Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Modeling Nanomaterial Physical Properties Theory A

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS

2019, VOL. 10, NO. 2, 116–143


https://doi.org/10.1080/19475411.2018.1541935

ARTICLE

Modeling nanomaterial physical properties: theory and


simulation
Tanujjal Boraa,b, Adrien Doussea,b, Kunal Sharmaa,b, Kaushik Sarmac, Alexander Baevd,
G. Louis Hornyaka,b and Guatam Dasguptae
a
Nanotechnology, Industrial Systems Engineering, School of Engineering and Technology, Asian Institute
of Technology, Pathumthani, Thailand; bCenter of Excellence in Nanotechnology, Asian Institute of
Technology, Pathumthani, Thailand; cMechatronics, Industrial Systems Engineering, School of Engineering
and Technology, Asian Institute of Technology, Pathumthani, Thailand; dInstitute for Lasers, Photonics and
Biophotonics, State University of New York (SUNY) at Buffalo, Buffalo, NY, USA; eDepartment of Civil
Engineering and Engineering Mechanics, Columbia University, New York, NY, USA

ABSTRACT ARTICLE HISTORY


A brief theory and simulation overview for the purpose of design is Received 10 August 2018
presented with examples applies to modeling the physical properties, Accepted 24 October 2018
behavior, and phenomena of nanomaterial. This review paper con- KEYWORDS
structs perspectives that consider coupling traditional domains of Nanomaterial; multiphysics;
simulation by novel pathways to produce accurate representations finite element; molecular
of nanomaterial properties, behavior and phenomena. It is all about dynamic; density functional
size scaling and how different approaches are able to simulate, inte- theory
grate or simply pass the baton to the next level of complexity. In
macroscopic world, the atomic or molecular information alone may
not be directly useful. Nor is the bulk information useful in the micro-
scopic world without intimate knowledge of molecular makeup.
Therefore, when designing Nanomaterials, knowledge of properties
spanning the complete range of size is the prerequisite of
a recommended self-consistent approach. In fact, regarding applica-
tions in both industry and academia, the simulation first approach
often can lead to great savings in time. This review paper focuses
mostly on optical and electronic properties but a section is added that
provides a segue into mechanical properties for future consideration.

Introduction
Material simulation methodologies have become powerful tools for scientists and engineers
especially now with the advent of faster computing capability and sophisticated computer
programs. The complete spectrum of size dependent properties and behaviors can be
modeled with great confidence. Ideally, and theoretically, modeling everything from first
principles would be a boon to the material simulation community, however, due to limited
computing capability, this approach is still off into the future. Nonetheless, several bold
computational and numerical (and hybrid) methods, originating from dramatically differing

CONTACT Tanujjal Bora tbora@ait.ac.th Nanotechnology, Industrial Systems Engineering, School of Engineering
and Technology, Asian Institute of Technology, Pathumthani 12120, Thailand; Guatam Dasgupta gd3@columbia.edu
Department of Civil Engineering and Engineering Mechanics, Columbia University, New York, NY 10027, USA
© 2018 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/
4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 117

perspectives based on bold assumptions, have been developed to circumvent the aforemen-
tioned limitations.
Significant progress has been made over the past decade in coupling classical Maxwell’s
equations to other established simulation approaches, especially in the area of predicting
nanoscale plasmonic and optical properties [1]. Maxwell-Schrödinger equation hybrid
systems have been created and solved by a unified Hamiltonian approach capable of
simulating classical electromagnetic fields and particles [2]. Authors of this paper proposed
a coupled method based on a unified Hamiltonian of electromagnetic and quantum
mechanics. The numerical methods of modified Finite Difference Time Domain (FDTD)
was then applied to overcome discrepancies between wavelength dependent electromag-
netic properties and those of electron waves [2]. The result yields a useful hybrid method
with coupled classical electromagnetic, quantum mechanics and numerical methods. This
method was then used to simulation nanomaterials and nanodevices, including those
based on quantum mechanics. In particular, the coupling of finite element methods
(FEM) with classical Maxwell’s equations based methods leads to the direct computation
of Maxwell equation, which has made a headway into modeling nanomaterials [3]. Recent
developments in nanoscience and nanotechnology, among other fields, ‘have made it
possible to conceive materials and structures with atomic level controllability and with
unprecedented properties. . .. These developments have opened doors to shape and sculpt
light at the nano-, micro and mesoscales in a desired fashion’ [4]. A new numerical method,
‘Extended’ Finite-Difference Time-Domain (EFDTD) method, has been developed to analyze
nanoscale systems [5]. EFDTD is used to solve Maxwell’s equations in combination with
Schrödinger’s equation to evaluate quantum effects such as those encountered in electron
tunneling through potential barriers. For future nanoscale circuit design and analysis, an
important role would be played by such approaches that can consider all physical inter-
actions to compute and model [5].
There are a wide variety of programs at hand for the researchers and engineers, both
commercially available and via open source. These include cross-platform finite element
programs for multi-physics simulation that are capable of solving for coupled phenom-
ena. For example, fluid dynamics coupled with heat transfer. Current molecular domain
simulations are capable of linking material atomic and molecular structure with their
microscopic properties and behavior. Varied simulation approaches based on funda-
mentally different premises can effectively work together to model material behavior,
especially in the case of nanomaterials. For example, classical electromagnetic (EM) and
thermodynamic theories can provide inputs for methods that have completely divergent
and unrelated theoretical roots; for example, inputs into numerical convergent methods.
Quantum physics based methods such as density functional theory (DFT) are capable of
providing such fundamental data into classical EM equations. Here we show examples of
such coupled and/or convergent efforts in the simulation of optical properties and of
plasmonic systems. Simulations of mechanical systems, which is deeply rooted in finite
element methods, are provided at the conclusion of the paper.

Size domains
Classical approaches of course are still valid as stand-alone methods that can adequately
predict nanomaterial properties, optical properties in particular. We consider three
118 T. BORA ET AL.

conventional and conveniently defined size domains. Macroscopic material properties


can be described by either classical geometric optics or electromagnetic wave theory
without input of atomic or molecular information or consideration of localized material
interactions with photons and electrons [6]. We and many others have shown that
classical electromagnetic approaches can be very useful in predicting nanoscale material
optical properties through the lens of macroscopic observables – i.e. as in the case of
composite materials [7,8]. Effective medium theories and scattering matrix approxima-
tions, applied without further input into finite element systems, have proven to be of
great value in predicting optical properties of composite materials and periodic arrays,
photonic structures, respectively [6]. Therefore, to a point defined by macroscopic size
limits, classical representations are completely valid and useful.
Regarding the microscopic perspective, atomistic-molecular approaches are tradition-
ally represented by ab initio, molecular dynamic, density functional theory (DFT), Monte
Carlo and tight-binding combinations among each other – essentially modeling physical
properties from the bottom up. In principle, information based within quantum
mechanics contains all relevant information about any material system. However, the
exact analytical solution to the Schrödinger’s equation are almost impossible for any
N-body system and we have to invoke approximations to help us along. DFT in particular
offers such an approach to solving Schrödinger’s equation for multiple-particle systems
by providing an approximate method that correlates motions of electrons. The non-
relativistic time-independent Schrödinger equation is given as,
^ jψi ¼ Ejψi;
H

where H^ is the Hamiltonian or total energy operator, jψi is the wavefunction describing
the system and E is the energy. DFT reformulates the Schrödinger equation into the
Kohn-Sham equations:
 
h2 2
 Ñ þ Veff ðrÞ ϕi ðrÞ ¼ εi ϕi ðrÞ;
2m

where fϕi ðrÞg is a set of single particle orbitals representing fictious non-interacting

electrons. The ground state electron density calculated from the orbitals ρðrÞ ¼

P
N
jϕiðrÞj2 is a central quantity in DFT. It uniquely determines the Hamiltonian operator
i
from which all the properties of the system’s ground state can be deduced. The Kohn-
Sham effective potential is given by:
δJ½ρðrÞ δEXC ½ρðrÞ
Veff ðrÞ ¼ VNe þ þ
δρ δρ
In this formulation the total energy can be written as:
E½ρðrÞ ¼ VNe ½ρðrÞ þ TS ½ρðrÞ þ J½ρðrÞ þ EXC ½ρðrÞ;

where VNe ½ρðrÞ is a functional representing the nuclear-electron attraction potential


energy, TS ½ρðrÞ is the exact kinetic energy of the non-interacting fictious reference
system with the same density as the real interacting one. J½ρðrÞ represents the classical
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 119

contribution to the electron-electron interaction. And EXC ½ρðrÞ is the correlation-


exchange energy. The Kohn-Sham equations are exact, however, DFT is necessarily
made approximate by the fact that the explicit and exact form of the exchange correla-
tion functional is not known. However, accurate approximations for EXC are possible and
eventually yield correct quantitative results.
From such DFT approximations, structural and electromagnetic characteristics of
materials can be investigated. As such, DFT has become one of the most popular and
resourceful methods in condensed matter and computational physics. In fact, DFT has
successfully provided quantum modeling of optical properties of nanomaterials as well
as structural information relating to phase diagrams [9]. However, DFT is still compli-
cated and therefore the FEM is introduced to simplify the simulation. FEM and facilitative
adapted versions are now used solve Kohn-Sham equations of DFT to provide realistic
three-dimensional representations [9].
Lastly, at the mesoscopic domain within which nanomaterials reside, applies electro-
magnetic tools that are superimposed upon some preconceived nanoscale scaffolding
[10]. Particles ranging in size from 1 to 100 nm are considered to be nanoparticles.
Nanomaterials have a countable number of atoms and are Par Excellence substrates for
current N-body problem solving programs. They also are well-suited to occupy the
position of the fictitious mesoscopic bipolar cavity envisioned by early electromagnetic
theoreticians to justify linking Maxwell’s equations to the macroscopic world. As stated
earlier, atomistic-molecular parameters are not always considered in simulations. Input
of bulk material derived optical constants are routinely input into classical (and FEM)
programs from which nanomaterial properties are predicted. However, for decades,
combined methods utilizing molecular dynamic simulations and the finite element
method have begun to provide simultaneous resolutions at the atomistic-molecular
and continuum field scales [11]. Furthermore, finite element methods have been applied
to the atomic scale via ‘multiscale simulations’ in the development of nanotechnology
by providing seamless and facilitated computation [12].
Finite element numerical methods, assuming importance in the early seventies, were
derived mainly to solve macroscopic engineering problems [13]. However, today, finite
elements can be applied even down to the molecular scale. Natural phenomena show
multiscale character in space and time. ‘Continuum-on-atomistic’ multiscale coupling
methods have achieved great popularity today [14]. Actually, when considering electro-
magnetic input parameters, finite element methods are able now to stand alone when
describing optical properties of nanoscale structures [15].

