Vous êtes sur la page 1sur 11

A circuit model for plasmonic resonators

Di Zhu,1 Michel Bosman,1 and Joel K. W. Yang1,2,*


1

Institute of Materials Research and Engineering, A*STAR, 3 Research Link, 117602, Singapore
2
Singapore University of Technology and Design, 20 Dover Drive, 138682, Singapore
*joel_yang@sutd.edu.sg

Abstract: Simple circuit models provide valuable insight into the properties
of plasmonic resonators. Yet, it is unclear how the circuit elements can be
extracted and connected in the model in an intuitive and accurate manner.
Here, we present a detailed treatment for constructing such circuits based on
energy and charge oscillation considerations. The accuracy and validity of
this approach was demonstrated for a gold nanorod, and extended for a
split-ring resonator with varying gap sizes, yielding good intuitive and
quantitative agreement with full electromagnetic simulations.
2014 Optical Society of America
OCIS codes: (240.6680) Surface plasmons; (260.5740) Resonance; (350.4990) Particles;
(310.6628) Subwavelength structures, nanostructures.

References and links


1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.

N. Engheta, A. Salandrino, and A. Al, Circuit Elements at Optical Frequencies: Nanoinductors, Nanocapacitors, and Nanoresistors, Phys. Rev. Lett. 95(9), 095504 (2005).
N. Engheta, Circuits with light at nanoscales: optical nanocircuits inspired by metamaterials, Science
317(5845), 16981702 (2007).
A. Al and N. Engheta, Input Impedance, Nanocircuit Loading, and Radiation Tuning of Optical Nanoantennas, Phys. Rev. Lett. 101(4), 043901 (2008).
A. Al and N. Engheta, Tuning the scattering response of optical nanoantennas with nanocircuit loads, Nat.
Photonics 2(5), 307310 (2008).
A. Al, M. Young, and N. Engheta, Design of nanofilters for optical nanocircuits, Phys. Rev. B 77(14),
144107 (2008).
M. G. Silveirinha, A. Al, J. Li, and N. Engheta, Nanoinsulators and nanoconnectors for optical nanocircuits,
J. Appl. Phys. 103(6), 064305 (2008).
A. Al and N. Engheta, All Optical Metamaterial Circuit Board at the Nanoscale, Phys. Rev. Lett. 103(14),
143902 (2009).
Y. Sun, B. Edwards, A. Al, and N. Engheta, Experimental realization of optical lumped nanocircuits at infrared wavelengths, Nat. Mater. 11(3), 208212 (2012).
M. Staffaroni, J. Conway, S. Vedantam, J. Tang, and E. Yablonovitch, Circuit analysis in metal-optics, Photon.
Nanostructures 10(1), 166176 (2012).
H. Caglayan, S.-H. Hong, B. Edwards, C. R. Kagan, and N. Engheta, Near-Infrared Metatronic Nanocircuits by
Design, Phys. Rev. Lett. 111(7), 073904 (2013).
H. Duan, A. I. Fernndez-Domnguez, M. Bosman, S. A. Maier, and J. K. Yang, Nanoplasmonics: classical
down to the nanometer scale, Nano Lett. 12(3), 16831689 (2012).
M. Bosman, E. Ye, S. F. Tan, C. A. Nijhuis, J. K. W. Yang, R. Marty, A. Mlayah, A. Arbouet, C. Girard, and M.
Y. Han, Surface Plasmon Damping Quantified with an Electron Nanoprobe, Sci. Rep. 3, 1312 (2013).
J. D. Jackson, Classical Electrodynamics, 3rd ed. (Wiley, 1999).
C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (Wiley, 1983).
J. W. Nilsson and S. A. Riedel, Electric Circuits, 9th ed. (Prentice Hall, 2011).
F. Wang and Y. R. Shen, General Properties of Local Plasmons in Metal Nanostructures, Phys. Rev. Lett.
97(20), 206806 (2006).
P. B. Johnson and R. W. Christy, Optical Constants of Noble Metals, Phys. Rev. B 6(12), 43704379 (1972).
S. A. Maier, Plasmonics: Fundamentals and Applications, 1st ed. (Springer, 2007).
C. P. Huang, X. G. Yin, H. Huang, and Y. Y. Zhu, Study of plasmon resonance in a gold nanorod with an LC
circuit model, Opt. Express 17(8), 64076413 (2009).
J. B. Pendry, A. J. Holden, D. J. Robbins, and W. J. Stewart, Magnetism from conductors and enhanced nonlinear phenomena, IEEE Trans. Microw. Theory 47(11), 20752084 (1999).
R. A. Shelby, D. R. Smith, and S. Schultz, Experimental verification of a negative index of refraction, Science
292(5514), 7779 (2001).
A. W. Clark, A. Glidle, D. R. S. Cumming, and J. M. Cooper, Plasmonic Split-Ring Resonators as Dichroic
Nanophotonic DNA Biosensors, J. Am. Chem. Soc. 131(48), 1761517619 (2009).

