You are on page 1of 8

Applied Surface Science 314 (2014) 807814

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Tailoring of UV/violet plasmonic properties in Ag, and Cu coated Al


concaves arrays
a

Magorzata Norek a, , Maksymilian Wodarski b , Wojciech J. Stepniowski


a
Department of Advanced Materials and Technologies, Faculty of Advanced Technologies and Chemistry, Military University of Technology,
Str. Kaliskiego 2, 00-908 Warszawa, Poland
b
Institute of Optoelectronics, Military University of Technology, Str. Kaliskiego 2, 00-908 Warszawa, Poland

a r t i c l e

i n f o

Article history:
Received 13 May 2014
Accepted 30 June 2014
Available online 8 July 2014
Keywords:
UV plasmonics
Aluminum concaves
Anodization
Copper
Silver
Reectivity spectra

a b s t r a c t
UV plasmonics is of particular interest because of large variety of applications, where the higher energy
plasmon resonances would advance scientic achievements, including surface-enhanced Raman scattering (SERS) with UV excitation, ultrasensitive label-free detection of important biomolecules absorbing
light in the UV, or the possibility for exerting control over photochemical reactions. Despite its potential,
UV plasmonics is still in its infancy, mostly due to difculties in fabrication of reproducible nanostructured materials operating in this high energy range. Here, we present a simple electrochemical method to
fabricate regular arrays of aluminum concaves demonstrating plasmonic properties in UV/violet region.
The method enables the preparation of concaves with well-controlled geometrical parameters such as
interpore distance (Dc ), and therefore, well controllable plasmon resonances. Moreover, the patterning
is suitable for large scale production. The UV/violet properties of Al concaves can be further ne-tuned
by Ag and Cu metals. The refractive index sensitivity (RIS) increases after the metals deposition as compared to RIS of pure Al nanohole arrays. The highest RIS of 404 nm/RIU was obtained for Cu coated Al
nanoconcaves with the Dc = 460.8 nm, which is similar or better than the RIS values previously reported
for other nanohole arrays, operating in visible/near IR range.
2014 Published by Elsevier B.V.

1. Introduction
The interest in understanding of light interaction with ordered
arrays of nanoholes in thin metal lms has been a topic of research
for number of years [14]. Surface plasmons (SPs) are essentially
the coupled oscillations of light and free electrons at the interface
between a metal, which has a negative dielectric constant, and a
positive dielectric material. They are classied as surface plasmon
polaritons (SPPs) [5,6] and localized surface plasmons (LSPs) [7].
While in metal nanoparticles of different size and shape only LSPs
is present [810], tunable and regular metal nanovoids can support
both types of resonances depending on their truncation [1114]. It
was observed that rectangular holes are considerably better than
circles or squares due to larger LSPs contribution [15].
Optimal plasmonic properties (the strongest and narrowest
resonances) are provided by metals with small imaginary part
of the dielectric constant [1618]. Most of the metals possess

Corresponding author. Tel.: +48 504628903.


E-mail addresses: mnorek73@gmail.com, mnorek@wat.edu.pl (M. Norek).
http://dx.doi.org/10.1016/j.apsusc.2014.06.192
0169-4332/ 2014 Published by Elsevier B.V.

inter- and intraband transitions that increase the imaginary part


of their dielectric constants. This phenomenon causes weaker and
broader plasmonic resonances in metals such as Ru, Pd, Pt or Ni
[1922]. Because of the transitions, gold (Au), silver (Ag), and copper (Cu) do not generate plasmon resonance in UV range. On the
other hand, Au, Ag, Cu metals demonstrate excellent plasmonic
performance in the visible and near-IR region [2332] and their
plasmonic properties have been already exploited in many technological devices such as ultra-sensitive detectors, eld-enhanced
uorescence spectroscopy, or surface enhanced Raman spectroscopy (SERS) [3336]. There are, however, important upcoming
applications which stipulate the research on searching for materials with optimal plasmonic performance in the ultraviolet (UV)
range. UV plasmonics can be exploited in UV surface enhanced
Raman scattering spectroscopy (UV SERS), where scattering efciency is greatly improved. Biomolecules such as proteins and
nucleic acids contain residues that absorb light in UV. Plasmonic
structures operating in UV could enable the label-free detection
of those biomolecules. SPR in UV region could be used to enhance
light extraction in blue/UV light emitted devices (LEDs). Last but
not least, potential applications UV plasmonics include solar cells,
photocatalysis, or UV optical waveguides.

808

M. Norek et al. / Applied Surface Science 314 (2014) 807814

Aluminum (Al) can be the material of choice for plasmonic


application in the UV region. Al has an interband transition near
1.4 eV [16], what does not affect its UV performance. The optical
spectra of Al show well-dened resonance peaks in UV or even
deep UV, which additionally are very sensitive to size and shape
of nanostructures [3739]. Though aluminum is easily oxidized,
the process is self-limiting giving a thin and stable oxide thickness
which makes it easy to work with even in atmospheres containing oxygen [4042]. Aluminum can exhibit strongly enhanced local
elds owing to its high electron density (3 conductive electron per
atom in contrast to 1 electron per atom in metals such as Au or
Ag), which was used in plasmonic nanoantennas for uorescent
enhancement in UV [43,44]. Moreover, Al is abundant and cheaper
than most other metals, which can be very appealing for sustainable
photovoltaics.
SPs are very sensitive to the near-surface dielectric constant
(index of refraction), what has been used in a wide variety of SPRbased sensors. In order to efciently utilize the plasmonic eld
enhancement, careful matching of plasmon resonance in a given
sensing medium to a particular wavelength of a target material is
necessary. In free-electron metallic nanoparticles SPRs can be tuned
by changing size and shape of the nanoparticles, or by preparation of hybrid or alloyed nanostructures [45]. In nanohole array the
SPR depends on the pitch size, but is also susceptible to change of
the holes size and shape [11,46]. In order to improve the sensing
performance and broaden the application of nanoplasmonics other
approaches, including modeling of complex structures with multiple SPRs [47], or development of other sensing mechanisms [48],
were proposed as well. Recently, it was demonstrated that coating of nanohole arrays in metallic thin lm with another plasmonic
metal offer better plasmonic performance and can be used for ne
tuning of SPRs aiming to excite different uorescent dyes [49].
Current progress in nanoplasmonics has been also focused
on searching for less-expensive fabrication methods. Typically, ion beam or electron beam lithography (EBL) is used
[4,13,15,24,25,27,34,40] to produce nanohole arrays. These methods are, however, not well suited for large-scale production (the
typical area of the periodic nanostructure is about 100 m). In this
work, we show that regular, hexagonally arranged Al nanoconcaves, produced by a well-known electrochemical method, can be
a promising material for plasmonics operating in UV/violet range.
The preparation method is easy, reproducible, and the nanoconcaves can be synthesized on arbitrarily large area (20 mm2 in this
paper). The material can be, thus, applied for high-throughput
based detections. Geometrical parameters, and therefore the optical properties of the Al hole arrays, can be easily controlled by
changing the operating conditions during electrochemical synthesis, such as voltage, temperature or time of the rst anodization.
Further tuning of plasmonic behavior can be achieved by coating of
the Al concaves with a thin layer of other plasmonic metals, such
as Ag, and Cu.