Optical properties of nanomaterials by classical EM and FEM


Classical treatments
Efforts have been made since Maxwell’s equations were established to explain electro-
magnetic theory from the atomic level – i.e. to link the microscopic with the macro-
scopic. However, at the microscopic level, molecular dipole fields are discrete (not
smooth) and variant over molecular dimensions and timescales – not exactly suitable
for treatment by classical electromagnetic theory. Therefore, the fictitious mesoscopic
cavity was invented to average molecular microscopic fields over larger size scales and
120 T. BORA ET AL.

lengthier timescales – albeit with dimensions significantly less than of a macroscopic


material [6]. Resultantly, the constitutive material equations are required to link
Maxwell’s equations to the macroscopic world via the relative dielectric function r .
Following O-F. Mossotti’s view of dielectric behavior, Maxwell devised the displacement
current and hence the space and time variant displacement field D
Dðr; tÞ ¼ o Eðr; tÞ þ Pðr; tÞ ¼ o r Eðr; tÞ
where D is the displacement field vector, E if the electric field and P is the polarization
density vector. Maxwell’s equations then are able to provide an interface between the
extreme domains through the dielectric constant by way of smooth shapes (spheres or
ellipsoids of revolution), the mesoscopic cavity. For the optical range, this cavity had to
be much smaller than the wavelength of light (hence, quasi-static, at the infinite
wavelength limit). Details of quantum electronic nature are not required in this mode
of treatment.
When modeling nanomaterial optical properties, certainly over the past several
decades, optical constants n (refractive index) and k (extinction coefficient) were nor-
mally extracted from well-established data tables [16]. Data from Johnson & Christy’s
famous work were obtained by measuring reflection and transmission in vacuum on
evaporated thin films with thickness ranging from 20 to 50 nm [16]. These data, though
excellent in providing approximate values, are held in question when directly applied to
nanomaterials. Not only do optical constants differ according to method of surface
preparation, for example, optical constants (n and k) determined by ellipsometry on
electroplated gold surfaces; but in addition, they are influenced by intrinsic (quantum)
size effects [17]. Optical constants of gold nanoparticles embed within titania matrix
were extracted by researchers using spectroscopic ellipsometry [18]. Their model was
able to demonstrate anomalously high near-infrared absorption by such small nanopar-
ticles, when optical constants from data tables were applied in simulations. Furthermore,
semiconductor nanoparticles are expected to undergo optical bandgap variation upon
diminishing size [19]. Semiconductor band gap energy Eg expands as particle size is
diminished leading to concomitant changes in refractive index and absorbtivity.

Effective medium treatments


Effective medium approximations or theories (EMT) are capable of describing macro-
scopic properties of composite materials based on proportional blending of optical or
electronic properties of components [20]. They are considered to be self-consistent,
phenomenologically based continuum averaging techniques. Therefore, EMTs provide
approximations of composite material behavior that are based on the relative volume
fractions of its components [21]. EMT’s are rooted within classical physics, in particular,
classical electromagnetic theories. In the middle of the 19th century, Ottavio-Fabrizio
Mossotti (1850) developed the first form of the expression later to be modified by Rudolf
Clausius (1879, the explicit form) to yield the well-known Clausius-Mossotti relationship –
later to be embellished further by Ludvig Lorenz and Hendrik Lorentz (polarization
leading to local field correction). The host medium was assumed to be vacuum or air
in those expressions and mainly applicable for spherical particles or cavities. Today,
numerical techniques based on discrete dipole approximations evolved from the
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 121

Clausius-Mossotti relationship. These EMT homogenization techniques were based on


classical and macroscopic constitutive relations derived from Maxwell’s equations [22].
Modern theories of polarization address computations of induced microscopic currents
in condensed media by application of density functional theory [22,23].
Two general forms of EMT are commonly used in simulations: the Maxwell-Garnett (1904)
and the Bruggeman homogenization forms. In the former, volume fraction of the composite
inclusion is expected to be small with respect to the host medium (usually a dielectric
material). In the latter, volume fractions of inclusion and host are on par. The Maxwell-
Garnett and Bruggeman are widely known and need not be displayed at this juncture.
We have used EMTs extensively in the past to model extrinsic size effects [24]
however this simulation method is limited with regard to accurately quantifying intrinsic
(quantum) size effects [7]. Extended effective medium theories try to incorporate size-
dependent electromagnetic effects like dynamic depolarization, extinction and retarda-
tion effects. Mie scattering theory is best applied to homogeneous spherical particles.
Extrinsic size effects are those governed by particle size, shape and orientation with
respect to the wavelength of light [24]. In other words, the optical constants n and
k (sequestered with the dielectric constant ) input into classical EMT are simply those of
the bulk material and are functions of the wavelength of the light,
 ¼ ðλÞ with d > 10 nm. Regarding intrinsic size effects, applicable to particles less
than 10 nm radius,  ¼ ðλ; dÞ, optical constants are expected to be a function of size
as well [24].

Maxwell-Garnett EMT and FEM


Many years ago, we conducted simulations from programs that we generated in
MATLAB and other platforms. However, our programs were limited to spherical or
ellipsoidal shapes due to the necessity of maintaining continuity within Maxwell’s
equations. Now with the adaptation of finite element methods, we can input equations
as well as multiple and variable shape data directly into FEM programs (e.g. COMSOL
Multiphysics with wave optics module). Here we show a simple example of facilitated
simulations created by FEM. In Figure 1, absorption spectra of spherical to cylindrical
gold nanoparticles embedded homogeneously within an isotropic porous alumina
parallel channel matrix, calculated by dynamical Maxwell-Garnett approximation, are
input into an FEM program [20,25]. Input parameters included length a (varied from 20
to 100 nm), diameter b (fixed at 20 nm), and volume fraction fi (fixed at 0.05). The
orientation of the particle array (i.e. the pore channels in the alumina membrane) was
held perpendicular to the electric field of the incident light. It is clear from the simula-
tion shown in the figure that a blue shift in absorption occurs with increasing aspect
ratio. Application from this kind of result can be useful in designing filters and wave-
guides and made facilitative by application of FEM to classical electromagnet equations.

Bruggeman EMT and FEM for optical properties- a case study


A wide variety of computational methods have been developed and implemented to
simulate the properties of anti-reflective coatings [26]. Effective medium theory (EMT)
[27], finite-difference time-domain (FDTD) [28], transfer matrix method (TMM), rigorous
122 T. BORA ET AL.

Figure 1. COMSOL Multiphysics FEM simulation results of absorbance spectrum varying length (a) of
gold nanoparticle as inclusion in alumina matrix. The transverse dimension (b) was held constant at
20 nm diameter. Volume fraction was fixed at 0.05 metal inclusion in the aluminum oxide host.
A blue shift in absorbance is associated with increasing aspect ratio.

coupled-wave analysis (RCWA) [29,30] and the finite element method (FEM) [31], are
commonly used for predicting the optical properties of solar cell applications for
example. The OPTOS (optical properties of textured optical sheets) simulation formalism
[32–34], recently developed by the Fraunhofer Institute, is a more advanced computa-
tional technique dedicated to simulating textured surfaces with low computational
costs. Here we detail a simple method based on the well-known Bruggeman effective
medium approximation to model randomly patterned nanostructured layers acting as
anti-reflection coatings for a simple air/glass interface and make predictions for their
optical properties. Figure 2 depicts examples such layer.
The characteristic length of the surface roughness, or equivalently the average
distance between particles, a is taken to be less than that of the wavelength of the
electromagnetic radiation of interest. This condition is necessary for effective med-
ium theories to apply. In this limit, the subwavelength nanostructures cannot be
resolved by the electromagnetic waves: as the light traverses the nanolayer it will
interact with the nanostructured layer as if it was a continuous stack of infinitely thin
homogeneous sublayers, each of which being characterized by an effective refractive
index. In other words, the nanolayer can be modelled as an inhomogeneous layer
with a gradually varying refractive index (Figure 3). In this regime scattering is
negligible.
The nanolayer acts as an impedance matching layer for the light. This gradual
transition from a medium of low refractive index (the incident medium: air) to
a medium of higher refractive index (the substrate: glass) reduces reflection losses
through a series of destructive (constructive) interference of the waves being reflected
(transmitted).
The representative unit elements of the nanolayers studied here are shown Figure 4.
They are characterized by the following nanoparticles profiles that we define as: linear,
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 123

Figure 2. Cross-section view of randomly distributed nanoparticles on a glass substrate.

Figure 3. Continuous stack of infinitely thin homogeneous sublayers or equivalently, a single layer,
inhomogeneous in refractive index.

Figure 4. Representative unit elements of each layer: the nanoparticles with linear (a), quadratic (b),
and elliptical (c) profiles are shown together with their respective bounding box representing the
volume of the unit element.

quadratic, and elliptical. The aspect ratio for each nanoasperity is fixed to 2. The particles
are distributed randomly with an average separation a (Figure 5).
To calculate the effective refractive indices, the Bruggeman formulation requires the
knowledge of the dielectric functions of the nanoparticle material and the surrounding
medium, a shape factor characterizing the aspect ratio and interaction of the particles
124 T. BORA ET AL.