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9809

23. E. Cubukcu, S. Zhang, Y. S. Park, G. Bartal, and X. Zhang, Split ring resonator sensors for infrared detection of
single molecular monolayers, Appl. Phys. Lett. 95(4), 043113 (2009).
24. A. Pors, M. Willatzen, O. Albrektsen, and S. I. Bozhevolnyi, From plasmonic nanoantennas to split-ring resonators: tuning scattering strength, J. Opt. Soc. Am. B 27(8), 16801687 (2010).
25. O. Sydoruk, E. Tatartschuk, E. Shamonina, and L. Solymar, Analytical formulation for the resonant frequency
of split rings, J. Appl. Phys. 105(1), 014903 (2009).
26. M. Amin, M. Farhat, and H. Bac, A dynamically reconfigurable Fano metamaterial through graphene tuning
for switching and sensing applications, Sci. Rep. 3, 2105 (2013).
27. B. Willingham and S. Link, A Kirchhoff solution to plasmon hybridization, Appl. Phys. B 113(4), 519525
(2013).

1. Introduction
Localized plasmon resonances are the collective oscillations of electrons in metal nanostructures. They can conveniently be modeled by lumped-circuit elements in the form of a resistor,
inductor, and capacitor (RLC). Doing so led to the concept of optical nanocircuits that was
first introduced by Engheta et al., and has since generated great interest [110]. In its first
report, a circuit model was constructed for the example of a nanosphere [1]. Although mathematically accurate, the determination of the complex impedance of the circuit from the
total displacement current and the average voltage across the nanostructure does not necessarily produce a physical model. For instance, one would expect that increasing the material resistivity would directly cause an increase in the extracted R and a reduction in the quality factor (Q-factor) of the resonance; or that changing the geometry of the structure to increase the
kinetic inductance would increase the extracted L. However, this basic physical intuition is
missing in existing treatments of the problem. A detailed treatment is thus needed to explicitly
and intuitively link the RLC values of the circuit model to the nanostructure properties, and
the physical processes during the resonance.
In this work, we detail the procedure for circuitizing plasmonic nanostructures from a
thermodynamic perspective. Here, each circuit element is associated with an energy component obtained by considering that a resonator requires the cyclic exchange of different forms
of energy. For plasmonic resonators, this cyclic exchange occurs between (1) the electric potential energy (arising from the accumulation of charges on surfaces of the nanostructure), and
(2) the sum of kinetic energy of the moving electrons, and the induced magnetic field energy
from the current flow. Dissipation or damping of the oscillation occurs through electron scattering (Joule heating) and radiative damping. Intuitively, the electric potential energy is captured by a capacitor (C); the electron kinetic energy by a kinetic inductor (LK) [9]; the magnetic field energy by a Faraday inductor (LF); and the losses by an ohmic resistor (Rohmic) and a
radiative-loss resistor (Rrad). In contrast to previous work [18, 10], where displacement current is the only current flow, we explicitly introduce a term for the conduction current due to
the moving electrons. Doing so links the currents in the RLC circuit to the charge flow in the
nanostructure. We compare our model to the previous work, and test the correctness of the
model by observing its ability to reproduce the spectral response for a nanorod plasmonic
resonator (including resonance frequency and Q-factor), and its extension to a split-ring resonator.
2. Circuitizing a plasmonic resonator
First, we consider the charge oscillation in the metal arising from the free-electron gas that is
enclosed by the nanostructure surface. Therefore, the total displacement current, iD, can be
separated into two parts as follows:
iD = i E = i ( 0 )E i 0 E = J iD0