process. Two samples (Al concaves) were prepared. To obtain the


rst sample (average interpore distance of 246.3 nm), hard anodization (HA) was applied according to Lee et al. [50]: the sample was
rst pre-anodized in 0.3 M oxalic acid solution at 40 V and 0 C for
10 min. Then, the voltage was gradually increased to 120 V, and the
anodization was carried out for 1 h. The anodization of the second
sample, with an average interpore distance of 456.7 nm, was performed in a mixture of phosphoric acid solution (0.1 M), water and
glycol (3:1, v/v), at temperature of 4 C, and for 20 h. As-obtained
alumina was chemically removed in a mixture of 6 wt% phosphoric
acid and 1.8 wt% chromic acid at 60 C for 120 min.
Morphology of nanoporous aluminum and microanalysis of
chemical composition of the samples coated with various metals
was studied using eld-emission scanning electron microscope FESEM (FEI, Quanta) and with Carl Zeiss Leo 1530 FE-SEM equipped
with energy dispersive X-ray spectrometer (EDS). BSE images were
taken with energy selective backscattered (EsB) detector at low
(3 kV) acceleration voltage.
To obtain geometrical parameters of the fabricated Al concaves
Fast Fourier transforms (FFTs) were generated based on three SEM
images taken at the same magnication for every anodizing voltage,
and were further used in calculations with WSxM software [51,52].
To estimate regularity ratio, three intensity proles were generated from each FFT image. The regularity ratio (RR) was estimated
according to the following formula (1) [53]:
RR =

H
W1/2 n

(1)

where n is the number of pores on the analyzed image, H the maximal intensity value of the FFT intensity prole, and W1/2 is the width
of the intensity prole at half of its height. Interpore distance was
estimated as an inverse of the FFTs radial average. The average
interpore distance (Dc ) was estimated from three FE-SEM images
for each sample. Pore diameter (Dd ) of the analyzed nanostructures
was estimated from three FE-SEM images for each operating conditions, using NIS-Elements software provided by Nikon Company.
Pores density (number of pores per 1 m2 ) was evaluated based
on six FE-SEM images for a given set of experimental conditions.
Detailed information on the FFTs image analysis can be found in
refs. [53,54].
The Ag, and Cu coating of as prepared Al concaves was done
by vacuum evaporation technique under high vacuum of 106 kPa
using resistively heated tungsten crucibles. The evaporation process was performed in the same experimental conditions for all
samples: the weights of metals were kept constant, the distance
from the crucibles to the substrates was always the same. The time
of the metals deposition lasted till the crucible was empty.
Reectivity measurement was performed using CCD spectrometer with ber-optics reection probe (Avantes) and deuterium/halogen light source (Ocean Optics Inc.). The probe was
set up at normal angle to measured sample. All reectance spectra
were collected at 2001100 nm wavelength range. Electropolished
Al coupon was used as a reference sample with 100% reection.

2. Experimental
3. Results and discussion
High-purity aluminum foil (99.9995% Al, Puratronic, AlfaAesar) with a thickness of about 0.25 mm was cut into coupons
(2 cm 2 cm). Before the anodization process the Al foils were
degreased in acetone and ethanol and subsequently electropolished in a 1:4 mixture of 60% HClO4 , and ethanol at 10 C, constant
current density of 0.5 A/cm2 , for 1 min. Next, the samples were
rinsed with distilled water, ethanol and dried. As prepared Al
coupons were insulated at the back and the edges with acid resistant tape. A Pt grid was used as a cathode, and the distance between
both electrodes was kept constant (ca. 1 cm). A two-electrode electrochemical cell with a Pt grid cathode was used in the anodizing

In Fig. 1 representative top view scanning electron microscopy


(SEM) images of the hemispherical Al concave arrays, synthesized
at 120 V (Fig. 1A) and at 195 V (Fig. 1B), are shown. After electrochemical anodization and thorough removal of the resulting oxide,
the periodic hexagonal hole arrays appear on the remaining Al foil.
The pores form a honeycomb structure, which is an exact replica of
the morphology of the AAO nanopore bottoms. Hexagonal shape of
pores is more regular in the case of the sample fabricated at 120 V
as compared to the one synthesized at 195 V. The size of the pores
increases with the applied voltage. On top of the Al concaves silver

M. Norek et al. / Applied Surface Science 314 (2014) 807814

809

Fig. 1. SEM micrographs (accelaration voltage of 3 kV) of the sample produced at 120 V (A) and 195 V (B), coated with Ag (C and D) and Cu (E and F) metals, respectively. The
scale bar applies to all images.