Figure 5. Dimensions corresponding to the linear (a), quadratic (b), and elliptical (c) profiles.

with the field: the depolarization factor. It also requires the volume fraction f of the
material considered.
! !
air   SiO2  
fair   þ fSiO2   ¼0
air þ Pi1  1  SiO2 þ Pi1  1 

where fair and fSiO2 are the volume fractions of air and silica respectively, air and SiO2
their dielectric functions, Pi are the depolarisation factors and  the effective dielectric
pffiffi
function from which we will obtain the effective refractive index n using n ¼ , as none
of the materials considered here absorb or exhibit magnetic properties.
Here, the volume fraction for each particle profile is a function of the height h and can be
calculated by simple geometrical considerations (Figure 6). Since the nanoparticles have
cylindrical symmetry we can define the functions rl , rq , and re describing the height-
dependent cross-sectional radii for the linear, quadratic and elliptical profile cases respectively:
a
rl ðhÞ ¼ ð1  h=aÞ
2

a 
rq ðhÞ ¼ 1  ðh=aÞ2
2
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a
re ðhÞ ¼ 1  ðh=aÞ2
2
For an infinitesimal slice of height dh the volume of material at height h is then
π ðri ðhÞÞ2 dh, and the corresponding total volume of material and air is ð2ri ð0ÞÞ2 dh, so
that the height-dependent volume fractions are then given by: fi ðhÞ ¼ π ðri ðhÞ=2ri ð0ÞÞ2
with i ¼ l; q; e.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 125

Figure 6. The calculated volume fractions for the linear (a), quadratic (b), and elliptical (c) profiles as
a function of layer height. The continuous curves are approximated by a discontinuous gradient,
composed of 4 steps of equal height: each step represent the layers used to simulate the ARCs.

Together with the refractive index of the base material considered, the functions f are
the only parameters that determine the anti-reflective properties of the layer.
We will approximate the modelling of the surface by decomposing the nanolayer using
only a finite number of sublayers, instead of a continuum (see Figure 7). This assumes that
the volume fraction of silica can be considered to decrease in a stepwise manner as
a function of height. To simplify the calculation of the depolarization factors, at each
sublayer we will also assume that the geometry of the material (e.g. a truncated cone for
the linear profile case) can be replaced by a sphere with equivalent volume fraction. The
depolarisation factors Pi of a sphere are all equal to 1=3, we then obtain the following
simplified Bruggeman expression:
   
1 SiO2  
fair þ fSiO2 ¼0
1 þ 2 SiO2 þ 2

The resulting height-dependent refractive indices are shown Figure 7. The gradient in
refractive index, between the air/glass interface, although discontinuous, typically allows
for a higher transmission of light, as will be discussed below.
To simulate the optical properties of the substrate/sublayers system, the Fresnel
equations may be used in combination with a transfer matrix method [35]. Another
method is to input the model into a FEM package, such as COMSOL and to solve for the
transmitted and reflected fields using Maxwell’s wave equation. As a general require-
ment for EM FEM calculations, a fine enough meshing (<λ/10) should be used as well as
a large enough computational domain (≈λ) (Figure 8) so that numerical artefacts such as
unwanted boundary reflections can be avoided. In our FEM simulation, the incidence of

Figure 7. The calculated effective refractive indices for the linear (a), quadratic (b), and elliptical (c)
profiles as a function of layer height, for a wavelength of 650 nm.
126 T. BORA ET AL.

Figure 8. FEM simulation domain.

the electromagnetic field is kept normal to the surface. The total height of the sublayer
stack has been set at 400 nm, bearing in mind that for the visible part of the solar
spectrum as a light source, the accuracy of the model is higher for longer wavelengths.
The refractive index gradient allows to minimize reflectance losses for all 3 profiles as
compared to bare glass (Figure 9). The fashion in which the mixture of air with silica is
increased from substrate to air has a strong impact on overall reflectivity. Numerical and
experimental studies of multilayer graded-index profile have shown that a quintic or 5th order
polynomial profile for the refractive index results in optimum anti-reflective properties [36–
38]. In the present case however, the relative performance of the modelled layers seems to
follow the refractive index contrast at the first sublayer/air interface. The elliptical profile
model results in a more severe contrast in which case we would expect, as is the case, a higher
proportion of light being reflected. For the quadratic profile, as the first step presents
a smoother transition into the medium, lower reflectivity is obtained. This effect is also
accompanied by less interference patterns (oscillations in the reflectance spectra due to
Fresnel reflections). The linear case presents to softest transition for the light, from air to
medium, and results in no interference pattern and a transmission enhancement > 4% across
a wide bandwidth of the spectrum.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 127

Figure 9. The reflectance spectrum of the system for the linear (black squares), quadratic (red circles)
and elliptical (blue triangle) profiles. The dashed line averaging approximatively 4% across the
spectrum gives the reflectivity of glass with nglass ¼ 1:5.

Multiscale theoretical frameworks


Material simulation overview
Simulations of nanomaterials’ optical response essentially follows historical development
of computational materials sciences, starting with classical Maxwellian electrodynamics
and Lorentz oscillator model, through more refined semi-classical approaches. Including
both wave function and electron density-based approaches, they can provide dielectric
function beyond Drude approximation, through various homogenization models, culmi-
nating with fully atomistic quantum methods. The hierarchy of state-of-the-art modeling
approaches therefore retains this historical development with full-wave classical electro-
dynamics modeling, applicable to nanoparticles and nanostructured materials on top,
fully atomistic time-dependent methodologies for calculating optical properties of small
clusters at the bottom, with the development of dielectric function beyond classical
electrodynamics approaches in-between (Figure 10). The ultimate goal is to develop
a multi-scale theoretical framework that integrates phenomenological models describ-
ing the essential physics of light-matter interactions with full-wave electromagnetic
analysis and comprehensive microscopic modeling.
To analyze optical properties of materials of interest one has to transform information
obtained at the quantum level to relevant macroscopic properties for using with the
Maxwellian constitutive relations. In general, three different tiers of theoretical modeling
can be applied to undertake this non-trivial task: (i) modeling of intermolecular interac-
tions and their influence on the property; (ii) linear scaling methodology, and (iii)
multiscale modeling. The first approach involves models based on reaction field theory
and polarizable continuum methods. Linear scaling is more elaborate method in which
rigorous quantum mechanics is transcended by successive enlargements of the mole-
cular models (the number of atoms) until the property value is converged. Ideally linear
scaling methods can be implemented to be proportional to the system size, thereby
extending the applicability range into the nanoscale regime, beyond ten thousand
atoms. This is not achieved merely through high-performance computing, but rests
128 T. BORA ET AL.

Figure 10. Hierarchy of systems and modeling methods.

critically on the development and adaptation of new computational schemes. Another,


conceptually different approach to close the gaps between micro- and macro-scales is
offered by multiscale modeling technology. The most important variation of contem-
porary multiscale modeling from atomic size elements to macroscopic homogenous
media is given by the combination of quantum mechanics and extended classical
physics models. This theoretical framework endows various models with rigorous
insights extracted from the quantum nature of materials at the atomic scale and builds
them into equations that are solved by advanced numerical techniques. In particular,
joining quantum mechanics (QM) with molecular mechanics (MM) has become an
important and popular area in ‘in silico’ research across a wide variety of fields (QM/
MM). The multiscale modeling paradigm is sketched in Figure 11.
The calculations of optical properties at the quantum level have extensively utilized
the analytical response theory [39]. Within the response theory paradigm, excitation
energies and excited state properties are generated by applying time-dependent per-
turbation theory, but without explicitly solving the Schrödinger equation for each
excited state. Response functions can be obtained with either the Ehrenfest method
or quasi-energy method [40]. Thus, polarizability tensors are computed directly for given
frequencies without approximations other than those, dictated by the computational
model. The formalism is completely generic with respect to the source of the perturbing
fields (external or internal) and type of field (electric, magnetic or mixed), and time
dependence (time-dependent or time-independent fields). The response theory yields
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 129

Figure 11. Coupling of theoretical tiers in multiscale modeling progresses from the quantum regime
to the molecular mechanics (MM) regime to plasmonic environment; the nanosize objects are
embedded in a dielectric host material with local reaction fields at the top level, where electro-
magnetic interactions between the active material and dielectric inclusions are incorporated into the
simulations.

the real and imaginary parts of molecular polarizabilities directly as a function of the
excitation frequencies without recourse to sum-over-states (SOS) formulation. One
critical extension of the response theory is the complex polarization propagator
approach [41], which is a resonant convergent response theory that accommodates
resonant properties.
An example of using fully atomistic description is ultra-small metal nanoparticles
(clusters). Such noble metal clusters can maximize direct enhancement of optical proper-
ties of bound molecules while minimizing losses, inherent to larger size noble metal
nanoparticles. In fact, small size clusters do not support plasmon resonance. Thus,
optical losses associated with plasmon resonances are not an issue. For the smallest
particles, which range from well-defined clusters of 4, 11, 20, 25 or 38 gold atoms [42–
45] to a few hundred atoms, both the absorption and field enhancement become
increasingly sensitive to the environment, and the concept of field enhancement
based on plasmonic behavior and the Drude model breaks down. Fully atomistic
description of gold atoms requires incorporation of relativistic effects into the calcula-
tion, which can be done at the minimal cost with the use of effective core pseudopo-
tentials [46].
One recent extension of the hybrid QM/MM approach is a capacitance-polarization
force field, used for simulating the optical response of metal nanostructures in complex
environments [10]. To date, such multiscale integrated approaches have mostly been
applied to fully organic systems, while organometallics, or systems incorporating nano-
particles with free charge carriers, have received relatively less attention. However,
nanoparticle-enhanced materials show unique capabilities for a wide range of applica-
tions in sensing, imaging, molecular electronics and energy conversion – within or
outside of the plasmonic regime. An obvious complication of metallic particle embed-
ding is the almost free motion of charges within the metallic cluster, and that the
130 T. BORA ET AL.

