. The dependence of the Rabi splitting with the square root of the dye layer absorption and the width inversion of the reflectometry lines at the resonance are observed. The experimental spectra are well fitted using a transfer matrix method with a complex index. The comparison between experiments and simulations for layers in different configurations is qualitatively related to the role of the plasmon polarization in the interaction with the aggregated dye chains. Finally, large Rabi splitting of 300meV are observed for pure dye layers. Characteristic behavior of the large coupling and the effect of asymmetrical absorption line shape are analyzed.
Article
Full-text available
We report on the resonant coupling between localized surface plasmon resonances (LSPRs) in nanostructured Ag films and an adsorbed monolayer of Rhodamine 6G dye. Hybridization of the plasmons and molecular excitons creates new coupled polaritonic modes, which have been tuned by varying the LSPR wavelength. The resulting polariton dispersion curve shows an anticrossing behavior which is very well fit by a simple coupled-oscillator Hamiltonian, giving a giant Rabi-splitting energy of ˜400meV . The strength of this coupling is shown to be proportional to the square root of the molecular density. The Raman spectra of R6G on these films show an enhancement of many orders of magnitude due to surface enhanced scattering mechanisms; we find a maximum signal when a polariton mode lies in the middle of the Stokes shifted emission band.
Article
In this review, the most recent progress in the development of noble metal nano-optical sensors based on localized surface plasmon resonance (LSPR) spectroscopy is summarized. The sensing principle relies on the LSPR spectral shifts caused by the surrounding dielectric environmental change in a binding event. Nanosphere lithography, an inexpensive and simple nanofabrication technique, has been used to fabricate the nanoparticles as the LSPR sensing platforms. As an example of the biosensing applications, the LSPR detection for a biomarker of Alzheimer's disease, amyloid-derived diffusable ligands, in human brain extract and cerebrospinal fluid samples is highlighted. Furthermore, the LSPR sensing method can be modified easily and used in a variety of applications. More specifically, a LSPR chip capable of multiplex sensing, a combined electrochemical and LSPR protocol and a fabrication method of solution-phase nanotriangles are presented here.
Article
This review article focuses on the basic physics of LSPR modes, and how they can be observed. For dipolar modes, observation is rather straightforward. However, higher order modes often require the use of more advanced experimental conditions or dedicated spectroscopic techniques such as electron energy-loss spectroscopy (EELS). Eventually, bespoke LSPR modes can be engineered when different cavities are brought together to interact, giving rise to super- or sub-radiant modes, as well as Fano resonances, which in the right conditions can evolve into plasmonic induced transparency.
Article
Recently, a large number of experimental and theoretical works have revealed a variety of plasmonic nanostructures with the capabilities of Fano resonance (FR) generation. Among these structures, plasmonic oligomers consisting of packed metallic nanoelements with certain configurations have been of significant interest. Oligomers can exhibit FR independently of the polarization direction based on dipole–dipole antiparallel modes without the need to excite challenging high-order modes. The purpose of this review article is to provide an overview of recent achievements on FR of plasmonic nanostructures in recent years. Meanwhile, more attention is given to the optical properties of plasmonic oligomers due to the high potential of such structures in optical spectra engineering.
Article
The optical properties of asymmetric ring structures are investigated theoretically by using the discrete dipole approximation method. The numerical results revealed that this kind of structure can achieve a giant localized field enhancement (LFE, 264) and a high LSPR sensitivity (corresponding FOM, 8.28) in the visible spectrum by Fano resonance, whose origin is discussed based on plasmon hybridization theory. Furthermore, the dependence of the Fano resonance on the polarization states of the incident light is also demonstrated. Giant LFE and high LSPR sensitivity enable this structure to be promising for surface enhanced Raman spectroscopy and sensing applications.
Article
The optical constants n and k were obtained for the noble metals (copper, silver, and gold) from reflection and transmission measurements on vacuum-evaporated thin films at room temperature, in the spectral range 0.5-6.5 eV. The film-thickness range was 185-500 Å. Three optical measurements were inverted to obtain the film thickness d as well as n and k. The estimated error in d was ± 2 Å, and that in n, k was less than 0.02 over most of the spectral range. The results in the film-thickness range 250-500 Å were independent of thickness, and were unchanged after vacuum annealing or aging in air. The free-electron optical effective masses and relaxation times derived from the results in the near infrared agree satisfactorily with previous values. The interband contribution to the imaginary part of the dielectric constant was obtained by subtracting the free-electron contribution. Some recent theoretical calculations are compared with the results for copper and gold. In addition, some other recent experiments are critically compared with our results.
Article
Plasmon resonances and the plasmon-induced field enhancement (FE) in sodium nanoring dimers are investigated by time-dependent density functional theory. For larger separations, the optical absorption, the induced charge response and the frequency dependent current demonstrate that there are two capacitive coupling plasmon modes. One feature of FE is that, in the surface region of the nanoring, it has a very large maximum. Another feature of FE is that, along the perpendicular bisector of the line segment joining the two nanoring center points in the middle region of the nanoring dimers, it has maxima. With the decrease of the gap distance, because of the electrons tunneling across the dimer junction and screening, collective excitation modes are changed, and the charge transfer plasmon modes emerge in the nanoring dimers. FE induced by any plasmon modes decreases in the gap region. Moreover, corresponding to different gap distances, the high-energy plasmon resonance peak almost does not shift, because this plasmon mode is mainly the collective excitation as a result of interactions among degenerate individual electronic states.
Article
Plasmonic–molecular resonance coupling was systematically studied using quasistatic approximation, Mie theory, and rigorous finite-difference time-domain calculations. The results indicate that the two types of coupling behaviors, plasmonic splitting and energy transfer, which are commonly manifested in experiments as peak splitting and a quenching dip, respectively, can be unified by considering a Au nanocrystal core coated with dye molecules. The dye coating is treated as a dielectric shell with Lorentzian-type absorption. By varying the oscillator strength and molecular transition line width, either plasmonic splitting or a quenching dip can be observed on the scattering spectrum of the dye-coated Au nanocrystal. The effects of the thickness of the dye coating, the spacing between the dye shell and the Au core, the partial dye coating, and the Au core shape on the coupled spectral shape were also ascertained. Our results will be useful for further exploring new phenomena in plasmon-based light–matter interactions as well as for developing highly selective and sensitive detection devices on the basis of plasmonic–molecular resonance coupling.