(Ag) and copper (Cu) metals were coated. Fig. 1(C) and (D) show
the top view SEM images of the sputtered Ag lm, demonstrating
that thin Ag lm is composed of tiny Ag nanoparticles which are dispersed on the cavities and the edges of the apertures of the concave
arrays. The thin Cu layer seems to uniformly cover the entire surface
of the concave arrays. In addition, some texture of the layer is visible
on the sample anodized at 195 V (Fig. 1E and F). The compositional
contrast of the samples conrmed the above observations (Fig. 1S in the Supplementary data). The sample covered with Ag shows
bright contrast coming from higher atomic number Ag, which corresponds to the Ag nanoparticles scattered on lower atomic number
Al nanoconcaves. On the other hand, there is no compositional contrast visible in the sample covered with Cu, giving indication that
the Al material is evenly coated by the Cu. The components of the Ag

and Cu coated Al nanobowl arrays are conrmed by energy dispersive X-ray spectroscopy (EDS) measurements (Fig. 2). In addition
to Al, Ag, and Cu elements, the presence of C and O were detected.
Oxygen comes from partial oxidation of the samples, which were
kept in air. The aluminum oxide with a thickness of 2.53.0 nm is
formed on the Al surface within few hours after exposure to air [41].
The process is self-limiting and its negative effect was not noticed
for the Al nanoholes arrays previously investigated [40,42]. The role
of Al2 O3 becomes important for very small nanoparticles (>10 nm).
It was observed that the Al2 O3 tends to red-shift the position of
SPRs in those tiny nanoparticles, but is rather undisruptive for the
extinction efciency [39,41]. Like aluminum, copper is prone to oxidation. However, some advantages of Cu surface oxidation, such as
enhancement of plasmonic response, were also noticed [55].

Fig. 2. EDS measurements for the samples coated with Ag (A), and with Cu (B). The analysis presents the results obtained for the Al concaves synthesized at 120 V.

810

M. Norek et al. / Applied Surface Science 314 (2014) 807814

Fig. 3. Lower magnication SEM micrographs (accelaration voltage of 20 kV) of the sample produced at 120 V (A) and 195 V (B) along with their respective FFT images (in
the upper, right corners).

Table 1
Pores diameter (Dp ), interpores distance (Dc ), regularity ratio (RR), and pores density
values for the samples anodized at 120 V and 195 V as determined by FFTs analysis.

120 V
195 V

Dp

Dc

RR

Pores density

188.3 16.4
356.6 27.0

238.2 0.1
460.8 11.6

1.84 0.29
1.60 0.40

14.8 1.4
4.2 0.2

To obtain geometrical parameters of pure Al nanoconcave


arrays, relatively large SEM image areas were used in Fast Fourier
transforms (FFTs) analysis (Fig. 3). The determined values are gathered in Table 1. The FFT image of the sample fabricated at 120 V
demonstrates six distinct points in the corners of hexagon, conrming good, hexagonal arrangement of pores. In the case of the
sample anodized at 195 V the points in the FFT image are more
blurred, suggesting lower degree of pores order. The degree of longrange alignment of nanopores is best reected in the regularity
ratio parameter (RR). The larger the RR, the better the alignment
[53,54]. The RR value of 1.84 0.29 was determined for the sample fabricated at 120 V, whereas for the sample prepared at 195 V
the RR value was 1.60 0.40. The worse pores arrangement in
the sample anodized at 195 V is most likely linked with smaller
domain size. Within a domain high local nanoconcaves regularity is observed. The domains are, however, separated by defects
and grain boundaries, where the regularities of pores drop signicantly. The bigger the domains, the higher the pores arrangement,
as indicated by the RR parameter. A domain size depends on volume expansion during the conversion of aluminum to alumina and
the experimental conditions (i.e. temperature, time of anodization) [56]. The volume expansion is strongly inuenced by a type
of used electrolyte. The larger the expansion the smaller the size
of the ordered domain [57]. During anodization process there is a
movement of negatively charged ions (OH , O2 , PO4 3 , C2 O4 2 ),
which are attracted to the positively polarized anode. The smaller
ions combine with Al3+ cations to form a nanoporous AAO. The
larger ions from electrolyte (PO4 3 , C2 O4 2 ) accumulate in the
walls of the growing oxide causing gradual deterioration of pores
arrangement [58]. The ionic mobility increases with the applied
potential. It can be, therefore, expected that the larger the ions and
the higher the applied voltages the stronger the deterioration. This
explains why anodization in phosphoric acid solution (relatively
large PO4 3 ions) and at relatively high voltages (195 V) results in
lower RR.

The resulting pores diameter (Dd ) and interpores distance (Dc ) determined by FFTs analysis are: 188.3 16.4 nm
and 238.2 0.1 nm for the sample anodized at 120 V, and
356.6 27.0 nm and 460.8 11.6 for the sample anodized at 195 V,
respectively. The increase of Dc is accompanied by the decrease
of pores density (number of pores per 1 m2 ): it decreases from
14.8 1.4 for sample anodized at 120 V down to 4.2 0.2 for the
sample anodized at 195 V (Table 1).
Investigations of plasmonic properties of materials usually
proceed by monitoring a dip in reectance (R), when an evanescent
light eld travels through a metal thin lm and excites SPs at the
metaldielectric interface. As a result, the normal reectivity of the
metal surface is greatly reduced on resonance due to optical absorption by the metal [2]. In Fig. 4 the reectivity spectra of the studied
samples are given. The overall reectivity of the patterned sample is
much lower than that of the continuous Al lm. At the same time in
the reectivity spectra pronounced minima are observed, the most
distinct one at a photon wavelength of 254 nm and 395 nm, for
the sample with Dc = 238.2 nm and Dc = 460.8 nm, respectively. The
reection intensity measured at the centre of the minima drops
down to 12.5% for the sample anodized at 120 V and to 2.7% for
the sample anodized at 195 V as compared to 100% reectivity of
unstructured Al (Table 2).
The minima are signatures of SPPs excitations, which are directly
connected to the symmetry of the lattice and the optical constants
of Al. At normal incidence, coupling photons of a given energy with
2-dimensional hexagonal periodic array gives SP resonances (SPRs)
at the following wavelength (Eq. (2)) [3,22]:
=

a0
4 2
(i
3

+ j2

+ ij)

m d
m + d

(2)

where a0 is the pitch size (interpore distance, Dc ), d is the


frequency-dependent permittivity of the dielectric material and m
is the real part of the frequency dependent permittivity of the metal
(Al in this case). The integers i, j specify the orders of SP resonances.
An iterative method was applied to solve Eq. (2), to account for
frequency dependent permittivity. For m calculation Drude model
was applied with constants provided by ref. [59]. The analytical
wavelength calculated for (i, j) = (0,1) or (1,0) ((0,1) ) was: 217,3 nm
for the sample with Dc = 238.2 nm, and 402.9 nm for the sample
with Dc = 460.8 nm. The theoretical (0,1) is pretty close to the measured reection minima in the sample anodized at 195 V, whereas

M. Norek et al. / Applied Surface Science 314 (2014) 807814

811

Fig. 4. Reectivity spectra of Al concaves fabricated at 120 V (A) and at 195 V (B) pure (red dashed lines) and coated with Ag (blue dotted lines) and Cu (magenta dash dot
lines) metals. For comparison, the reectivity of unstructured Al in also provided (black lines). The insets show larger magnication of the reectivity spectra demonstrating
the minima shift after the metals deposition. (For interpretation of the color information in this gure legend, the reader is referred to the web version of the article.)