interaction with the QM shell should be regulated by ‘atomic capacitances’ rather than
by fixed charges in addition to polarizabilities. The so-called quantum mechanics/capa-
citance molecular mechanics (QM/CMM) model provides a practical approach for theo-
retical modeling of optical properties of molecular moieties interacting with inorganic
surfaces [47,48]. Here, the heterogeneous MM part is split into metallic and non-metallic
parts, assuming a capacitance-polarization model for the electrostatics and polarization
of the metallic part (resistance and inductance) and distributed charges for the non-
metallic part. The implementation of a heterogeneous MM region captures the essential
physical features of a material containing metal nanoparticles and molecules interacting
with them.
Several obstacles have yet to be overcome to reach the full capacity of the QM/CMM
method. When coupled to the response theory, it can effectively model the full spec-
trum of properties of hybrid inorganic:organic nanocomposites . However, the homo-
genization of force fields used in the dynamics and property parts of the calculations
over the entire system represents a substantial challenge. The bond capping, charge
balance and treatment of short range interaction and charge transfer between the QM
and MM systems are notoriously challenging issues. Recent developments of computa-
tional methods and techniques are targeting these challenges. For example, fully
atomistic treatment of plasmonic nanoparticles has become a reality thanks to novel
algorithms and increased computing power [49].
On the other end of the spectrum, the full-wave classical electrodynamics algorithms
use both differential and integral equation based methods and acceleration approaches,
to predict the high-frequency electromagnetic response of nano- and mesoscale materi-
als [50–58]. In particular, the finite difference frequency domain (FDFD), finite element
analysis (FEA), discrete dipole approximation (DDA), and extended beam propagation
methods (BPM) that incorporate forward- and backward-propagation effects can be
used to compute key measureable quantities such as transmittance and reflectance
spectra. Many of them are integrated into state-of-the-art commercial software packages
(e.g. COMSOL [59]). Many of these packages allow for custom modifications comprising
extended dynamical equations that describe quantum aspects of nanoscale material
behavior.
Refining the dielectric function of plasmonic materials, particularly emerging and alter-
native materials, beyond the limits of conventional classical electrodynamics-based models,
bridges quantum and classical approaches and allows to obtain fundamental new knowl-
edge of material responses at the nanoscale. Over the decades, technological progress has
led to a steady miniaturization that has resulted in features with near-atomic size. However,
for the classical dielectric function to remain a valid concept, the size of the cluster/particle
should be sufficiently large, larger than the mean-free path of charge carriers. Classical
electrodynamics is based on averaging process that turns the rapidly fluctuating micro-
scopic fields found near individual atoms into macroscopic fields averaged over a volume of
space that may contain myriad of such atoms, or dipoles, so much so that the medium loses
its granularity and becomes a continuum, necessitating only the mere application of
boundary conditions [60]. Obviously, this simplified picture fails if the macroscopic theory
is applied to systems with features that are only a few atomic diameters in size. This is
already the case for typical nanowire and/or nanoparticle systems that are now easily
fabricated with features so small and so closely spaced that the electronic wave functions
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 131

spilling outside their respective surfaces may begin to overlap. The diameter of a typical
noble metal atom is approximately 3 Å, while the electronic cloud forming at and shielding
a flat, noble metal surface may extend several Angstroms into free space. The linear optical
response of metals in classical electrodynamics is almost exclusively modeled under the
assumption that a metal (especially a noble one) can be represented by the ionic core
surrounded by a cloud of free electrons that is driven by an incident electric field. The
resulting dielectric response depends only on the frequency of the incoming light. However,
this model is inadequate to describe experimental observations. It has been enhanced by
hydrodynamic models that incorporate nonlocal effects, through terms like electron pres-
sure [61–66] and surface and bulk nonlinearities [67–70]. Ultimately the sub-nanometer gap
between metals enables quantum tunneling [–41,71–79], which also is incorporated into
the dynamical models [74].
Finally, effective medium analysis or homogenization can be further applied to obtain
effective frequency-dependent bulk dielectric function that can be used for larger scale
(macroscopic) material, device, and systems design. The effective medium analysis can
identify optimum attributes of constituents such as size, spacing, ordering, and orienta-
tion that best enhances the target property. Here, the basic quantum description of
electronic structure can be combined with nonlinear Maxwell’s equations and the
density matrix approach. In this way one obtains not only a connection to field interac-
tion with the macro-structure of the bulk material, but also how the ingoing fields
transform to local fields, which is crucial to properly assess experimental results. Effective
medium theories such as the Maxwell-Garnett approach are of great use here to obtain
macroscopic bulk effective parameters of a composite material from its microscopic
characteristics.
It has to be mentioned that traditional plasmonic materials, such as gold and silver
(less so platinum and aluminum) have long been the ‘golden’ standard for applications
utilizing the operation wavelengths falling into the visible range. However, it has been
realized that the intrinsic lossyness of these materials impedes their application in the
emerging fields such as metaphotonics [80]. Recently, alternative low-loss plasmonic
materials have emerged such as titanium nitride (TiN), indium tin oxide (ITO), Al:ZnO
(AZO), In:CdO, B-hyperdoped Si, and copper chalcogenides [81]. They allow for tunable
local field enhancement in the short wave infrared region (SWIR). These materials
generally exhibit a smaller imaginary contribution to the dielectric function than noble
metals, especially in the IR range. Modeling of the alternative materials and devices
based on them adheres to the strategies outlined above.

Practical applications of simulations


Anti-reflection coatings
We return to discussions about anti-reflection coatings (ARCs). Anti-reflection coatings
are technologically important subwavelength structures, that are also found naturally.
The eyes of 19 diurnal butterfly species were analyzed by Stavenga et al. [82] and
reported to have arrays of periodical conical structures each having height about
230 nm acting as perfect anti-reflecting layer providing protection from predators.
Today several nanostructures with variety of shapes have been produced artificially to
132 T. BORA ET AL.

explore their application as ARC in many areas, such as solar cells, optical lenses,
automobiles etc [83,84]. Optical modelling methods have developed rapidly in recent
times with advanced computational ability and has been applied widely for ARC design.
Amongst numerous modelling methods, finite element method (FEM), finite-difference-
time-domain (FDTD), transfer matrix method (TMM) and rigorous coupled-wave analysis
or Fourier modal method (RCWA/FMM) are the commonly used methods in the field of
optical modelling for ARC [85]. Effective medium theory (EMT) is another method widely
used for determining the effective index of refraction of the subwavelength ARC
structures [86]. But EMT is limited to structures much smaller (typically one tenth) of
the light wavelengths [87], and hence needs to be considered carefully.
Several researchers have used FDTD methods for optical modelling of ARC, mostly
when considering ARC based on optically thin dielectric materials. Lesina et al. [88] has
reported design of ARC for silicon solar cells based on plasmonic silver (Ag) nanoparti-
cles with silica overlayer. Using FDTD methods they have investigated the effect of
nanoparticle size, their location within the ARC layer and surface coverage on the ARC
performance. Feng et al. [89] has compared FDTD method with TMM method for multi-
ple thin film based ARC and reported that FDTD methods have better accuracy than
TMM. Similar comparisons have been reported by other researchers as well [90,91]
showing FDTD with higher accuracy over TMM methods. However, FDTD methods are
more time consuming compared to TMM methods.
Unlike FDTD method, which is a time domain optical modeling method, FEM is
a frequency-based optical modeling method also reported widely for ARC modelling.
High accuracy and use of boundary conditions reducing the unwanted edge reflec-
tion are some of the advantages of FEM [92,93]. FEM is also capable of multiphysics
modelling. Lee et al. [94] has reported a 2D FEM simulation for silicon nitride (Si3N4)
as ARC and examined the geometry-dependent effective reflectance of the Si3N4 sub-
wavelength structure over the wavelength ranging from 400 nm to 1000 nm. Using
optimum conditions for Si3N4 ARC layer they have calculated almost 1% enhance-
ment in the efficiency of a P-N junction solar cell. Geometry and arrangement of sub-
wavelength ARC structures are also investigated by Hishikawa et al. [95] using FEM
modelling and showed randomly arranged triangular pyramids are more effective
as ARC.
Similarly, other frequency-based optical models, such as RCWA and TMM, has been
used to investigate a wide range of ARC materials. A variety of geometries including
moth-eye like conical arrays, inverted moth-eye structures, hexagonal closed-packed,
tapered and truncated cones, nanopillars, and core/shell structures have been evaluated
by using RCWA methods [96–99]. TMM method is used for layered thin films and
employs continuous boundary conditions for reflectance estimation. Combination of
TMM and EMT was used by Forberich et al. [100] to calculate optical absorptions in solar
cells with ARC based on moth-eye structure. The reflectance profile of the ARC layer was
evaluated by EMT, while TMM was used to simulate the optical absorptions of the solar
cells. A seven-layer ARC for solar cells with 1% – 6% reduction in reflectance was
reported by Kuo et al. [101], where TMM was used to model the reflectance of the
layered ARC.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 133

Optical filters
Optical filters are devices which allow transmission of selective wavelengths based on
their design. They are commonly used in various optical and opto-electronic devices.
FDTD simulation was used by Hosseini et al. [102] for the design of a low pass optical
filter based on a plasmonic nanostrip waveguide capable of proving sub-wavelength
confinement. The simulated filter showed flat low pass filter with 3 dB attenuation at the
optical communication wavelength (λ = 1.55 µm). Another FDTD simulation was
reported by Lee et al. [103] for simulating transmission characteristics of transparent
conducting electrodes composed of silver or aluminum grids with widths from 25 nm
to 5 μm.
A high performance bandpass filter using single-layer resonant Si gratings was
reported recently, where filter transmittance was simulated by using FEM method
[104]. A narrowband tunable antireflection optical filter was reported by Bibbo et al.
[105] where the filter is composed of three layers with a metasurface consisting plas-
monic nanoparticles. FEM simulation was used for the parameters of the materials used
in the filter design and to estimate the refractive index of the layers. The simulation also
showed that the resonance peak of the plasmonic layer can be tuned by controlling the
refractive index of the dielectric layer. A LiNbO3 layer was considered as the dielectric
layer in this case. In another work, FEM was used for modelling diffraction grating based
optical filters for fluorescence detection of biomolecules. Authors have evaluated three
filter concepts using the FEM modelling [106], namely placement of the diffraction
grating, embedded diffraction grating in a low refractive index medium and combina-
tion with a flat absorbing filter. The first one showed minimal improvements in terms of
filter efficiency, while the other to produces high efficiencies (in the order of 105).