(1)

where J is the conduction current, corresponding to charge flow, and iD0 is the free-space
displacement current, caused by charge accumulation (see Fig. 1(a)-1(c)); is the complex
permittivity of the material, and 0 is the free-space permittivity; E is the electric field.

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9810

This separation leads to a unit-cell circuit with three branches, containing iD, J and
iD0 (see Fig. 1(d)). By separating the conduction and the free-space displacement current,
we rewrite the standard source-free Maxwells equations into the following form:
D0 =

0 H = 0
= ( 0 )E

with

J = i ( 0 )E = E
H = iD0 + J
E = i0 H

(2)

where H is the magnetic field; 0 is the free-space permeability; is the charge density within
the metal; and is the optical conductivity, equal to i(0).

Fig. 1. The process of forming a lumped circuit model for a metal nanorod. (a) The total displacement current density, i(D), can be separated into two parts: (b) the conduction current
density, (J), due to free electrons, and (c) free-space displacement current density, i(D)0,
that stems from the surface charges accumulated on the nanorod [Fields in (c) were scaled to
accentuate the free-space displacement current]. The electron kinetic energy and Joule heating
caused by the conduction current in (b) are modeled using an inductor and a resistor in series;
while the electric potential energy stored in the displacement current (c) is modeled using a capacitor. (d) Schematic of a unit-cell circuit. (e) Overall RLC lumped circuit formed by defining the lumped current as the net current in each branch in (d) [R, L, and C are per-unit values, while R, L and C are lumped values.]

The insight gained by modeling R and L in series is immediately seen from the Drude
model [9]. Here, the permittivity of metal is given by m = 0 (1 P2 / ( 2 + i )) , with

P2 = ne 2 / (m 0 ) (n is the free-electron density, m is electron mass, e is electron charge, and


is the inelastic electron scattering rate (damping parameter)). From Eq. (2), we express the
optical resistivity as 1/ = m/(ne2)im/(ne2) = RiL. R and L are identified as the intrinsic resistivity and inductivity of the material. Therefore, a circuit consisting of R in series with
L makes good intuitive sense. We then substitute the expression for current density, |J| = nev,
where v is the drift velocity of the electron, and integrate over the volume of the metal nanostructure, Vmetal, to determine the time-averaged energy stored in L and power dissipated in R.
We obtain WL = |J|2Ldv = mv2ndv, which is the expression for the total kinetic energy of
electrons; and PR = |J|2Rdv = mv2ndv, which represents the power dissipated due to the
inelastic scattering of electrons.

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9811

To transform the unit-cell circuit to an overall macroscopic lumped circuit, it is necessary to define the lumped currents, I0, I1, and I2. These are the total currents in the three
branches shown in Fig. 1(d) obtained via surface integrals of iD, J, and iD0:
I 0 = iD n d A = ( iD) dv = 0

(3.1)

I1 = J n d A = J dv = i dv = iQ

(3.2)

I 2 = ( iD0 ) n d A = ( iD0 ) dv = i dv = iQ

(3.3)