Table 2
Position of reectivity minima ((0,1) SPR mode), reectance intensity at the (0,1) and refractive index sensitivity (RIS) values for the samples anodized at 120 V and 195 V.
Al concaves

Al concaves with Ag

Al concaves with Cu

120 V

(0,1) [nm]
Reectance at (0,1) [%]
RIS [nm/RIU] (the correlation coefcient)

254
12.5
189 (R2 = 0.9964)

259
6.2
191 (R2 = 0.9973)

262
1.7
199 (R2 = 0.9988)

195 V

(0,1) [nm]
Reectance at (0,1) [%]
RIS [nm/RIU] (the correlation coefcient)

395
2.7
296 (R2 = 0.9984)

407
7.1
399 (R2 = 0.9953)

425
2.2
404 (R2 = 0.9985)

is smaller than the minima related to the sample anodized at 120 V


(254 nm). The (0,1) SPR mode is shifted to higher wavelength
(lower energy) most probably due to additional contributions from
localized surface plasmons. The Al concaves possess very sharp
edges (Fig. 2-S in the Supplementary data), which can sustain
LSPs. The interaction between the LSPs and SPPs can generate a
strong plasmonic eld, which can add to or subtract from the
SPR. Apparently, the interaction lowered the plasmonic resonance
energy in the sample characterized by better pores arrangement
as determined by FFTs analysis (RR = 1.84). For the sample fabricated at 195 V, the potential LSPs effect is offset by poorer pores
arrangement (RR = 1.60) and relatively larger contribution from
grain boundaries. In the boundaries the shape of hexagons is very
irregular, or even pentagons and heptagons are present as an effect
of lattice defects, and therefore, the LSPs-SPPs interaction is effectively suppressed.
The (0,1) reectivity dips are slightly shifted to longer wavelengths after Ag and Cu metals deposition on the Al concaves (the
insets in Fig. 4). The largest shift of around 30 nm is observed for
Cu coated Al concaves with the pitch size of 460.8 nm (Table 2).
Additionally, the peak related to the samples covered by Cu (particularly to the one anodized at 120 V) becomes clearly broadened,
which is probably due to partial oxidation of Cu layer [28]. These
results indicate that the reectivity dip is dominated by the Al
concaves and is only ne-tuned by the thin Ag, and Cu coating.
It was previously observed that in order to get a substantial contrast between the surrounding metal and the holes, the metal must
be opaque (optically thick), which means that the layer thickness (or nanoholes depth) must be several times the skin-depth
of the metal [3,60]. The skin depth is the distance where the electric led falls to 1/e. Typical skin-depth for noble metals in the
visible spectrum is of 20 nm. When metal lm is very thin, it is partially transparent for the incident light, and no nanohole arrays is

necessary to obtain signicant reection drop, especially if the surface is resonantly corrugated [61]. The lm thickness formed by Ag
and Cu on the Al concaves can be considered as optically thin and,
therefore, plasmonic sensing performance of the combined metals
is basically determined by the aluminum nanoconcaves. Similar
results were obtained for hybrid plasmonic nanoparticles systems:
plasmonic resonances of copper- [62] and gold-shell [26] coated silver nanoparticles were only slightly shifted to longer wavelength
with respect to pure Ag NPs. Furthermore, thicker coating induced
stronger red-shift of LSPR peak. In contrast, dielectric coremetallic
shell spheroid nanoparticle demonstrated the LSPR shift to shorter
wavelength with the increase of the shell thickness [10]. Simulation results showed that various plasmonic resonance modes of
nanohole arrays in metallic lms (such as Ag, Al, Cu) are shifted
within 35 nm after coating them with Au thin layer (10 nm) [49],
which is in agreement with the results obtained in this work.
The intensity at the reectivity minima is further reduced after
the metals deposition (Table 2). For the Cu coated Al concaves
with Dc = 238.2 nm the reectivity as low as 1.7% is obtained, as
compared to unstructured, pure Al. These excellent antireecting
properties of regular, hexagonally arranged Al nanoconcaves can
be, thus, considered for applications in solar photovoltaics. In order
to increase absorption in silicon thin lms, texturing of the active
layer itself or application of various structures, including plasmonic
structures, have been widely investigated [6366].
SPR sensitivity of the coated and pure Al nanoconcave arrays was
measured as resonance wavelength shift per refractive index unit
(RIU). After immersing the samples in liquids with different refractive index the reectivity dips shift to longer wavelengths (Fig. 3-S
in the Supplementary data). In Fig. 5 (0,1) as a function of refractive index is presented, and refractive index sensitivity (RIS) values
are determined from the slope of linear t to the points measured
at the centre of the reectivity minima (Table 2). It can be seen

812

M. Norek et al. / Applied Surface Science 314 (2014) 807814

Fig. 5. Reectivity dips ((0,1) ) as Figs. 4 and 5 will appear in black and white in print and in color on the web. Based on this, the respective gure captions have been updated.
Please check, and correct if necessary a function of refractive index for Al concaves synthesized at 120 V (A), and at 195 V (B), pure (squares) and Ag (circles) and Cu (diamonds)
coated, along with the respective sensitivities predicted by SP theory (Eq. (2)) (black, dotted lines). The red, blue, and green lines represent linear ts to the measured data
to obtain refractive index sensitivity (RIS) values. (For interpretation of the color information in this gure legend, the reader is referred to the web version of the article.)