Solar cells
Optical simulations for solar cell design are becoming very common these days. Such
simulations help researchers to save time and predict the characteristics of the device
ahead of the actual experiments. To improve solar cell performance one of the common
approaches is to employ antireflection coatings on solar cell surface to increase optical
transmission. Optical simulations related to antireflection coatings have been discussed
earlier. Another approach to improve solar cell efficiency is by light trapping within the
cell to maximize the optical absorption resulting in more photocurrent. Several strate-
gies have been reported for light trapping purpose that can be very useful for solar cell
applications [107]. Light trapping by periodically textured surface with realistic interface
morphologies for amorphous silicon solar cell was investigated by Jovanov et al. [108].
A similar micro-textured light trapping layer was reported for thin film organic solar cells
[109]. In both the cases FEM simulation was carried out to evaluate the light trapping
behavior of such textured layers and their impact on the solar cell output. More than
60% of less optical reflection due to the light trapping was predicted from the simula-
tion resulting in almost 20% improvement in the short-circuit current of the cell.
A nanostructured grating was proposed as a light trapping method for solar cells by
Das et al. [110]. Using FDTD methods authors have calculated about 2% reduction in the
reflection when a triangular shaped nanostructured grating with height of about 300 nm
134 T. BORA ET AL.

and the period of about 830 nm was used. This value of reflectance is almost 28% lower
than the flat type substrates.
RCWA modelling have been also widely used for optical simulation of solar cells [111–
113]. RCWA modelling provides high accuracy, while FEM and FDTD are time and CPU
consuming approaches with advantages of multiphysics modelling. Apart from light
trapping, several simulations are done on the structure of the solar cell materials aiming
for high efficiency. Si nanowire (NW) based solar cells have been extensively investigated
by Foldyna et al. [114,115] using RCWA modelling approach. With increasing NW length
from 1 to 10 μm, they have demonstrated improvement in short circuit current density
from 25 mA/cm2 to 34 mA/cm2. The optical absorption profile of these Si NW was
investigated by Lin et al. [116] by using TMM method. From the calculations, they have
proposed an optimum dimension for the Si NWs with diameter 540 nm and periodicity
600 nm. Using these optimum conditions almost 70% enhancement in the efficiency of
the Si NW solar cell was finally predicted.

‘Post scriptum mechanicus’ and looking forward


We have focused heavily on optical aspects in this paper but must not neglect other
aspects of physical properties – those areas arising from civil engineering, mechanics
and now, nanomechanics. The finite element treatment, as allude to earlier primarily to
solve engineering problems. It originated in the 1940’s by A. Hrennikoff and R. Courant
[117,118]. By the early 1960’s, Feng Kang of China and western researchers developed
a numerical method to solve partial differential equations called the finite difference
method based upon the variational principle – better known as the finite element
method (FEM). In the 1980’s Feng Kang integrated dynamical systems like the
Hamiltonian and wave systems into analytical techniques. The FEM is considered to be
a landmark of computational mathematics. This segue is provided to link

Nanomechanics
Average property determination requires spatial averaging. Consequently, regarding the
non-local, where the stress at a point cannot be determined alone from the strain at that
point, constitutive relations emerge [119]. The resulting stochastic differential equations
for equilibrium demands advanced formulations [120]. The spatially discretized finite
and boundary element models necessitate stochastic shape functions and stochastic
Green’s functions, respectively [121,122]. Symbolic methods become indispensable for
carrying out operations of lengthy algebraic expressions [123]. Nanomechanics must
utilize the most accurate computational schemes. Conventional commercial packages
provide a number of finite element solvers that fail to address the effects of Poisson’s
ratio and especially incompressibility [124,125]. The response computation for stochastic
nanomechanics formulations calls for computation intensive Monte Carlo simulations.
Therein considerable economy can be achieved by ordering the Monte Carlo samples
and carrying out an incremental strategy in the sample space [126].
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 135

Mechanics modeling with FEM


Finite element analysis has been accomplished recently in describing the performance of
a cantilever mechanical energy harvestor based on nanoscale layer of zinc oxide (ZnO)
piezoelectric nanomaterial [127]. The purpose of the simulation was to show the feasibility
of applying an energy harvester that is able to detect low frequency vibrations under
ambient conditions (for example, flowing fluid within a pipe). The link of the device was to
a wireless sensor node system. Application of COMSOL Multiphysics, for example, was able
to simulate resonant frequency modes at selected frequencies. The images below depict
input parameters and simulations at various frequencies (Figure 12).
Simulation of the cantilever at three eigen-frequency modes are shown in Figure 13.
The eigen-frequencies obtained were 575.5, 1810.6 and 3957.3 Hz. Since the energy
harvester is based on the lowest frequency, that frequency was considered to be the
fundamental frequency. The maximum voltage output obtained (by simulation) was
0.32 V at 575.5 Hz input (Figure 14). This is due to the impact of a piezoelectric zinc
oxide layer, considered to be deposited at the microcantilever surface, which is sub-
jected to large deflection at the resonance frequency.

Conclusions
We have presented a brief overview of progress in nanomaterial design theory and
simulation. Coupled classical electromagnetic theory with finite element methods have
shown facilitative simulation capability in designing anti-reflection coatings – especially
with regard to modeling of coatings that consist of layers made of nanoparticles of
complex shape. With the advent of ever more powerful open source or commercially
available programming packages and new and more powerful computational methods,
such as those of Materials Studio™ [128], nanomaterial modeling will gain in popularity
and be available to more researchers. There are ingenious ways to create hybrid
methods that link Schrödinger’s equation with Maxwell’s equations and continuing on
to the macroscopic level. An all-encompassing unified method is still off into the future
however enough technology is available to accurately model nanomaterial physical
properties, especially those of optical properties. Ironically, it will be, we predict, by

Figure 12. Top Left: Basic engineering drawing to provide dimensional input parameters of
cantilever. Top right: Mesh model developed in COMSOL Multiphysics FEM program.
136 T. BORA ET AL.

Figure 13. Top left: First-mode resonant frequency at 575.5 Hz. Top right: Second-mode resonant
frequency at 1810.6 Hz. Bottom left: Third-mode resonant frequency at 3957.3 Hz.

Figure 14. Voltage vs. frequency of the potential energy harvester as simulated by COMSOL
Multiphysics program. The maximum voltage was derived from the resonant frequency at 575.5 Hz.

quantum computing that quantum mechanics is utilized to completely characterize


material behavior, at the least for mesoscopic systems.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 137

Disclosure statement
No potential conflict of interest was reported by the authors.

References
[1] D.M. Solis, J.M. Taboada, L. Landesa, J.L. Rodriguez, and F. Obelleiro, Squeezing Maxwell’s
equations into the nanoscale, Prog. Electromagn. Res. 154 (2015), pp. 35050. doi:10.2528/
PIER15110103.
[2] Y.P. Chen, W.E.I. Sha, L.J. Jiang, Y.M. Wu, and W.C. Chew, A unified Hamiltonian solution to
Maxwell-Schrödinger equations for modeling electromagnetic field-particle interaction,
Comput. Phys. Commun. 215 (2017), pp. 63–70. doi:10.1016/j.cpc.2017.02.006.
[3] P. Monk, Finite Element Methods for Maxwell’s Equations, Oxford University Press, New York,
2003.
[4] N. Engheta, 150 years of Maxwell’s equations, Sci. 349 (6244) (2015), pp. 136–137.
doi:10.1126/science.aaa7224.
[5] J. Yang and W. Sui; Solving Maxwell-Schrödinger equations for analyses of nano-scale
devices, 2007 European Microwave Devices; Munich, Germany; IEEE Explore, doi:10.1109/
EUMC.2007.4405149 (2007)
[6] W.S. Mohammed and G.L. Hornyak, Optical behavior of periodic media: A classical electro-
magnetic (mesoscopic) approach, in Chapter 9, Handbook of Nanoscience, Engineering and
Technology, 3rd ed., W.A. Goddard III, D. Brenner, S.E. Lyshevski, and G.J. Iafrate, eds., 193
-238, CRC Press, Boca Raton, FLA USA, 2012.
[7] G.L. Hornyak, C.J. Patrissi, and C.R. Martin, Fabrication, characterization and optical properties
of gold nanoparticle/porous alumina composites: The nonscattering Maxwell-Garnett limit,
J. Phys. Chem. B 101 (9) (1997), pp. 1548–1555. doi:10.1021/jp962685o.
[8] S. Link, M.B. Mohamed, and M.A. El-Sayed, Simulation of the optical absorption spectra of gold
nanorods as a function of their aspect ratio and the effect of the medium dielectric constant,
J. Phys. Chem. B 103 (16) (1999), pp. 3073–3077. doi:10.1021/jp990183f.
[9] J.-L. Fattebert, R.D. Hornung, and A.M. Wissink, Finite element approach for density functional
theory calculations on locally refined meshes, Lawrence Livermore National Laboratory Report
UCRL-TR-228360, February 26, 2007
[10] G.M. Odegard, T.S. Gates, L.M. Nicholson, and K.E. Wise, Equivalent continuum modeling of
nanostructured materials, Composite Sci. Technol. 62 (2002), pp. 1869–1880. doi:10.1016/
S0266-3538(02)00113-6.
[11] J.A. Smirnova, L.V. Zhigilei, and B.J. Garrison, Combined molecular dynamics and finite
element method applied to laser induced pressure wave propagation, Comput. Phys.
Commun. 118 (1999), pp. 11–16. doi:10.1016/S0010-4655(98)00175-1.
[12] B. Liu, Y. Huang, H. Jiang, S. Qu, and K.C. Hwang, The atomic scale finite element method,
Comput. Methods Appl. Mech. Eng. 191 (2004), pp. 1849–1864. doi:10.1016/j.
cma.2003.12.037.
[13] T.J.R. Hughes, R.L. Taylor, and W. Kanok-Nukulchai, A simple and efficient finite element for
plate bending, Int. J. Numer. Methods. Eng. 11 (10) (1977), pp. 1529–1543. doi:10.1002/
nme.1620111005.
[14] M.H. Ulz, Coupling the finite element method and molecular dynamics in the framework of the
heterogeneous multiscale method for quasi-static isothermal problems, J. Mech. Phys. Solids.
24 (2015), pp. 1–18. doi:10.1016/j.jmps.2014.10.002.
[15] S. Burger, L. Zschiedrich, J. Pomplun, M. Blome, and F. Schmidt, Advanced finite-element
methods for design and analysis of nanooptical structures: applications, SPIE 8642, pp. 864205
(2013).
[16] P.B. Johnson and R.W. Christy, Optical constants of the noble metals, Phys. Rev. B 6 (1972), pp.
4370. doi:10.1103/PhysRevB.6.4370.
138 T. BORA ET AL.