US

US

US

where US denotes the upper-half surface of the nanorod; Q is the total charge accumulated on
the upper half surface. Since I0 = 0 and I1 = I2 = iQ = I, the unit-cell circuit in Fig. 1(d)
is transformed into an overall lumped circuit shown in Fig. 1(e) in accordance to Kirchhoffs
current law. This source-free circuit models the impulse-response of the RLC circuit that undergoes transient oscillation and damping, reminiscent to the response of resonators as excited
experimentally by a beam of energetic electrons, using electron energy-loss spectroscopy
(EELS) [11, 12], and in pulse-response calculations as done in finite-difference time-domain
(FDTD) simulations, e.g. using Lumerical FDTD Solutions.
Next, we consider the power flow in the plasmonic resonator to identify the role and value
of each circuit element. Applying Poyntings Theorem for harmonic fields in phasor notation,
in which all fields have a time dependence exp( it ) , we obtain [13]:
1
1
1 2
J E dv =
J dv

metal
V
V
2
2

1
= ( H + iD0 ) E dv
2 V

1
1
2
2
= S dv +
iD0 dv + i 0 H dv

V
V
V
2
i 0

(4)

where Vmetal is the volume of the nanostructure, and V is the volume of the nanostructure and
its surroundings (the entire simulation space in practice); S = E H* is the complex Poynting
vector. Practically, the integral is performed for a simulation space that is sufficiently large,
where the boundary of V is at least one half wavelength away from the nanostructure.
Separating out the real and imaginary components in Eq. (4) yields the following equations:
1 1
1
1
1
2
2
2
iD0 dv + i Im( )
J dv i0 H dv = 0

V
V
V
metal
2 i 0
2
2


P
P
PE

1
1
1
1
2
( S)in dv = Re( )
J dv + ( S) out dv

V
Vmetal

2
2
2 V


Pin

Pohmic

(5.1)

(5.2)

Prad

In an RLC circuit, the power delivered at each circuit element can be expressed as Z|I|2,
where Z is the impedance and I is the current. As the capacitor and inductor have opposite
signs for their impedances (i/C and iL), their reactive powers are 180 Deg out of phase,
denoting the cyclic power flow between them. On the other hand, a resistor has a real impedance (R) and power, corresponding to the power dissipation.
Equation (5.1) consists of three reactive power terms that sum to zero, describing the
energy oscillation between the inductive and capacitive circuit elements at resonance. (1) PE
is positive (capacitive), and the energy stored in the corresponding capacitor (C) is the total
#207066 - $15.00 USD
(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9812

electrostatic energy (|PE/| = 0|E|2dv). (2) Because of the negative imaginary optical resistivity for metals, PK is negative (inductive), and the energy stored in the corresponding inductor (LK, kinetic inductance) is the electron kinetic energy (|PK/| = Im[1/]|J|2dv/(i)). (3)
The last term, PM, is also inductive, and emanates from a Faraday inductance (LF), because it
arises from the total magnetic energy (|PM/| = 0|H|2dv).
On the other hand, Eq. (5.2) consists of the balance of three real power terms in accordance to the conservation of energy. Note that
parts:

( S)in dv and

S dv in Eq. (4) was separated into two

( S)out dv , denoting the power flow in and out of V respec-

tively. Thus, the left hand side, Pin, is the power input; and on the right hand side, Pohmic and
Prad are the ohmic and radiative losses, respectively.
Finally, linking all the power (Eq. (5)) and current (Eq. (3)) terms, the corresponding circuit parameters can be expressed as follows based on the simple formula in circuit theory, i.e.
P = |I|2Z:
2
2
2
1
Rohmic = 2 Pohmic / I = Re
J dv / I
V
metal

Rrad = 2 Pout / I = ( S) out dv / I


2

(6.2)

C = i I / (2 PE ) =| Q |2 /( 0 E dv)
2

LK = 2 PK / ( i I ) =

1
2
2
1
J dv / I
Im

Vmetal

LF = 2 PM / ( i I ) = 0 H dv/ I
2

(6.1)

(6.3)
(6.4)
(6.5)

These expressions provide a means for extracting the RLC components from the corresponding physical quantities.
3. Circuit model for nanospheres