that RIS increases for Ag and Cu coated Al concaves as compared


to pure Al. For the sample with Dc = 238.2 nm, RIS changes from
189 nm/RIU for pure Al concaves to 199 nm/RIU for Cu coated Al
nanoconcaves arrays. The same trend can be noticed for the sample
with Dc = 460.8 nm: RIS 399 nm/RIU and 404 nm/RIU for Ag and
Cu coated Al concaves, respectively, as compared to 296 nm/RIU
for pure Al nanohole arrays. The increase of the plasmonic sensitivity was observed in Au coated Ag nanoparticles [26]. The plasmonic
eld enhancement was also demonstrated by simulation in Aucoated Ag, Cu and Al nanohole arrays in visible region [49]. Since
the metal coating does not change the Dc of the respective nanohole
arrays, the increase of RIS (as well as the (0,1) shift) may result
from the reduction of the holes size after metal coverage, or/and
from subsequent interface effects between two metals layers. Eq.
(2) neglects those effects, although there are strong experimental
premises that they affect the positions of SPRs [1115].
In Fig. 5 theoretical sensitivities for uncoated Al concaves,
obtained via solution of Eq. (2) by changing d values for different solvents, are also provided (dotted black lines). For both
samples the theoretical RIS is always higher than the experimental
one. It equals 230 nm/RIU for the sample with Dc = 238.2 nm and
427 nm/RIU for the sample with Dc = 460.8 nm. The disagreement
can be a consequence of domain-like morphology of Al concaves
and lessened pores arrangement outside the domains, giving rise to
strong scattering effects (energy losses produced by scattering from
roughness and irregularities on metal surface and at the concave
edges). Those effects are stronger when relative contribution from
grain boundaries with respect to that from domains is larger. Moreover, the bulk SPR sensitivity of nanohole arrays depends on holes
number contributing to the resonance. For nanohole arrays with a
low hole number the bandwidth is wider compared to nanohole
arrays with a large hole number [67,68]. Wider bandwidth results
in lower bulk SPR sensitivity. In this work the effect of holes number can also be noticed: the sample anodized at 120 V with larger
pores density (14.8 m2 ) demonstrates narrower bandwidths of
reectivity minima as compared to the Al concaves synthesized
at 195 V (pores density of 4.2 m2 ). In other words, a number of
holes contributed to the SPR were lower in the case of the second
sample, which might have resulted in lower than expected SPR sensitivity (bigger difference between theoretical and experimental
RIS). Besides, previously it was demonstrated that the RIS depends
strongly on whether it is measured at the transmission dip (thus
reection peak) or the transmission peak (reection dip) [13]. From

nite-difference time-domain (FDTD) calculations, the optical pattern at the transmission peak showed extensive surface plasmons
eld near the aperture edges (so-called short-range SPR), which
resulted in a small sensitivity, much smaller than the theoretical
one. Since in this work the sensitivity was determined from the
reection dips (because they were better discernible), the wavelength sensitivity was mainly constituted by the short-range SPR.
As an effect, lower than theoretically expected RIS was obtained for
both samples. The experimental RIS becomes closer to the theoretical one after the deposition of Ag and Cu metals. Particularly for Cu
coated Al concaves with Dc = 460.8 nm it increases up to 404 nm/RIU
as compared to 296 nm/RIU determined for pure Al holes arrays.
This indicates that the metals coating has a positive effect on plasmonic eld enhancement and can be used for tailoring of plasmonic
performance of Al concave arrays in UV region.
The RIS values evaluated in this work are comparable or even
better than previously determined for various nanohole arrays
operating in visible/near IR range. The RIS of 252 nm/RIU was
measured for Ag nanohole arrays prepared by colloidal lithography [32], 167 nm/RIU and 286 nm/RIU for periodic Au nanohole
arrays fabricated by EBL [27] and soft interference lithography
[69], respectively, and 212 nm/RIU for Ag nanoparticles prepared
by nano-sphere lithography [8]. Au nanohole arrays with the
lattice parameter (periodicities) of 590 nm demonstrated RIS of
400 nm/RIU [25], whereas RIS 409 nm/RIU was obtained for 500nm-period Au nanhole arrays [23].
It is anticipated that upon optimization of electrochemical process performed at high voltages the sensitivity performance of
the Al nanoconcave arrays, particularly those fabricated at 195 V
with the pitch size of 500 nm, will be better. As an example, Al
nanohole square arrays with similar periodicity fabricated by EBL
and nanoimprint lithography (NIL) techniques provided more regular nanohole arrays, which, in turn, resulted in higher RIS values:
487 and 516 nm/RIU, respectively [40,42]. Perfect Al concaves
arrangement is also possible by application of imprinting or pretexturing techniques prior to anodization [50]. On the other hand, the
electrochemical synthesis of Al nanoconcaves with smaller pitches
gives better pores arrangements and SPRs at lower wavelengths
(higher energies). Owing to the proportionality of the resonance
wavelength to the pitch size (Eq. (2)), plasmonic sensitivity is
consequently lower in the Al concaves with the smaller Dc , but
the material is very promising with respect to many important
applications in UV range, such as improvement of blue/UV light

