07.09.2021 Views

2015-3

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Tuning the effective plasma frequency

of nanorod metamaterials from visible to

telecom wavelengths

Cite as: Appl. Phys. Lett. 107, 121110 (2015); https://doi.org/10.1063/1.4931687

Submitted: 11 June 2015 . Accepted: 13 September 2015 . Published Online: 25 September 2015

M. E. Nasir, S. Peruch, N. Vasilantonakis, W. P. Wardley, W. Dickson, G. A. Wurtz, and A. V. Zayats

ARTICLES YOU MAY BE INTERESTED IN

Nanowire metamaterials with extreme optical anisotropy

Applied Physics Letters 89, 261102 (2006); https://doi.org/10.1063/1.2422893

From plane waves to local Gaussians for the simulation of correlated periodic systems

The Journal of Chemical Physics 145, 084111 (2016); https://doi.org/10.1063/1.4961301

Density matrices in full configuration interaction quantum Monte Carlo: Excited states,

transition dipole moments, and parallel distribution

The Journal of Chemical Physics 146, 244105 (2017); https://doi.org/10.1063/1.4986963

Appl. Phys. Lett. 107, 121110 (2015); https://doi.org/10.1063/1.4931687 107, 121110

© 2015 AIP Publishing LLC.


APPLIED PHYSICS LETTERS 107, 121110 (2015)

Tuning the effective plasma frequency of nanorod metamaterials

from visible to telecom wavelengths

M. E. Nasir, a) S. Peruch, N. Vasilantonakis, W. P. Wardley, W. Dickson, G. A. Wurtz,

and A. V. Zayats

Department of Physics, King’s College London, Strand, London WC2R 2LS, United Kingdom

(Received 11 June 2015; accepted 13 September 2015; published online 25 September 2015)

Hyperbolic plasmonic metamaterials are important for designing sensing, nonlinear, and emission

functionalities, which are, to a large extent, determined by the epsilon-near-zero behaviour

observed close to an effective plasma frequency of the metamaterial. Here, we describe a method

for tuning the effective plasma frequency of a gold nanorod-based metamaterial throughout the

visible and near-infrared spectral ranges. These metamaterials, fabricated by two-step anodization

in selenic acid and chemical post-processing, consist of nanorods with diameters of around 10 nm

and interrod distances of around 100 nm and have a low effective plasma frequency down to a

wavelength range below 1200 nm. Such metamaterials open up new possibilities for a variety of

applications in the fields of bio- and chemical sensing, nonlinearity enhancement, and fluorescence

control in the infrared. VC 2015 AIP Publishing LLC.[http://dx.doi.org/10.1063/1.4931687]

The ability to design the optical properties of metamaterials

has already achieved a significant impact in a variety

of photonic, data processing, and sensing applications. 1–6

Various types of metamaterials have been explored in the

radio-frequency (RF), terahertz, infrared, and optical regions

based on split-ring resonators, nanorod pairs, coaxial structures,

fishnets, and other approaches, allowing both the electric

and magnetic resonances to be engineered, providing

control of the permittivity and permeability of the nanostructured

media. 7–9 The adaptation of the above mentioned

approaches at longer wavelengths is straightforward, while

their scaling down to the visible spectral range is difficult

due to challenging top-down fabrication required at the subwavelength

scale.