In this section, we analytically derive the circuit parameters of a metal nanosphere using our
model, for the ease of direct comparison with the previous work [1].
Quasi-static approximation for the electric field inside and outside a nanosphere under an
external excitation electric field E0 writes [1, 14]:
Eint = 3 0 E0 / ( + 2 0 )
Eext = E0 + Edip = E0 + [3u( p u) p] / (4 0 r 3 )

(7)

where p = 40a3(0)E0/( + 20); u = r/r; 0 is the free space permittivity; is the permittivity of the nanosphere; a is the radius of the sphere; and r is the position vector, with r = |r|.
Removing the excitation field E0 gives the source-free fields inside and outside the nanosphere:
Ei = Eint E0 = ( 0 )E0 / ( + 2 0 )
Eout = Edip

(8)

From Eq. (3), we obtain the current and charge values:


I = J n d A = Ei n d A= Ei a 2 ,
US

US

Q = I / ( i )

#207066 - $15.00 USD


(C) 2014 OSA

(9)

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9813

where = i(0) is the optical conductivity.


From Eq. (6.1), we obtain the resistance value:

Rohmic

4 3
2
a Ei
4
2
2
1
1 3
1
J dv / I = Re
Re
= Re
=
2
metal
V
2
3 a

Ei a

(10)

From Eq. (6.4), we obtain the inductance value:


4 3
2
a Ei
1
1
4
2
2
1
1 3
1
LK =
J dv / I =
Im
Im
Im (11)
=
2
metal
V
2

3 a

Ei a

From Eq. (6.3), we obtain the capacitance value:


C =| Q |2 /( 0 E dv) =| Q |2 /( 0
2

Vmetal

It is easy to see that 0

Vmetal

V Vmetal

Ei dv + 0
2

V Vmetal

4
2
Ei dv = a 3 0 Ei
3

Eout dv)

(12)

, but the calculation of

Eout dv needs to be done in spherical coordinates ( r is the radius, is the polar

angle, and is the azimuthal angle) .


Note that,
2

Eout = Edip =

(3 | p | cos | p | cos ) 2 + (| p | sin ) 2


(4 0 r 3 ) 2

(13)

| p |2 +3 | p |2 cos2
=
16 2 0 2 r 6

where p = 40a3(0)E0/( + 20) = 40a3Ei.


Take the integration in spherical coordinates, we get

Eout dv = 0
2

V Vmetal

a 0 0

| p |2 +3 | p |2 cos2
sin r 2 drd d
16 2 0 2 r 6

| 4 0 a 3Ei |2 8
| p |2
=
=
= 0 a 3 Ei
3
6 a 0
6 a 3 0
3

(14)

Therefore, the capacitance is


C =| Q |2 /( 0

Vmetal

Ei dv + 0
2

V Vmetal

Eout dv)

2
4
8
2
2
= Ei a 2 / / ( a 3 0 Ei + 0 a 3 Ei )
3
3
a
| |2
=
4 2 0
Hence, the circuit parameters of a nanosphere are extracted as

(15)

4
a
1
1
Re , LK =
Im , and C =
| |2
(16)
3 a
4 2 0
3 a


As direct comparison, the previous work by Engheta et al. shows the following circuit parameters [1]:
Rohmic =

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9814

Rsph = (a Im[ ]) 1 , Lsph = ( 2 a Re[ ]) 1 , and Cfringe = 2 a 0

(17)

It is mentioned that the resonance condition for a circuit, LsphCsph = , requires the wellknown condition of plasmonic resonance for a nanosphere Re[] = 20 [1, 14].
Similarly, we check whether our circuit satisfies this resonance condition. Assuming =
+ i,
1
1 a
Im 2 | |2 =
Im *
3 a
3 3 0
4 0
(18)
0
1
Im [i ( + i 0 ) ] =
=
3 3 0
3 2 0
It can be seen that, when = 20, LKC = 2, which satisfies the resonance condition.
Furthermore, Enghetas model has a parallel connection of R, L and C, whose Q-factor is
[15]
4