M. Norek et al. / Applied Surface Science 314 (2014) 807814

extraction in LEDs, anti-reecting substrates for active layer in solar


cells, or biochemical sensing.
4. Conclusions
UV/violet plasmonic properties of Al nanoconcave arrays coated
with Ag, and Cu metals were investigated. Two types of the concaves were synthesized: one with the interpore distance (Dc ) of
238.2 nm and the second with Dc = 460.8 nm, as determined by
FFTs of SEM images. The properties were analyzed by monitoring reectivity spectra. Overall reectivity drops signicantly for
patterned samples and at the same time pronounced minima are
observed, the most distinctive one at photon wavelength dependent on the pitch size (254 nm for the sample with Dc = 238.2 nm
and 395 nm for the sample with Dc = 460.8 nm). The minima were
ascribed to (0,1) plasmonic mode ((0,1) ) based on Eq. (2). For the
sample anodized at 120 V the calculated (0,1) was smaller than the
experimentally observed reection minima, most probably owing
to a negative inuence of LSPs. In the sample characterized by
lower degree of pores ordering (anodized at 195 V) the interaction between LSPs was most likely reduced and the calculated (0,1)
was close to the measured one. After deposition of Ag, and Cu
the resonances is only slightly shifted, suggesting that plasmonic
properties of the composed material are mainly determined by
pure Al nanoconcave arrays and are only ne-tuned by other plasmonic metals. Refractive index sensitivity (RIS) increased after the
metals deposition as compared to pure Al nanohole arrays, which
was tentatively attributed to subsequent interface effects between
two metals layers. The highest RIS of 404 nm/RIU was obtained
for Cu coated Al nanoconcaves with Dc = 460.8 nm, which is similar or better than the RIS values previously determined for other
nanohole arrays, operating in visible/near IR range. The intensity
measured at reectivity dip was as low as 1.7% for Cu coated Al concaves array with Dc = 238.2 as compared to unstructured Al (100%),
which can be very interesting for prospective applications in solar
photovoltaics. The preparation of Al nanoconcave arrays is simple,
the geometrical parameters, such as pores and pitch size, can be
easily controlled by electrochemical parameters during synthesis
(voltage, anodization time, etc.). The material is reproducible and
suitable for high-throughput based detections. Furthermore, the
results obtained in this work suggest that plasmonic properties of
Al concaves arrays can be tailored by deposition of other plasmonic
metals on top of the concaves, which is very promising for rapidly
growing UV nanoplasmonics.
Acknowledgments
The research was nanced by National Science Centre (Decision
number: DEC-2012/07/D/ST8/02718). The work has been nancially supported by the Polish Ministry of Science and Higher
Education, Project: LAPROMAW (POIG.02.01.00-14-071/08/00).
Appendix A. Supplementary data
Supplementary data associated with this article can be
found, in the online version, at http://dx.doi.org/10.1016/j.apsusc.
2014.06.192.
References
[1] T.W. Ebbesen, H.J. Lezec, H.F. Ghaemi, T. Thio, P.A. Wolff, Extraordinary transmission through sub-wavelength hole arrays, Nature 391 (1998) 667669.
[2] J.V. Coe, J.M. Heer, S. Teeters-Kennedy, H.K.R. Tian, Rodriguez extraordinary
transmission of metal lms with arrays of subwavelength holes, Annu. Rev.
Phys. Chem. 59 (2008) 179202.
[3] C. Genet, T.W. Ebbesen, Light in tiny holes, Nature 445 (2007) 3946.

813

[4] L. Martn-Moreno, F.J. Garca-Vidal, H.J. Lezec, K.M. Pellerin, T. Thio, J.B. Pendry,
T.W. Ebbesen, Theory of extraordinary optical transmission through subwavelength hole arrays, Phys. Rev. Lett. 86 (2001) 11141117.
[5] S. Lal, S. Link, N.J. Halas, Nano-optics from sensing to waveguiding, Nat. Photon.
1 (2007) 641648.
[6] W.L. Barnes, A. Dereux, T.W. Ebbesen, Surface plasmon subwavelength optics,
Nature 424 (2003) 824830.
[7] J. Zhao, X.Y. Zhang, C.R. Yonzon, A.J. Haes, R.P. Van Duyne, Localized surface
plasmon resonance biosensors, Nanomedicine 1 (2006) 219228.
[8] M.D. Malinsky, K.L. Kelly, G.C. Schatz, R.P. Duyne, Chain length dependence
and sensing capabilities of the localized surface plasmon resonance of silver
nanoparticles chemically modied with alkanethiol self-assembled monolayers, J. Am. Chem. Soc. 123 (2001) 14711482.
[9] H. Chen, X. Kou, Z. Yang, W. Ni, J. Wang, Shape- and size-dependent refractive
index sensitivity of gold nanoparticles, Langmuir 24 (2008) 52335523.
[10] H. Wang, D.W. Brandl, F. Le, P. Nordlander, N.J. Halas, Nanorice A hybrid plasmonic nanostructure, Nano Lett. 6 (2006) 827832.
[11] A. Hajiaboli, M. Kahrizi, V.V. Truong, Optical behavior of thick gold and silver
lms with periodic circular nanohole arrays, J. Phys. D: Appl. Phys. 45 (2012)
485105 (8 pp).
[12] W.H. Chuang, J.Y. Wang, C.C. Yang, Y.W. Kiang, Differentiating the contributions
between localized surface plasmon and surface plasmon polariton on a onedimensional metal grating in coupling with a light emitter, Appl. Phys. Lett. 92
(2008) 133115 (3 pp).
[13] K.-L. Lee, W.-S. Wang, P.-K. Wei, Comparisons of surface plasmon sensitivities
in periodic gold nanostructures, Plasmonics 3 (2008) 119125.
[14] V. Giannini, Y. Francescato, H. Amrania, C.C. Phillips, S.A. Maier, Fano resonances
in nanoscale plasmonic systems: a parameter-free modeling approach, Nano
Lett. 11 (2011) 28352840.
[15] K.L. Molen, K.J. Klein Koerkamp, S. Enoch, F.B. Segerink, N.F. Hulst, L. Kuipers,
Role of shape and localized resonances in extraordinary transmission through
periodic arrays of subwavelength holes: experiment and theory, Phys. Rev. B
72 (2005) 045421 (9 pp).
[16] J.M. Sanz, D. Ortiz, R. Alcaraz de la Osa, J.M. Saiz, F. Gonzlez, A.S. Brown,
M. Losurdo, H.O. Everitt, F. Moreno, UV plasmonic behavior of various metal
nanoparticles in the near and far-eld regimes: geometry and substrate effects,
J. Phys. Chem. C 117 (2013) 1960619615.
[17] K. Aslan, C.D. Geddes, Directional surface plasmon coupled luminescence for
analytical sensing applications: which metal, what wavelength, what observation angle? Anal. Chem. 81 (2009) 69136922.
[18] P.R. West, S. Ishii, G.V. Naik, N.K. Emani, V.M. Shalaev, A. Boltasseva, Searching
for better plasmonic materials, Laser Photon. Rev. 4 (2010) 795808.
[19] T. Teranishi, M. Eguchi, M. Kanehara, S. Gwo, Controlled localize surface plasmon resonance wavelength for conductive nanoparticles over the ultraviolet
to near-infrared region, J. Mater. Chem. 21 (2011) 1023810242.
[20] L. Cui, D.Y. Wu, A. Wang, B. Ren, Z.Q. Tian, Charge-transfer enhancement
involved in the SERS of adenine on Rh and Pd demonstrated by ultraviolet to visible laser excitation, J. Phys. Chem. C 114 (2010) 16588
16595.
B. Kasemo, Plasmonic properties of supported
[21] C. Langhammer, Z. Yuan, I. Zoric,
Pt and Pd nanostructures, Nano Lett. 6 (2006) 833838.
[22] E.T. Papaioannou, V. Kapaklis, E. Melander, B. Hjrvarsson, D. Spiridon, S.D.
Pappas, P. Patoka, M. Giersig, P. Fumagalli, A. Garcia-Martin, G. Ctistis, Surface plasmons and magneto-optic activity in hexagonal Ni anti-dot arrays, Opt.
Express 19 (2011) 2386723877.
[23] K. Nakamoto, R. Kurita, O. Niwa, T. Fujii, M. Nishida, Development of a massproducible on-chip plasmonic nanohole array biosensor, Nanoscale 3 (2011)
50675075.
[24] K. Nakamoto, R. Kurita, O. Niwa, Electrochemical surface plasmon resonance
measurement based on gold nanohole array fabricated by nanoimprinting technique, Anal. Chem. 84 (2012) 31873191.
[25] A.G. Brolo, R. Gordon, B. Leathem, K.L. Kavanagh, Surface plasmon sensor based
on the enhanced light transmission through arrays of nanoholes in gold lms,
Langmuir 20 (2004) 48134815.
[26] P. Dong, Y. Lin, J. Deng, J. Di, Ultrathin goldshell coated silver nanoparticles
onto a glass platform for improvement of plasmonic sensors, Appl. Mater.
Interfaces 5 (2013) 23922399.
[27] P. Jia, H. Jiang, J. Sabarinathan, J. Yang, Plasmonic nanohole array sensors
fabricated by template transfer with improved optical performance, Nanotechnology 24 (2013), 195501 (6pp).
[28] G.H. Chan, J. Zhao, E.M. Hicks, G.C. Schatz, R.P. van Duyne, Plasmonic properties
of copper nanoparticles fabricated by nanosphere lithography, Nano Lett. 7
(2007) 19471952.
[29] R.K. Li, H. To, G. Andonian, J. Feng, A. Polyakov, C.M. Scoby, K. Thompson, W.
Wan, H.A. Padmore, P. Musumeci, Surface-plasmon resonance-enhanced multiphoton emission of high-brightness electron beams from a nanostructured
copper cathode, Phys. Rev. Lett. 110 (2013), 074801 (5pp).
[30] D.B. Pedersen, S. Wang, Surface plasmon resonance spectra of 2.8 0.5 nm
diameter copper nanoparticles in both near and far elds, J. Phys. Chem. C 111
(2007) 1749317499.
[31] I. Gryczynski, J. Malicka, K. Nowaczyk, Z. Gryczynski, J.R. Lakowicz, Effects of
sample thickness on the optical properties of surface plasmon-coupled emission, J. Phys. Chem. B 108 (2004) 1207312083.
[32] X. Zhang, Z. Li, S. Ye, S. Wu, J. Zhang, L. Cui, A. Li, T. Wang, S. Li, B. Yang, Elevated Ag nanohole arrays for high performance plasmonic sensors based on
extraordinary optical transmission, J. Mater. Chem. 22 (2012) 89038910.