A different class of metamaterials based on arrays of

aligned plasmonic nanorods has also been studied. 10 This

type of anisotropic metamaterial has unique electromagnetic

properties determined by the electromagnetic interaction

between the rods forming the array and can be described

within the effective medium theory (EMT) via the effective

permittivity tensor components e x ¼ e y 6¼ e z , corresponding

to the direction perpendicular to (x,y) and along (z) nanorod

axes. The metamaterial behaves as an indefinite metamaterial

(Re(e z ) ¼ 1) at RFs and can exhibit hyperbolic dispersion

in the spectral range where Re[e x,y (k)] > 0 and

Re[e z (k)] < 0. For a plasmonic nanorod-based metamaterial,

the hyperbolic regime has a short-wavelength cut-off typically

in the visible spectral range, but no long-wavelength

limit. The optical properties of such anisotropic metamaterials

are governed by their ability to support bulk plasmonpolaritons

which define the behaviour of the extraordinary

modes supported by the metamaterial. 11 Both ordinary and

extraordinary modes are related to the inter-rod coupling

via cylindrical surface plasmons (CSPs) supported by the

nanorods in the array. 12,13 These peculiar dispersion properties

exhibited by the metamaterial have been used for

a) mazhar.nasir@kcl.ac.uk

negative refraction index engineering and super-resolution

applications, deep-subwavelength wave guiding, spontaneous

emission control, nonlinearity enhancement, ultrasensitive

refractive index sensing, etc. 4–6,10,11,14–16

Many of the above described properties are defined by the

ÞŠ 0) of the metamaterial,

which is the onset of hyperbolic dispersion and around

which the so-called epsilon-near-zero (ENZ) regime takes

place. In contrast to other types of metamaterials, which can

be easily fabricated at long wavelengths, but which are problematic

in the visible, control over the bulk plasma frequency

of nanorod-based metamaterials can easily be achieved in the

visible spectral range by changing the interaction between the

nanorods in the array. This can be accomplished by changing

either the period of the array or the nanorod diameter, using a

bottom-up, template-based fabrication approach. Typically,

metamaterials comprised of arrays of aligned metallic nanorods

are synthesised in a porous alumina template fabricated

using conventional electrolytes, such as sulphuric or oxalic

acid, and have an effective plasma frequency limited to the

visible spectral range of 530–650 nm (Refs. 17 and 18) for Aubased

metamaterials and to shorter wavelengths for those fabricated

using Ag. The use of other plasmonic metals results in

increased losses due to the intrinsic material properties. Thus,

in contrast to other types of metamaterials, it is not straightforward

to fabricate such nanorod-based metamaterials with an

effective plasma frequency in the infrared and telecom spectral

ranges.

In this paper, we describe the design, fabrication and

characterisation of metamaterials based on periodic arrays of

vertically aligned Au nanorods which allows the effective

plasma frequency to be tuned throughout the visible and into

the infra-red spectral range. Highly ordered nanoporous alumina

templates with 15 nm pore diameter and 120 nm

interpore spacing over large cm-sized areas are fabricated by

two-step anodization in selenic acid. The effective plasma

frequency of Au-nanorod based metamaterials obtained with

such templates can then be tuned by controlling the nanorod

effective plasma frequency (Re½e z ðx ef f

p

0003-6951/2015/107(12)/121110/5/$30.00 107, 121110-1

VC 2015 AIP Publishing LLC


121110-2 Nasir et al. Appl. Phys. Lett. 107, 121110 (2015)

diameter via post processing of the pores while keeping both

the inter-rod spacing and rod length constant.

The optical properties of plasmonic nanorod metamaterials

are determined by the coupling of cylindrical surface

plasmons supported by individual nanorods in the array. On

the macroscopic level, this is well described by the Maxwell-

Garnett type EMTs which can be validated through comparison

with the experimental and microscopic numerical modelling.

19 The effective permittivities in both the in-plane and

out of plane directions are given by

ð1 þ pÞe Au þ ð1 pÞe h

e xy ¼ e h

ð1 pÞe Au þ ð1 þ pÞe h

e z ¼ pe Au þ ð1

pÞe h ; (1)

where p ¼ p(r/d) 2 is the nanorod concentration with r being

radius of the nanorods and d being distance between the

nanorods, e Au and e h are the permittivities of Au and the host

medium (Al 2 O 3 ), respectively. This model is valid away

from the Brillouin zone edge of the nanorod array with

k 0 p d , where k 0 is the wavevector of the incident light in

the plane of the metamaterial slab. 11 The effective plasma

frequency of Au nanorod metamaterials, which is determined

by Re(e z ) ¼ 0, is affected by both the host medium properties

and the nanorod concentration. 11 This frequency separates

the elliptic dispersion regime, where the metamaterial

behaves as a strongly anisotropic dielectric, and the hyperbolic

regime where the metamaterial supports bulk plasmonpolaritons

due to its metallo-dielectric behaviour.