[17] L.B. Scaffardi and J.O. Tocho, Size dependence of refractive index of gold nanoparticles,
Nanotechnology 17 (5) (2006), pp. 1309–1315. doi:10.1088/0957-4484/17/5/024.
[18] C. Pecharroman, E.D. Gaspera, A. Martucci, R. Escobar-Galindod, and P. Mulvaney,
Determination of optical constants of gold nanoparticles from thin-film spectra, J. Phys.
Chem. C 119 (17) (2015), pp. 9450–9459. doi:10.1021/jp512611m.
[19] K.K. Chattopaghyay and N.S. Das, Size-dependent optical properties of nanoparticles ana-
lyzed by spectroscopic ellipsometry; Chapter 4, Handbook of Nanoparticles, Springer
International Publishing, Switzerland, 2016.
[20] C.A. Foss, G.L. Hornyak, J.A. Stockert, and C.R. Martin, Template-synthesized nanoscopic gold
particles: optical spectra and the effects of particle size and shape, J. Phys. Chem. 98 (11)
(1994), pp. 2963–2971. doi:10.1021/j100062a037.
[21] M. Wang and N. Pan, Predictions of effective physical properties of complex multiphase
materials, Mater. Sci. Eng. R 63 (1) (2008), pp. 1–30. doi:10.1016/j.mser.2008.07.001.
[22] V.A. Markel, Introduction to the Maxwell Garnett approximation, J. Opt. Soc. Am. 33 (7) (2016),
pp. 1244–1256. doi:10.1364/JOSAA.33.001244.
[23] R. Resta and D. Vanderbilt, Theory of polarization: A modern approach; Physics of
Ferroelectrics: A Modern Perspective, K. M. Rabe, C. H. Ahn, J-M Triscone, eds., Springer,
2007, pp. 31–68.
[24] U. Kreibig and M. Vollmer, Optical Properties of Metal Clusters, Springer-Verlag, Berlin, 1995.
[25] G.L. Hornyak, Characterization and optical theory of nanometal/porous alumina composite
membranes, Dissertation, Department of Chemistry, Colorado State University, C.R. Martin
(Advisor), 1997.
[26] K. Han and C.-H. Chang, Numerical modeling of sub-wavelength anti-reflectivestructures for
solar module applications, Nanomaterials 4 (2014), pp. 87–128. doi:10.3390/nano4010087.
[27] P. Lalanne and D. Lemercier-Lalanne, On the effective medium theory of subwavelength
periodic structures, J. Mod. Opt. 43 (10) (1996), pp. 2063–2085. doi:10.1080/
09500349608232871.
[28] P. Spinelli, M.A. Verschuuren, and A. Polman, Broadband omnidirectional antireflection coat-
ing based on subwavelength surface Mie resonators, Nat. Commun. 3 Article number: 692,
(2012), pp. 1–5. doi:10.1038/ncomms1691.
[29] A. Gombert, K. Rose, A. Heinzel, W. Horbelt, C. Zanke, B. Bläsi, and V. Wittwer, Antireflective
submicrometer surface-relief gratings for solar applications, Sol. Energy Mater. Sol. Cells 54
(1–4) (1998), pp. 333–342. doi:10.1016/S0927-0248(98)00084-1.
[30] J. Gjessing, A.S. Sudbø, and E.S. Marstein, Comparison of periodic light-trapping structures in
thin crystalline silicon solar cells, J. Appl. Phys. 110 (033104) (2011). doi:10.1063/1.3611425
[31] O. Isabella, H. Sai, M. Kondo, and M. Zeman, Full-wave optoelectrical modeling of optimized
flattened light-scattering substrate for high efficiency thin-film silicon solar cells, Prog.
Photovoltaics: Res. Appl. 22 (2014), pp. 671–689. doi:10.1002/pip.v22.6.
[32] N. Tucher, J. Eisenlohr, P. Kiefel, O. Höhn, H. Hauser, M. Peters, C. Müller, J.-C. Goldschmidt,
and B. Bläsi, 3D optical simulation formalism OPTOS for textured silicon solar cells, Opt.
Express 23 (2015), pp. A1720–A1734. doi:10.1364/OE.23.0A1720.
[33] J. Eisenlohr, N. Tucher, O. Höhn, H. Hauser, M. Peters, P. Kiefel, J.-C. Goldschmidt, and B. Bläsi,
Matrix formalism for light propagation and absorption in thick textured optical sheets, Opt.
Express 23 (2015), pp. A502–A518. doi:10.1364/OE.23.00A502.
[34] N. Tucher, B. Muller, P. Jakob, J. Eisenlohr, O. Hohn, H. Hauser, J.C. Goldschmidt, M. Hermle,
and B. Blasi, Optical performance of the honeycomb texture – A cell and module level analysis
using the OPTOS formalism, Sol. Energy Mater. Sol. Cells 173 (2017), pp. 66–71. doi:10.1016/j.
solmat.2017.06.004.
[35] Z.-Y. Li and -L.-L. Lin, Photonic band structures solved by a plane-wave-based transfer matrix,
Phys. Rev. E 67 (4) (2003), pp. 046607. doi:10.1103/PhysRevE.67.046607.
[36] W.H. Southwell, Gradient-index antireflection coatings, Opt. Lett. 8 (11) (1983), pp. 584–586.
doi:10.1364/OL.8.000584.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 139

[37] W.H. Southwell, Pyramid-array surface-relief structures producing antireflection index


matching on optical surfaces, J. Opt. Soc. America 8 (1991), pp. 549–553. doi:10.1364/
JOSAA.8.000549.
[38] J.-Q. Xi, M.F. Schubert, J.K. Kim, E.F. Schubert, M. Chen, S.-Y. Lin, W. Liu, and J.A. Smart,
Optical thin-film materials with low refractive index for broadband elimination of Fresnel
reflection, Nat. Photonics 1 (2007), pp. 176–179. doi:10.1038/nphoton.2007.26.
[39] J. Olsen and P. Jorgensen, Linear and nonlinear response functions for an exact state and for
an MCSCF state, J. Chem. Phys. 82 (1985), pp. 3235–3264. doi:10.1063/1.448223.
[40] P. Norman, D.M. Bishop, H.J.A. Jensen, and J. Oddershede, Quasi energy formulation of
damped response theory, J. Chem. Phys. 123 (2005), pp. 194103. doi:10.1063/1.2107627.
[41] A. Jiemchooroj and P. Norman, Electronic circular dichroism spectra from the complex
polarization propagator, J. Chem. Phys. 126 (2007), pp. 134102. doi:10.1063/1.2716660.
[42] M. Zhu, C.M. Aikens, F.J. Hollander, G.C. Schatz, and R. Jin, Correlating the crystal structure of
a thiol-protected Au25 cluster and optical properties, J. Am. Chem. Soc. 130 (18) (2008), pp.
5883. doi:10.1021/ja801173r.
[43] Y. Yanagimoto, Y. Negishi, H. Fujihara, and T. Tsukuda, Chiroptical activity of BINAP-stabilized
undecagold clusters, J. Phys. Chem. B 110 (24) (2006), pp. 11611–11614. doi:10.1021/
jp061670f.
[44] C. Gautier and T. Burgi, Chiral gold nanoparticles, Chem. Phys. Phys. Chem. 10 (3) (2009), pp.
483–492. doi:10.1002/cphc.200800709.
[45] O. Toikkanen, V. Ruiz, G. Ronholm, N. Kalkkinen, P. Liljeroth, and B.M. Quinn, Synthesis and
stability of monolayer-protected Au38 clusters, J. Am. Chem. Soc. 130 (33) (2008), pp.
11049–11055. doi:10.1021/ja802317t.
[46] Z. Rinkevicius, J. Autschbach, A. Baev, M. Swihart, H. Ågren, and P.N. Prasad, Novel pathways
for enhancing nonlinearity of organics utilizing metal clusters, J. Phys. Chem. 114 (2010), pp.
7590–7594. doi:10.1021/jp102438m.
[47] L.L. Jensen and L. Jensen, Atomistic electrodynamics model for optical properties of silver
nanoclusters, J. Phys. Chem. C 113 (2009), pp. 15182–15190. doi:10.1021/jp904956f.
[48] Z. Rinkevicius, X. Li, J.A. Sandberg, K. Mikkelsen, and H. Agren, A hybrid density functional
theory/molecular mechanics approach for linear response properties in heterogeneous
environments, J. Chem. Theory Comput. 10 (2014), pp. 989. doi:10.1021/ct500394t.
[49] M. Barbry, P. Koval, F. Marchesin, R. Esteban, A.G. Borisov, J. Aizpurua, and D. Sánchez-Portal,
Atomistic near-field nanoplasmonics: reaching atomic-scale resolution in nanooptics, Nano
Lett. 15 (2015), pp. 3410–3419. doi:10.1021/acs.nanolett.5b00759.
[50] A. Baev, E. Furlani, P.N. Prasad, A. Grigorenko, and N.W. Roberts, Laser nano-trapping and
manipulation of nano-scale objects using sub-wavelength apertured plasmonic media, J. Appl.
Phys. 103 (2008), pp. 084316. doi:10.1063/1.2912493.
[51] W.-C. Law, K.-T. Yong, A. Baev, R. Hu, and P.N. Prasad, Nanoparticle enhanced surface plasmon
resonance biosensing: application of gold nanorods, Opt. Express 17 (2009), pp. 19041.
doi:10.1364/OE.17.019041.
[52] S. Shukla, K.T. Kim, A. Baev, Y.-K. Yoon, N.M. Litchinitser, and P.N. Prasad, Fabrication and
characterization of gold-polymer nanocomposite plasmonic nanoarrays in porous alumina
template, ACS Nano 4 (2010), pp. 2249. doi:10.1021/nn9018398.
[53] G.S. He, J. Zhu, K.-T. Yong, A. Baev, H.-X. Cai, R. Hu, Y. Cui, and P.N. Prasad, Scattering and
absorption cross-section’s spectral measurements of gold nanorods in water, J. Phys. Chem. C
114 (2010), pp. 2853. doi:10.1021/jp907811g.
[54] W.-C. Law, K.-T. Yong, A. Baev, and P.N. Prasad, Sensitivity improved surface plasmon reso-
nance biosensor for cancer biomarker detection based on plasmonic enhancement, ACS Nano
5 (2011), pp. 4858. doi:10.1021/nn202666w.
[55] G.S. He, W.-C. Law, A. Baev, S. Liu, M.T. Swihart, and P.N. Prasad, Nonlinear optical absorption
and stimulated Mie scattering in metallic nanoparticle suspensions, J. Chem. Phys. 138 (2013),
pp. 024202. doi:10.1063/1.4773340.
[56] F. Alali, Y.H. Kim, A. Baev, and E.P. Furlani, Plasmon-enhanced metasurfaces for controlling
optical polarization, ACS Photonics 1 (6) (2014), pp. 507–515. doi:10.1021/ph5000192.
140 T. BORA ET AL.