814

M. Norek et al. / Applied Surface Science 314 (2014) 807814

[33] F. Xie, J.S. Pang, A. Centeno, M.P. Ryan, D.J. Riley, N.M. Alford, Nanoscale control
of Ag nanostructures for plasmonic uorescence enhancement of near-infrared
dyes, Nano Res. 6 (2013) 496510.
[34] R. Gordon, D. Sinton, K.L. Kavanagh, A.G. Brolo, A new generation of sensors based on extraordinary optical transmission, Acc. Chem. Res. 41 (2008)
10491057.
[35] M.R. Stewart, N.H. Mack, V. Malyarchuk, J.A.N.T. Soares, T.-W. Lee, S.K. Gray,
R.G. Nuzzo, J.A. Rogers, Quantitative multispectral biosensing and 1D imaging
using Quasi-3D plasmonic crystals, PNAS 46 (2006) 1714317148.
[36] Y. Wang, N. Lu, W. Wang, L. Liu, L. Feng, Z. Zeng, H. Li, W. Xu, Z. Wu, W. Hu, Y.
Lu, L. Chi, Highly effective and reproducible surface-enhanced Raman scattering
substrates based on Ag pyramidal arrays, Nano Res. 6 (2013) 159166.
[37] J. Martin, J. Proust, D. Grard, J. Plain, Localized surface plasmon resonances
in the ultraviolet from large scale nanostructured aluminum lms, Opt. Mater.
Express 3 (2013) 954959.
[38] G. Maidecchi, G. Gonella, R.P. Zaccaria, R. Moroni, L. Anghinol, A. Giglia, S.
Nannarone, L. Mattera, H.-L. Dai, M. Canepa, F. Bisio, Deep ultraviolet plasmon
resonance in aluminum nanoparticle arrays, ACS Nano 7 (2013) 58345841.
[39] J. Hu, L. Chen, Z. Lian, M. Cao, H. Li, W. Sun, N. Tong, H. Zeng, Deep-ultravioletblue-light surface plasmon resonance of Al and Alcore /Al2 O3shell in spherical and
cylindrical nanostructures, J. Phys. Chem. C 116 (2012) 1558415590.
[40] V. Canalejas-Tejero, S. Herranz, A. Bellingham, M.C. Moreno-Bondi, C.A. Barrios,
Passivated aluminum nanohole arrays for label-free biosensing applications,
Appl. Mater. Interfaces 6 (2014) 10051010.
Localized surface plasmon
[41] C. Langhammer, M. Schwind, B. Kasemo, I. Zoric,
resonances in aluminum nanodisks, Nano Lett. 8 (2008) 14611471.
[42] J.L. Skinner, L.L. Hunter, A.A. Talin, J. Provine, D.A. Horsley, Large-area subwavelength aperture arrays fabricated using nanoimprint lithography, IEEE Trans.
Nanotechnol. 7 (2008) 527531.
[43] M.W. Knight, L. Liu, Y. Wang, L. Brown, S. Mukherjee, N.S. King, H.O. Everitt,
P. Nordlander, N.J. Halas, Aluminum plasmonic nanoantennas, Nano Lett. 12
(2012) 60006004.
[44] X. Jiao, S. Blair, Optical antenna design for uorescence enhancement in the
ultraviolet, Opt. Express 20 (2012) 2990929922.
[45] M.B. Cortie, A.M. McDonagh, Synthesis and optical properties of hybrid and
alloy plasmonic nanoparticles, Chem. Rev. 111 (2011) 37174373.
[46] K.L. Molen, F.B. Segerink, N.F. Hulst, L. Kuipers, Inuence of hole size on the
extraordinary transmission through subwavelength hole arrays, Appl. Phys.
Lett. 85 (2004) 43164318.
[47] L. Wang, B. Xu, W. Bai, J. Zhang, L. Cai, H. Hu, G. Song, Multiple surface plasmon
resonances in compound structure with metallic nanoparticles and nanohole
arrays, Plasmonics 7 (2012) 659663.
[48] Z. Zhu, L. Shi, X. Liu, J. Zi, Z. Wang, A mechanically tunable plasmonic structure composed of a monolayer array of metal-capped colloidal spheres on an
elastomeric substrate, Nano Res. 3 (2010) 807812.
[49] L. Wu, X. Zhou, P. Bai, Plasmonic metals for nanohole-array surface plasmon led-enhanced uorescence spectroscopy biosensing, Plasmonics (2014),
http://dx.doi.org/10.1007/s11468-014-9667-6.
[50] W. Lee, R. Ji, U. Gsele, K. Nielsch, Fast fabrication of long-range ordered porous
alumina membranes by hard anodization, Nature 5 (2006) 741747.
[51] I. Horcas, R. Fernandez, J.M. Gomez-Rodriguez, J. Colchero, J. Gomez- Herrero,
A.M. Baro, WSXM: a software for scanning probe microscopy and a tool for
nanotechnology, Rev. Sci. Instrum. 78 (2007), 013705 (8pp).