LK C =

Q=

R
Re[ ]
=
,
L
Im[ ]

(19)

while our model derives that R and L are in series, whose Q-factor is [15]
L
Im[1 / ]
=
R
Re[1 / ]
We test the Q-factor using the Drude model,
Q =

= 0 (1 P2 / ( 2 + i ))
= 0 0P2 / ( 2 + 2 ) + i 0P2 / ( ( 2 + 2 )),

1 / = 1 / ( i ( 0 )) = i / ( 0P2 ) + / ( 0P2 )

(20)

(21)
(22)

where P2 = ne 2 / ( m 0 ) (n is the free-electron density, m is electron mass, e is electron


charge), and is the inelastic electron scattering rate (damping parameter).
Enghetas model gives Q-factor:
Q = Re[ ] / Im[ ] =

2
(1 2 ) 2

P P

(23)

Our model gives Q-factor:


Q = Im[1 / ] / Re[1 / ] = /
(24)
In comparison, Wangs derivation gives Q-factor (independent of the plasmon frequency)
[16]:
Q=

d / d

=
2
+ 3 / 2

(25)

Given that P2 >> 2 and >> , all three models result in the same prediction, i.e. increasing damping factor decreases Q-factor. Our model importantly connects R and L in series, so that they are directly related to the optical resistivity and conductivity of the metal.
4. Circuit model for nanorods
FDTD simulations were performed using Lumerical FDTD Solutions to evaluate the accuracy
of our model. We considered a gold nanorod with hemispherical ends (Fig. 2(a)) whose permittivity is given by the Drude model Au() = 0(1p2/(2 + i)), with p = 9eV, and =

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9815

0.07 eV [17]. The radius (r) was constant at 10 nm, and the length (l) was varied from 100 nm
to 260 nm. In the simulations, we determined the absorption cross-section (abs), scattering
cross-section (scat), extinction cross-section (ext = abs + scat), total electric potential energy
(0|E|2dv), total magnetic energy (0|H|2dv), total charge ((||dv), and total current density (|J|2dv) at resonance. Rohmic, C, LK and LF were calculated according to Eq. (6). Equivalently, the radiative resistance was estimated using Rrad = Rohmicscat/abs. The resonance frequency for the circuit, (LC), and its Q-factor, (L/C)/R, were calculated and compared to
simulated values in Fig. 2(c) and 2(d). The resonance energy shows perfect agreement between FDTD simulations and the RLC circuit model (see Fig. 2(c)). Assuming that the circuit
is driven by an external current source, by considering the power dissipation, Re[Z], where
Z = (RiL)||1/(iC), the extinction spectra of the nanorods were closely reproduced by the
circuit model (see Fig. 2(b)). As the circuit model only works for the fundamental mode, the
high order mode at ~2.3 eV for the longer nanorod was not accounted for.
It is worth noting that our simple derivation using Poyntings Theorem for harmonic fields
does not consider the dispersion in the metal [13], and therefore slightly overestimates the Qfactor (Fig. 2(d)). According to our model, the Q-factor limit of the material is LK/Rohmic =
Im[1/] /Re[1/] = ( 0)/, where and are the real and imaginary parts of the permittivity, respectively. Using the Drude model, this value can be further simplified to /. In
contrast, by accounting for dispersion, this Q-factor limit is given by (d/d)/(2) = /( +
3/2) [16], which converges to the circuit-model predictions when << , + 3/2 . For
the sake of simplicity, we excluded the effect of dispersion in this work. If needed, a more
accurate expression for ohmic resistance with dispersion can be derived from Poyntings
Theorem for dispersive media [13, 18].

Fig. 2. Comparison between FDTD simulation and RLC circuit model. (a) Geometry of the simulated nanorod. (b) Plot of the nomarlized extinction cross-section vs. resonance energy for
two different nanorods. The circuit model can reproduce the spectra by considering the power
dissipation in the circuit. Solid lines show FDTD simulation results for nanorods with 100 nm
(blue) and 260 nm (black). Dashed lines are the reproduced extinction spectra using the circuit
model. (c) Resonance enrgy (Eres) and (d) Q-factor calculated from RLC circuit model (black
solid line) well matches FDTD simulation results (red triangles).