[57] M. Maldonado, H.T.M.C.M. Baltar, A.S.L. Gomes, R. Vaia, K. Park, J. Che, M. Hsiao, C.B. de
Araújo, A. Baev, and P.N. Prasad, Coupled-Plasmon induced optical nonlinearities in aniso-
tropic arrays of gold nanorod clusters supported in a polymeric film, J. Appl. Phys. 121 (2017),
pp. 143103. doi:10.1063/1.4980027.
[58] I.V.A.K. Reddy, A. Baev, E.P. Furlani, P.N. Prasad, and J.W. Haus, Interaction of structured light
with a chiral plasmonic metasurface: Giant enhancement of chiro-optic response, ACS
Photonics 5 (2018), pp. 734–740. doi:10.1021/acsphotonics.7b01321.
[59] www.comsol.com
[60] J.D. Jackson, The Classical Electromagnetic Field, Wiley, New York, 1999.
[61] E.L. Linder, Effects of electron pressure on plasma electron oscillations, Phys. Rev. 49 (1936),
pp. 753. doi:10.1103/PhysRev.49.753.
[62] F.J. Garcia de Abajo, Nonlocal effects in the plasmons of strongly interacting nanoparticles,
dimers, and waveguides, J. Phys. Chem. B 112 (2008), pp. 17983.
[63] J.M. McMahon, S.K. Gray, and G.C. Schatz, Optical properties of nanowire dimers with
a spatially nonlocal dielectric function, Nano Lett. 10 (2010), pp. 3473. doi:10.1021/
nl101606j.
[64] C. Ciraci, R.T. Hill, J.J. Mock, Y. Urzhumov, A.I. Fernandez-Dominguez, S.A. Maier, J.B. Pendry,
A. Chilkoti, and D.R. Smith, Probing the ultimate limits of plasmonic enhancement, Sci. 337
(2012), pp. 1072. doi:10.1126/science.1224823.
[65] N.A. Mortensen, S. Raza, M. Wubs, T. Søndergaardand, and S.I. Bozhevolnyi, A generalized
non-local optical response theory for plasmonic nanostructures, Nat. Commun. 5 (2014), pp.
3809. doi:10.1038/ncomms5972.
[66] Y. Luo, A.I. Fernandez-Dominguez, A. Wiener, S.A. Maier, J.B. Pendry, and S. Plasmons,
Nonlocality: a simple model, Phys. Rev. Lett. 111 (2013), pp. 93901.
[67] N. Bloembergen, R.K. Chang, S.S. Jha, and C.H. Lee, Optical harmonic generation in reflection
from media with inversion symmetry, Phys. Rev. 174 (3) (1968), pp. 813. doi:10.1103/
PhysRev.174.813.
[68] M. Corvi and W.L. Schaich, Hydrodynamics model calculation of second harmonic generation
at a metal surface, Phys. Rev. B 33 (1986), pp. 3688. doi:10.1103/PhysRevB.33.3688.
[69] M. Scalora, M.A. Vincenti, D. de Ceglia, V. Roppo, M. Centini, N. Akozbek, and M.J. Bloemer,
Second- and third-harmonic generation in metal-based structures, Phys. Rev. 82 (2010), pp.
043828. doi:10.1103/PhysRevA.82.043828.
[70] M. Scalora, M. Vincenti, D. de Ceglia, N. Akozbek, V. Roppo, M. Bloemerand, and J.W. Haus,
Dynamical model of harmonic generation in centrosymmetric semiconductors at visible and UV
wavelengths, Phys. Rev. 85 (2012), pp. 053809. doi:10.1103/PhysRevA.85.053809.
[71] J.W. Haus, L. Li, N. Katte, C. Deng, M. Scalora, D. de Ceglia, and M.A. Vincenti, Nanowire
metal-insulator-metal plasmonic devices, ICPS 2013: International conference on photonics
solutions, in Proceedings Of SPIE Vol. 8883, eds., P. Buranasiri and S. Sumriddetchkajorn,
2013, pp. 888303, Bellingham: SPIE.
[72] J.W. Haus, D. de Ceglia, M.A. Vincenti, and M. Scalora, Quantum conductivity for
metal-insulator-metal nanostructures, J. Opt. Soc. America B 31 (2014), pp. 259.
doi:10.1364/JOSAB.31.000259.
[73] J.W. Haus, D. de Ceglia, M.A. Vincenti, and M. Scalora, Nonlinear quantum tunneling effects in
nano-plasmonic environments, J. Opt. Soc. Am. B 31 (2014), pp. A13. doi:10.1364/
JOSAB.31.000A13.
[74] M. Scalora, J.W. Haus, D. de Ceglia, and M.A. Vincenti, Nonlocal and quantum tunneling
contributions to harmonic generation in nanostructures: electron cloud screening effects, Phys.
Rev. 90 (2014), pp. 013831. doi:10.1103/PhysRevA.90.013831.
[75] J. Zuloaga, E. Prodan, and P. Nordlander, Quantum description of the plasmon resonances of
a nanoparticle dimer, Nano Lett. 9 (2009), pp. 887. doi:10.1021/nl803811g.
[76] J. Zuloaga, E. Prodan, and P. Nordlander, Quantum plasmonics: optical properties and tun-
ability of metallic nanorods, ACS Nano 4 (2010), pp. 5269. doi:10.1021/nn101589n.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 141

[77] D.C. Marinica, A.K. Kazansky, P. Nordlander, J. Aizpurua, and A.G. Borisov, Quantum plasmo-
nics: Nonlinear effects in the field enhancement of a plasmonic nanoparticle dimer, Nano Lett.
12 (2012), pp. 1333. doi:10.1021/nl300269c.
[78] R. Esteban, A.G. Borisov, P. Nordlander, and J. Aizpurua, Bridging quantum and classical
plasmonics with a quantum-corrected model, Nat. Commun. 3 (2012), pp. 825. doi:10.1038/
ncomms1806.
[79] T.V. Teperik, P. Nordlander, J. Aizpurua, and A.G. Borisov, Quantum effects and nonlocality in
strongly coupled plasmonic nanowire dimers, Opt. Express 21 (2013), pp. 27306–27325.
[80] A. Baev, P.N. Prasad, H. Ågren, M. Samoć, and M. Wegener, Metaphotonics: an emerging field
with opportunities and challenges, Phys. Rep. 594 (2015), pp. 1–60. doi:10.1016/j.
physrep.2015.07.002.
[81] G.V. Naik, V.M. Shalaev, and A. Boltasseva, Alternative plasmonic materials: beyond gold and
silver, Adv. Mater. 25 (2013), pp. 3264–3294. doi:10.1002/adma.201205076.
[82] D.G. Stavenga, S. Foletti, G. Palasantzas, and K. Arikawa, Light on the moth-eye corneal
nipple array of butterflies, Proc. R. Soc. B, 273 (2006), pp. 661–667
[83] H.K. Raut, V.A. Ganesh, A.S. Nair, and S. Ramakrishna, Anti-reflective coatings: a critical,
in-depth review, Energy Environ. Sci. 4 (2011), pp. 3779–3804. doi:10.1039/c1ee01297e.
[84] Z.W. Han, Z. Wang, X.M. Feng, B. Li, Z.Z. Mu, J.Q. Zhang, S.C. Niu, and L.Q. Ren, Antireflective
surface inspired from biology: a review, Biosurf. Biotribol. 2 (2016), pp. 137–150. doi:10.1016/j.
bsbt.2016.11.002.
[85] M. Nevière and E. Popov, Light Propagation in Periodic Media: Differential Theory and Design,
CRC Press, New York, NY, USA, 2003.
[86] B.L. Good, S. Simmons, and M. Mirotznik, General optimization of tapered anti-reflective
coatings, Opt. Express 24 (2016), pp. 16618–16629. doi:10.1364/OE.24.029465.
[87] R. Bouffaron, L. Escoubas, J.J. Simon, P. Torchio, F. Flory, G. Berginc, and P. Masclet, Enhanced
antireflecting properties of micro-structured top-flat pyramids, Opt. Express 16 (2008), pp.
19304–19309. doi:10.1364/OE.16.019304.
[88] A.C. Lesina, G. Paternoster, F. Mattedi, L. Ferrario, P. Berini, L. Ramunno, A. Paris, A. Vaccari,
and L. Calliari, Modeling and characterization of antireflection coatings with embedded silver
nanoparticles for silicon solar cells, Plasmonics 10 (2015), pp. 1525–1536. doi:10.1007/s11468-
015-9957-7.
[89] N.N. Feng, G.R. Zhou, and W.P. Huang, Space mapping technique for design optimization of
antireflection coatings in photonic devices, J. Lightwave Technol. 21 (2003), pp. 281–285.
doi:10.1109/JLT.2003.808641.
[90] Z.F. Li, E. Ozbay, H.B. Chen, J.J. Chen, F.H. Yang, and H.Z. Zheng, Resonant cavity based
compact efficient antireflection structures for photonic crystals, J. Phys. D 40 (2007), pp.
5873–5877. doi:10.1088/0022-3727/40/19/012.
[91] G.R. Zhou, X. Li, and N.N. Feng, Design of deeply etched antireflective waveguide terminators,
IEEE J. Quantum Electron. 39 (2003), pp. 384–391. doi:10.1109/JQE.2002.807185.
[92] I. Andonegui and A.J. Garcia-Adeva, The finite element method applied to the study of
two-dimensional photonic crystals and resonant cavities, Opt. Express 21 (2013), pp.
4072–4092. doi:10.1364/OE.21.017108.
[93] G. Demésy, F. Zolla, A. Nicolet, M. Commandré, and C. Fossati, The finite element method
as applied to the diffraction by an anisotropic grating, Opt. Express 15 (2007), pp.
18089–18102.
[94] H.M. Lee, K.C. Sahoo, Y. Li, J.C. Wu, and E.Y. Chang, Finite element analysis of antireflective
silicon nitride sub-wavelength structures for solar cell applications, Thin Solid Films 518 (2010),
pp. 7204–7208. doi:10.1016/j.tsf.2010.04.078.
[95] Y. Hishikawa, T. Kinoshita, M. Shima, M. Tanaka, S. Kiyama, S. Tsuda, and S. Nakano; Optical
confinement and optical loss in high-efficiency a-Si solar cells; Proceedings of the Conference
Record of the Twenty-Sixth IEEE, Photovoltaic Specialists Conference; Anaheim, CA, USA;
29 September–3 October 1997; 615–618 (1997)
[96] S.A. Boden and D.M. Bagnall, Tunable reflection minima of nanostructured antireflective
surfaces, Appl. Phys. Lett. 93 (133108) (2008), pp. 1–133108. 3. doi:10.1063/1.2993231.
142 T. BORA ET AL.