In order to fabricate the metamaterial, Au nanorods are

electrodeposited in porous anodic alumina oxide (AAO) templates,

synthesized by a two-step anodization process. An aluminium

film of 800 nm thickness was sputtered on a

multilayer substrate comprised of a glass slide with a 10 nm

thick adhesion layer of tantalum pentoxide and a 7 nm thick

Au film acting as a weakly conducting layer. Tantalum pentoxide

is deposited by sputtering tantalum using a 20% oxygen/80%

argon mixture. The porous alumina structures were

synthesized by a two step-anodization in 0.3 M selenic acid

at 40–48 V at 0 C. The temperature of the sample and electrolyte

was controlled throughout the anodization process.

After an initial anodization step, the porous layer formed

was removed by etching in a solution of H 3 PO 4 (3.5%) and

CrO 3 (20 g l –1 )at70 C leaving an ordered, patterned surface.

The samples were then subjected to a second anodization

step under the same conditions as in the first step and

subsequently etched in 30 mM NaOH to tune the diameters

of pores from 15 nm to 50 nm. Gold electrodeposition was

performed using a three-electrode geometry and a noncyanide

solution. The length of the nanorods was controlled

by the electrodeposition time.

In general, the geometrical parameters of the porous alumina

template used for nanorod metamaterial fabrication are

determined by the electrolyte used for anodization and by

the applied anodizing potential. The pore diameter and interpore

spacing of the porous alumina template is proportional

to the applied voltage with proportionality constants

k d ¼ 1.29 nm V 1 and k int ¼ 2.52 nm V 1 , respectively. 20–24

Therefore, for a chosen electrolyte, the diameter of the pores

can be increased in a pore widening process for a given

voltage, resulting in an increase in the nanorod concentration

p. However, due to the limitations of currently used electrolytes,

a different approach is required in order to provide

ENZ-related functionalities in the infrared, including at telecom

wavelengths.

Self-ordered porous alumina is typically obtained in

acidic electrolytes, such as sulphuric, oxalic, phosphoric,

malonic, and tartaric acids. 25–28 In these electrolytes, the diameter

of pores is determined by the applied voltage with a

proportionality constant equal to or greater than 1 nm/V and

the interpore distance determined by a proportionality constant

2.52 nm V 1 . In selenic acid, 24 the pore diameter of

the self-ordered porous alumina template has a weaker dependence

on the applied voltage (a proportionality constant

is 0.3 nm/V) while the interpore distance has about the

same dependence (proportionality constant 2.33 nm V 1 ).

Figure 1(b) shows the SEM image of the self-ordered porous

alumina template formed in 0.3 M selenic acid at 48 V,

keeping the temperature at 0 C. The regular pore arrays

have been observed within micrometre sized domains. The

diameter of the pores is 15 nm with an interpore separation

of approximately 112 nm. The porosity, determining nanorod

concentration p, of the anodic alumina obtained in selenic

acid is 3.4%, much lower than achievable with sulphuric

acid (down to 12.6%). 22,29 It is very interesting to note that

both sulphuric acid and selenic acid possess similar chemical

structures (Fig. 1(c)) but totally different anodization behaviour.

In the case of acids having similar chemical structure,

acid strength decreases as the size of the central atom

increases. The atomic size of selenium is bigger than sulphur;

therefore, selenic acid is less soluble than sulphuric acid

under the same anodization conditions. This low solubility

accounts for the much smaller pore diameter observed here.