Besides full numerical calculations based on Eq. (6), it is convenient to approximate circuit parameters using simple analytical expressions. More specifically, it can be seen that the
ohmic resistance and the kinetic inductance are functions of the optical conductivity () and
current density distribution (|J|2dv/|I|2). If currents are uniformly distributed in the metal
structures, |J|2dv/|I|2 simply reduces to l/A, the ratio between length and cross-sectional area
of the nanorod. However, this uniform distribution is generally not satisfied; and therefore we

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9816

use an effective length, leff to account for the non-uniformity. As a result, the impedance of the
nanostructure can be expressed as:
Z rod = (1 / )leff / A = Re[1 / ]leff / A + i Im[1 / ]leff / A = R i L

(26)

where leff (l2r)/2, because the fundamental mode approximates a half-wavelength standing
wave of current across the nanorod. Figure 3(a) shows a characteristic current-density distribution from FDTD simulation along the axial direction in a 100 nm nanorod. The Faraday
inductance for a nanorod geometry can be calculated using LF = (0l/2)ln(l/2r) [19]. Figure 3
shows the comparison between the circuit parameters obtained from the full numerical calculation based on Eq. (6) and simple analytical approximations. As can be seen in Fig. 3, simple
formulas give adequate accuracy, and this is particularly useful for a quick assessment of the
resonance behavior for plasmonic resonators. The deviation in the estimate for Faraday inductance in Fig. 3(d) is due to the fact that the equation used to calculate LF from Ref [19]. is for
a perfect cylinder, and does not account for the semispherical ends in the simulated structure.

Fig. 3. Using simple formulas to estimate the circuit parameters. (a) Current inside the nanorod
forms a standing wave for the fundamental mode. (b) Ohmic resistance, (c) kinetic inductance,
(d) Faraday inductance, and (e) resonance energy calculated from full numeric calculation
(black solid lines) and simple analytical formula (red dashed lines) are with good agreement.

5. Circuit model for split-ring resonators


To demonstrate the applicability of our approach to other structures, we extend our model to
split-ring resonators (SRRs) [2023]. Most applications of SRRs are based on the so called
LC mode, whose resonance is sensitive to the gap size. Based on the similar resonance nature between the dipolar mode in a nanorod and the LC mode in an SRR, an SRR is topologically equivalent to a nanorod, i.e. a nanorod can be reshaped to form an SRR [24]. Therefore,
its equivalent circuit can be intuitively constructed by adding a gap capacitance, as shown in
Fig. 4(a). Consequently, the total capacitance of a split ring is the sum of the self-capacitance
(Cself) from the nanorod, and the gap capacitance (Cgap).
We simulated 20-nm wide square cross-sectional strips, as analytical expressions for the
gap capacitances of these structures are readily available. Figure 4(a) shows the simulated
structures. The side length for the SRR was kept constant at 100 nm, and the gap was varied
from 10 nm to 50 nm. For comparison, nanorods of equivalent lengths of 270 to 310 nm were
simulated (the equivalent total length of an SRR is calculated as 4(side length width) gap
size). To verify our circuit model, we calculated the total capacitance for the nanorod and

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9817

SRR using the total charge (||dv) and electric potential energy (0|E|2dv) based on Eq.
(6). The gap capacitances were analytically estimated using a formula adapted from Ref [25]:
Cgap = 0