[52] WSxM, http://www.nanotec.es

[53] W.J. Stepniowski,


Z. Bojar, Synthesis of anodic aluminum oxide (AAO) at
relatively high temperatures. Study of the inuence of anodization conditions on the alumina structural features, Surf. Coat. Technol. 206 (2011)
265272.

[54] W.J. Stepniowski,


A. Nowak-Stepniowska,
A. Presz, T. Czujko, R.A. Varin, The
effect of time and temperature on the arrangement of anodic aluminum oxide
nanopores, Mater. Charact. 91 (2014) 19.

[55] O. Pena-Rodrgueaz,
U. Pal, Effects of surface oxidation on the liner optical
properties of Cu nanoparticles, J. Opt. Soc. Am. B 11 (2011) 27352739.
[56] F. Li, L. Zhang, R.M. Metzger, On the growth of highly ordered pores in anodized
aluminum oxide, Chem. Mater. 10 (1998) 24702480.
[57] G. Sulka, Highly ordered porous alumina formation by self-organized anodizing, in: A. Eftekhari (Ed.), Nanostructures Materials in Electrochemistry,
WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2008, pp. 1116.

[58] M. Michalska-Domanska,
M. Norek, W.J. Stepniowski,
B. Budner, Fabrication of
high quality anodic aluminum oxide (AAO) on low purity aluminuma comparative study with the AAO produced on high purity aluminum, Electrochim.
Acta 105 (2013) 424432.
[59] A.D. Rakic, A.B. Djurisic, J.M. Elazar, M.L. Majewski, Optical properties of
metallic lms for vertical-cavity optoelectronic devices, Appl. Opt. 37 (1998)
52715283.
[60] A. Degiron, H.J. Lezec, W.L. Barnes, T.W. Ebbesen, Effects of hole depth on
enhanced light transmission through subwavelength hole arrays, Appl. Phys.
Lett. 81 (2002) 43274329.
[61] N. Bonod, S. Enoch, L. Li, E. Popov, M. Nevire, Resonant optical transmission
through thin metallic lms with and without holes, Opt. Express 11 (2003)
482490.
[62] J.-P. Lee, D. Chen, X. Li, S. Yoo, L.A. Bottomley, M.A. El-Sayed, S. Park, M. Liu,
Well-organized raspberry-like Ag@Cu bimetal nanoparticles for highly reliable and reproducible surface-enhanced Raman scattering, Nanoscale 5 (2013)
1162011624.
[63] W. Wang, J. Zhang, Y. Zhang, Z. Xie, G. Qin, Optical absorption enhancement
in submicrometre crystalline silicon lms with nanotexturing arrays for solar
photovoltaic applications, J. Phys. D: Appl. Phys. 46 (2013), 195106 (6pp).
[64] V.E. Ferry, L.A. Sweatlock, D. Pacici, H.A. Atwater, Plasmonic nanostructure design for efcient light coupling into solar cells, Nano Lett. 8 (2008)
43914397.
[65] Y. Wang, T. Sun, T. Paudel, Y. Zhang, Z. Ren, K. Kempa, Metamaterial-plasmonic
absorber structure for high efciency amorphous silicon solar cells, Nano Lett.
12 (2012) 440445.
[66] K.R. Catchpole, A. Polman, Plasmonic solar cells, Opt. Express 16 (2008)
2179321800.
[67] M. Najiminaini, F. Vase, B. Kaminska, J.J.L. Carson, Effect of surface plasmon
energy matching on the sensing capability of metallic nanohole arrays, Appl.
Phys. Lett. 100 (2012) 063110.
[68] F. Przybilla, A. Degiron, C. Genet, T.W. Ebbesen, F. de Lon-Prez, J. BravoAbad, F.J. Garca-Vidal, L. Martn-Moreno, Efciency and nite size effects in
enhanced transmission through subwavelength apertures, Opt. Express 16
(2008) 95719579.
[69] J. Henzie, M.H. Lee, T.W. Odom, Multiscale patterning of plasmonic metamaterials, Nat. Nanotechnol. 2 (2007) 549554.

You might also like