With a decrease in anodization temperature, the solubility

of selenic acid decreases leading to the unusual selfordered

porous alumina with smaller than expected pore

diameter, while the period remains unaffected. This unique

porous alumina structure has enabled the metamaterials’

plasma frequency to be tuned throughout the visible and

near-infrared spectral ranges. Additionally, since the geometry

of the gold nanorods is determined by the porous alumina

template, the diameter of nanorods may be varied by changing

the pore diameter in a pore widening process while keeping

the inter-rod spacing the same. Alternatively, we can

also tune the pore diameter while keeping the interpore distance

constant by raising the temperature during the anodization

process. Figure 1(e) shows the effect of varying the

nanorod diameter while maintaining constant nanorod spacing

and length. With a reduction in nanorod concentration, a

monotonous long-wavelength shift of the dominating extinction

peak is observed, accompanied by a broadening of the

peak due to the red-shift of the effective plasma frequency

(as the concentration of metal becomes smaller the interaction

between the nanorods becomes weaker).

Metamaterials comprised of Au nanorods embedded in

an AAO matrix were designed with the effective plasma frequency

in the infrared spectral range (Fig. 2). The extinction

spectra show the two typical dominating resonances, which

are well separated from each other spectrally. A short wavelength

resonance is associated with plasmonic excitations by


121110-3 Nasir et al. Appl. Phys. Lett. 107, 121110 (2015)

FIG. 1. (a) Schematics of the metamaterial based on a nanorod array. (b) SEM image of porous alumina synthesized in selenic acid at 48 V (sample B). (c)

Comparison of sulphuric and selenic acid molecules. (d) SEM image of the nanorods after removal of alumina matrix (sample B). Please note that disorder is

introduced in the nanorod array after the matrix removal due to the reduced self-supporting properties of thin nanorods. (e) The effect of nanorod concentration

on the extinction of the metamaterial (the Au nanorods in AAO matrix, p-polarized light, 40 angle of incidence).

light polarised perpendicular to the nanorod axes, while the

strong long-wavelength absorption near the effective plasma

frequency takes place for incident light polarised along the

nanorod axes. 30 These two resonances, of a different nature,

remain separated even when the AAO is removed, when for

the samples with higher nanorod concentrations they typically

overlap. For smaller nanorod diameters with the same

period, the peak in the ENZ region is shifted further in the

FIG. 2. Extinction spectra of metamaterials (a–c) A (nanorods of 40 nm diameter, 115 nm period, and 200 nm height, embedded in AAO matrix) and (d–f)

B (nanorods of 25 nm diameter, 115 nm period, and 350 nm height, embedded in AAO matrix) at different angles of light incidence: (a) and (d) experiment,

(b) and (e) full-vectorial microscopic modelling, (c) and (f) EMT modelling. For sample A an electron mean free path restriction of 23 nm was used for electrochemical

Au; 18 the metamaterial has a 7 nm thick Au underlayer and a 700 nm thick AAO overlayer. For sample B, the electron mean free path restriction is

8 nm for electrochemical Au; the metamaterial has a 7 nm thick Au underlayer and a 650 nm thick AAO overlayer.


121110-4 Nasir et al. Appl. Phys. Lett. 107, 121110 (2015)

FIG. 3. Effective medium parameters

of (a) metamaterial A and (b) metamaterial

B. Please note that Im(e z ) curves

in AAO and air overlap in (a) and (b).

(c) Experimentally measured and (d)

simulated reflectance dispersions of

the metamaterial B for p-polarised

light. The apparent change of contrast

in (c) is due to the change of the detector

in the measurements from the visible

(top panel) to IR (bottom panel).

The angular range corresponds to

30 –70 in glass. The light line in air

and the effective plasma frequency are

marked with solid and dashed lines,

respectively.

infrared spectral range to 1200 nm. This is related to two

simultaneous effects: an increase in the nanorod aspect ratio

which affects the CSP of the individual nanorods, and a

reduction of the coupling between the CSPs on neighbouring

nanorods.

The metamaterials’ optical properties have been modelled

using both an analytical transfer matrix method (TMM)

combined with an EMT based homogenization (Eq. (1)) as

well as full-vectorial microscopic numerical simulations

based on a finite element method (FEM). The tabulated permittivity

of Au was used 31 incorporating a correction to the

mean free path of electrons in electrochemically deposited

gold. 18 In the case of the numerical modelling, the unit cell

of a periodic square array with the periodicity given by the

inter-rod distance was modelled. The experimental data is in

good agreement with both numerical and EMT modelling,

reproducing the spectral position of the extinction peaks.