2 ( h + w)
hw
8l
+ 0 ( h + g + w) + 0
log

g
g

(27)

where h, w, g and l are the height, width, gap size, and side length of the split ring. From Fig.
4(b) we can see that adding the gap capacitance to the capacitance of the equivalent nanorod
gives a good approximation to the total capacitance of the SRR. Furthermore, we show the
comparison for other circuit parameters in Fig. 5. Given the same length, width and height,
and the similar resonance nature of the fundamental modes, we intuit that the kinetic inductance and ohmic resistance of SRRs are the same for that of the nanorods (Fig. 5(a) and 5(c)).
The slight differences of the circuit parameters (C, LK and Rohmic) between the SRR and nanorod may come from the irregular current flow and charge distribution in the SRR caused by
the sharp corners. On the other hand, the Faraday inductance and radiative resistance are
largely dependent on the shape of the resonators. As shown in Fig. 5(b) and 5(d), bending a
nanorod significantly reduces its radiative resistance (i.e. scattering) [24] and Faraday inductance. From the circuit model, the resonance frequency is (LC), and the Q-factor is R1(L/C)

. Because of the significant increase in total capacitance, the resonance of the SRR red shifts
relative to the nanorods. Decreasing gap size increases the capacitance at the gap, and reduces
the resonance frequency (see Fig. 5(e)). Interestingly, although increased capacitance would
reduce the Q-factor, the effect is counteracted by the suppressed radiative resistance of the
SRR (Fig. 5(f)). Similar to the case of the nanorods (Section 4), the Q-factor is overestimated
because dispersion was not considered.

Fig. 4. Extending the circuit model from a nanorod to an SRR. (a) Constructing a circuit model
for SRRs from nanorods with equivelent lengths. (b) The total capacitance for SRRs calculated
from FDTD simulation (black triangles) is approximately equal to the sum (black dot-dash
lines) of the self-capacitance of a nanorod calculated from FDTD simulation (red dashed line)
and the gap capacitance estimated using Eq. (27) (blue dotted line).

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9818

Fig. 5. Comparison of the circuit parameters between SRRs and nanorods with equivalent
lengths. (a) Kinetic inductance and (c) ohmic resistance are comparable for the SRRs and nanorods; while (b) Faraday inductance and (d) radiative resistance are geometry dependent, and
their values reduce significantly after bending. (e) Smaller gaps increase gap capacitance, and
the resoannce red shifts accordingly. (f) According to circuit theory, the Q-factor is calculated
as R1(L/C)1/2. Though SRRs have larger total capacitance, their radiative resistance is significantly suppressed, giving an even higher Q-factor compared to nanorods. [Black solid lines and
dots are for SRRs; red dashed lines and squares are for nanorods; lines are from circuit model,
and symbols are from FDTD simulations.]

The extension from nanorods to SRRs demonstrates that our model can be tailored to account for plasmonic nanostructures with different geometries through intuitive manipulation
of the equivalent circuit. Moreover, the elementary RLC circuit models for single structures
will serve as building blocks for more complex plasmonic structures, such as coupled nanoantennas [11], Fano systems [26], and offer insight to other hybridized plasmonic modes [27].
6. Conclusion
In conclusion, we provide a detailed procedure to extract physical RLC circuit parameters for
plasmonic resonators. Based on Poyntings Theorem, we identified each energy component
and defined its corresponding circuit element. Complete numerical formulas as well as simple
analytical approximations were derived and tested, demonstrating the effectiveness of our
circuit model. Although we considered only nanorods and split-ring resonator structures, our
treatment would enable circuit models to be extracted from more complex plasmonic resonators, and provides valuable prediction of the spectral response due to geometry and material
properties.
Acknowledgments
The authors acknowledge the funding support from Agency for Science, Technology and Research (A*STAR) Young Investigatorship (grant number 0926030138), SERC (grant number
092154099), and National Research Foundation grant award No. NRF-CRP 8-2011-07. The
authors thank Huang Shaoying from the Singapore University of Technology and Design for
fruitful discussions.

#207066 - $15.00 USD


(C) 2014 OSA

Received 25 Feb 2014; revised 2 Apr 2014; accepted 3 Apr 2014; published 16 Apr 2014
21 April 2014 | Vol. 22, No. 8 | DOI:10.1364/OE.22.009809 | OPTICS EXPRESS 9819

Vous aimerez peut-être aussi