[97] M.S. Mirotznik, B.L. Good, P. Ransom, D. Wikner, and J.N. Mait, Broadband antireflective
properties of inverse motheye surfaces, IEEE Trans. Antennas Propag. 58 (2010), pp.
2969–2980. doi:10.1109/TAP.2010.2052575.
[98] J.W. Leem, Y.M. Song, and J.S. Yu, Six-fold hexagonal symmetric nanostructures with various
periodic shapes on glass substrates for efficient antireflection and hydrophobic properties,
Nanotechnology 22 (2011), pp. 485304. doi:10.1088/0957-4484/22/48/485304.
[99] J.W. Leem, Y. Yeh, and J.S. Yu, Enhanced transmittance and hydrophilicity of nanostructured
glass substrates with antireflective properties using disordered gold nanopatterns, Opt. Express
20 (2012), pp. 4056–4066. doi:10.1364/OE.20.004056.
[100] K. Forberich, G. Dennler, M.C. Scharber, K. Hingerl, T. Fromherz, and C.J. Brabec, Performance
improvement of organic solar cells with moth eye anti-reflection coating, Thin Solid Films 516
(2008), pp. 7167–7170. doi:10.1016/j.tsf.2007.12.088.
[101] M.L. Kuo, D.J. Poxson, Y.S. Kim, F.W. Mont, L.K. Kim, E.F. Schuhert, and S.Y. Lin:, Realization of
a near-perfect antireflection coating for silicon solar energy utilization, Opt. Lett. 33 (2008), pp.
2527–2529. doi:10.1364/OL.33.002527.
[102] A. Hosseini, H. Nejati, and Y. Massoud, Design of a maximally flat optical low pass filter using
plasmonic nanostrip waveguides, Opt. Express 15 (2007), pp. 15280–15286.
[103] K. Lee, S.H. Song, and J. Ahn, FDTD simulation of transmittance characteristics of
one-dimensional conducting electrodes, Opt. Express 22 (2014), pp. 6269–6275. doi:10.1364/
OE.22.006269.
[104] M. Niraula, J.W. Yoon, and R. Magnusson, Single-periodic-film optical bandpass filter, Opt. Lett.
40 (2015), pp. 5062–5065. doi:10.1364/OL.40.005062.
[105] L. Bibbò, K. Khan, Q. Liu, M. Lin, Q. Wang, and Z. Ouyang, Tunable narrowband antireflection
optical filter with a metasurface, Photonics Res. 5 (2017), pp. 297784. doi:10.1364/
PRJ.5.000500.
[106] M. Kovačič, J. Krč, B. Lipovšek, and M. Topič, Modelling of diffraction grating based optical
filters for fluorescence detection of biomolecules, Biomed. Opt. Express 5 (2014), pp.
2285–2300. doi:10.1364/BOE.5.002285.
[107] Z. Tang, W. Tress, and O. Inganäs, Light trapping in thin film organic solar cells, Mater. Today
17 (2014), pp. 389–396. doi:10.1016/j.mattod.2014.05.008.
[108] V. Jovanov, U. Palanchoke, P. Magnus, H. Stiebig, J. Hüpkes, P. Sichanugrist, M. Konagai,
S. Wiesendanger, C. Rockstuhl, and D. Knipp, Light trapping in periodically textured amor-
phous silicon thin film solar cells using realistic interface morphologies, Opt. Express 21 (2013),
pp. A595–606. doi:10.1364/OE.21.00A595.
[109] B. Lipovšek, A. Čampa, F. Guo, C.J. Brabec, K. Forberich, J. Krč, and M. Topič, Detailed optical
modelling and light-management of thin-film organic solar cells with consideration of
small-area effects, Opt. Express 25 (2017), pp. A176–A190. doi:10.1364/OE.25.00A176.
[110] N. Das and S. Islam, Design and analysis of nano-structured gratings for conversion efficiency
improvement in GaAs solar cells, Energies 9 (2016), pp. 690. doi:10.3390/en9090690.
[111] M. Agrawal, M. Frei, Y. Bhatnagar, T. Repmann, K. Witting, J. Schroeder, and C. Eberspacher;
Comprehensive experimental and numerical optimization of surface morphology of trans-
parent conductive oxide films for tandem thin film photovoltaic cells, Proc. of 35th IEEE
Photovoltaic Specialists Conference (PVSC), Honolulu, HI, USA, 301–304 (2010)
[112] M. Peters, M. Rudiger, B. Bläsi, and W. Platzer, Electro-optical simulation of diffraction in solar
cells, Opt. Express 18 (2010), pp. A584–A593. doi:10.1364/OE.18.025008.
[113] I. Semenikhin, M. Zanuccoli, M. Benzi, V. Vyurkov, E. Sangiorgi, and C. Fiegna, Computational
efficient RCWA method for simulation of thin film solar cells, Opt. Quantum Electron. 44
(2012), pp. 149–154. doi:10.1007/s11082-012-9560-5.
[114] M. Foldyna, L. Yu, B. O’Donnell, and P.R.I. Cabarrocas; Optical absorption in vertical silicon
nanowires for solar cell applications, Proceedings of SPIE 2011, Society of Photo-Optical
Instrumentation Engineers, 8111, SPIE Press, Bellingham (2011)
[115] M. Foldyna, L. Yu, B. O’Donnell, and P.R.I. Cabarrocas, Theoretical short-circuit current density
for different geometries and organizations of silicon nanowires in solar cells, Solid Energy
Materials Solar Cells 10 (117) (2013), pp. 645–651. doi:10.1016/j.solmat.2012.10.014.
INTERNATIONAL JOURNAL OF SMART AND NANO MATERIALS 143

[116] C. Lin and M. Povinelli, Optical absorption enhancement in silicon nanowire arrays with a large
lattice constant for photovoltaic applications, Opt. Express 17 (2009), pp. 19371–19381.
doi:10.1364/OE.17.019371.
[117] A. Hrennikoff, Solution of problems of elasticity by the framework method, J. Appl. Mech. 8
(1941), pp. 169–175.
[118] R. Courant, Variational methods for the solution of problems of equilibrium and vibrations,
Bull. Am. Math. Soc. 49 (1943), pp. 1–23. doi:10.1090/S0002-9904-1943-07818-4.
[119] K. Sobczyk, Stochastic Wave Propagation, latest ed., Elsevier, Amsterdam, 2012.
[120] K. Sobczyk, Stochastic Differential Equations, latest ed., Kluwer Academic Press, Dordrecht,
2001.
[121] G. Dasgupta, Stochastic shape functions and stochastic strain-displacement matrix for
a stochastic finite element stiffness matrix, Acta Mech. 195 (1–4) (2008), pp. 379–395.
doi:10.1007/s00707-007-0569-y.
[122] G. Dasgupta, Green’s function for random media, J. Chin. Inst. Eng. 23 (2000), pp. 377–384.
doi:10.1080/02533839.2000.9670558.
[123] G. Dasgupta, Finite Element Concepts, Springer Verlag, New York, 2018.
[124] G. Dasgupta, Locking-free compressible quadrilateral finite elements: poisson’s ratio-dependent
vector interpolants, Acta. Mech. 225 (1) (2014), pp. 309–330. doi:10.1007/s00707-013-0927-x.
[125] G. Dasgupta, Incompressible and locking-free finite elements from Rayleigh mode vectors:
quadratic polynomial displacement fields, Acta. Mech. 223 (8) (2012), pp. 1645–1656.
doi:10.1007/s00707-012-0654-8.
[126] G. Dasgupta, Approximate dynamic responses in random media, Acta Mech. 3 (1992), pp.
99–114 (Special Volume edited by D. E. Be Sok’s and F. Zieglerc).
[127] K. Sarma, Energy Harvesting from Vibration Source Using Piezo-MEMS/NEMS Cantilever,
Master’s Thesis, Asian Institute of Technology, December 2017.
[128] http://accelrys.com/products/collaborative-science/biovia-materials-studio/.

You might also like