The EMT theory in particular, allows efficient extraction of

the effective medium parameters of the metamaterial and the

effective plasma frequency (Figs. 3(a) and 3(b)).

The real part of the effective permittivity components,

e x,y and e z , have the same sign for short wavelengths, where

the metamaterials operate in the elliptical dispersion regime,

while the effective plasma frequency is reached at wavelengths

around 795 nm and 1280 nm for metamaterials A and

B, respectively. For longer wavelengths, the metamaterials

are hyperbolic, supporting bulk plasmon polaritons. 11 If the

AAO matrix is removed so that the nanorods are in air, the

effective plasma frequency is shifted to shorter wavelengths

of about 625 nm and 850 nm, for metamaterials A and B,

respectively. Still the effective plasma frequency is sufficiently

away from the absorption resonance of e x,y .

The reflection spectra (Figs. 3(c) and 3(d)) show a set of

waveguided modes supported by the anisotropic metamaterial

slab in the visible spectral range where the dispersion is

elliptic in nature. The measured modes lie between light

lines of the air superstrate and the glass substrate, thus they

are leaky in the substrate and can be excited in attenuated

total internal reflection (ATR) measurements. 11 For the angle

of incidence corresponding to wavevectors smaller than the

light line in air, the modes are Fabry-Perot-type resonances

of the planar metamaterial slab. The mode structure drastically

changes in the IR spectral range below the effective

plasma frequency, where the dispersion is hyperbolic, and,

for the p-polarised light, the metamaterial has a strong metallic

behaviour supporting plasmon-polaritonic modes. 11 The

measured mode dispersions (Fig. 3(c)) are in a good agreement

with simulated dispersions (Fig. 3(d)) of the metamaterial

slab. Thus, the planar metamaterial waveguides can be

designed to support different types of modes in different dispersion

ranges using the control over the effective plasma

frequency developed here.

In conclusion, we have demonstrated hyperbolic

nanorod based metamaterials which allow improved flexibility

in designing the effective plasma frequency

throughout the visible and near-infrared ranges using

anodization in selenic acid. In particular, this approach

allows the effective plasma frequency to be designed in

the technologically important telecom wavelength range.

Such metamaterials are important for extending the spectral

reach of these versatile metamaterials and are

expected to impact applications in polarisation optics,

design of nonlinear optical response, sensing applications

and fluorescence control in the infrared.

This work was supported, in part, by EPSRC (UK)

and the ERC iPLASMM Project (No. 321268). A.Z.

acknowledges support from the Royal Society and the

Wolfson Foundation. G.W. acknowledges the support from

the EC FP7 Project No. 304179 (Marie Curie Actions).

The data access statement: all data supporting this

research are provided in full in the results section.


121110-5 Nasir et al. Appl. Phys. Lett. 107, 121110 (2015)

1 R. Kirchain and L. Kimerling, Nat. Photonics 1, 303 (2007).

2 D. Dregely, F. Neubrech, H. Duan, R. Vogelgesang, and H. Giessen, Nat.

Commun. 4, 2237 (2013).

3 A. S. Urban, S. Carretero-Palacios, A. A. Lutich, T. Lohm€uller, J.

Feldmann, and F. J€ackel, Nanoscale 6, 4458 (2014).

4 V. V. Yakovlev, W. Dickson, A. Murphy, J. McPhillips, R. J. Pollard, V.

A. Podolskiy, and A. V. Zayats, Adv Mater. 25, 2351 (2013).

5 A. V. Kabashin, P. Evans, S. Pastkovsky, W. Hendren, G. A. Wurtz, R.

Atkinson, R. Pollard, V. A. Podolskiy, and A. V. Zayats, Nat. Mater. 8,

867 (2009).

6 M. E. Nasir, W. Dickson, G. A. Wurtz, W. P. Wardley, and A. V. Zayats,

Adv. Mater. 26, 3532 (2014).

7 C. M. Soukoulis and M. Wegener, Nat. Photonics 5, 523 (2011).

8 S. P. Burgos, R. de Waele, A. Polman, and H. A. Atwater, Nat. Mater. 9,

407 (2010).

9 C. Garcıa-Meca, J. Hurtado, J. Martı, A. Martınez, W. Dickson, and A. V.

Zayats, Phys. Rev. Lett. 106, 067402 (2011).

10 C. R. Simovski, P. A. Belov, A. V. Atrashchenko, and Y. S. Kivshar, Adv.

Mater. 24, 4229 (2012).

11 N. Vasilantonakis, M. E. Nasir, W. Dickson, G. A. Wurtz, and A. V.

Zayats, Laser Photonics Rev. 9, 345 (2015).

12 G. A. Wurtz, W. Dickson, D. O’Connor, R. Atkinson, W. Hendren, P.

Evans, R. Pollard, and A. V. Zayats, Opt. Express 16, 7460 (2008).

13 K.-T. Tsai, G. A. Wurtz, J.-Y. Chu, T.-Y. Cheng, H.-H. Wang, A. V.

Krasavin, J.-H. He, B. M. Wells, V. A. Podolskiy, J.-K. Wang, and A. V.

Zayats, Nano Lett. 14, 4971 (2014).

14 G. A. Wurtz, R. Pollard, W. Hendren, G. Wiederrecht, D. Gosztola, V. A.

Podolskiy, and A. V. Zayats, Nat. Nanotechnol. 6, 107 (2011).

15 A. D. Neira, N. Olivier, M. E. Nasir, W. Dickson, G. A. Wurtz, and A. V.

Zayats, Nat. Commun. 6, 7757 (2015).

16 V. A. Podolskiy, P. Ginzburg, B. Wells, and A. V. Zayats, Faraday

Discuss. 178, 61 (2015).

17 P. Evans, W. R. Hendren, R. Atkinson, G. A. Wurtz, W. Dickson, A. V.

Zayats, and R. J. Pollard, Nanotechnology 17, 5746 (2006).

18 R. Pollard, A. Murphy, W. Hendren, P. Evans, R. Atkinson, G. A.

Wurtz, A. V. Zayats, and V. A. Podolskiy, Phys. Rev. Lett. 102,

127405 (2009).

19 J. Elser, R. Wangberg, V. A. Podolskiy, and E. E. Narimanov, Appl. Phys.

Lett. 89, 261102 (2006).

20 J. O’Sullivan and G. Wood, Proc. R. Soc. London, Ser. A 317, 511

(1970).

21 K. Nielsch, J. Choi, K. Schwirn, R. B. Wehrspohn, and U. G€osele, Nano

Lett. 2, 677 (2002).

22 J. Martın, C. V. Manzano, O. Caballero-Calero, and M. Martın-Gonzalez,

ACS Appl. Mater. Interfaces 5, 72 (2013).

23 W. Lee, R. Ji, U. G€osele, and K. Nielsch, Nat. Mater. 5, 741 (2006).

24 O. Nishinaga, T. Kikuchi, S. Natsui, and R. O. Suzuki, Sci. Rep. 3, 2748

(2013).

25 N. Myung, J. Lim, J. Fleurial, M. Yun, W. West, and D. Choi,

Nanotechnology 15, 833 (2004).

26 A. Cai, H. Zhang, H. Hua, and Z. Zhang, Nanotechnology 13(5), 627

(2002).

27 W. Lee, K. Nielsch, and U. G€osele, Nanotechnology 18, 475713 (2007).

28 G. D. Sulka, in Nanostructured Materials in Electrochemistry (Wiley-

VCH Verlag GmbH & Co. KGaA, 2008), pp. 1–116.

29 H. Masuda, F. Hasegwa, and S. Ono, J. Electrochem. Soc. 144, L127

(1997).

30 R. Atkinson, W. Hendren, G. A. Wurtz, W. Dickson, A. V. Zayats, P.

Evans, and R. Pollard, Phys. Rev. B 73, 235402 (2006).

31 P. B. Johnson and R. W. Christy, Phys. Rev. B 6, 4370 (1972).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!