0% found this document useful (0 votes)
137 views21 pages

CIGS - FDTD - Scaps pg-10

The document discusses using computational simulations to model optoelectronic behavior in ultrathin Cu(In,Ga)Se2 solar cells. Precise optical and electrical models were developed to simulate thin and ultrathin CIGS cells. Various light management techniques were studied including texturing and plasmonic nanoparticles. The simulations aimed to enhance light absorption in ultrathin cells and surpass performance of thin film cells.

Uploaded by

u1802119
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
137 views21 pages

CIGS - FDTD - Scaps pg-10

The document discusses using computational simulations to model optoelectronic behavior in ultrathin Cu(In,Ga)Se2 solar cells. Precise optical and electrical models were developed to simulate thin and ultrathin CIGS cells. Various light management techniques were studied including texturing and plasmonic nanoparticles. The simulations aimed to enhance light absorption in ultrathin cells and surpass performance of thin film cells.

Uploaded by

u1802119
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 21

PROCEEDINGS OF SPIE

SPIEDigitalLibrary.org/conference-proceedings-of-spie

Optoelectronic simulations for novel


light management concepts in Cu(In,
Ga)Se2 solar cells

António J. Oliveira, João R. Barbosa, André Violas,


Jennifer Teixeira, Kevin Oliveira, et al.

António J. N. Oliveira, João R. S. Barbosa, André Violas, Jennifer P. Teixeira,


Kevin Oliveira, Tomás Lopes, José M. V. Cunha, Marco A. Curado, Margarida
Monteiro, Célia Rocha, Carlos Vinhais, Paulo A. Fernandes, Pedro M. P.
Salomé, "Optoelectronic simulations for novel light management concepts in
Cu(In,Ga)Se2 solar cells," Proc. SPIE 11681, Physics, Simulation, and
Photonic Engineering of Photovoltaic Devices X, 1168108 (5 March 2021);
doi: 10.1117/12.2577650

Event: SPIE OPTO, 2021, Online Only

Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use


Optoelectronic simulations for novel light management concepts in
Cu(In,Ga)Se2 solar cells
António J. N. Oliveira*a,b, João R. S. Barbosaa, André Violasa, Jennifer P. Teixeiraa, Kevin Oliveiraa,
Tomás Lopesa,c,d,e, José M. V. Cunhaa, b, f, Marco A. Curadoa,g, Margarida Monteiroa, Célia Rochaa,
Carlos Vinhaisa,h, Paulo A. Fernandesa,f,i, Pedro M. P. Saloméa,b
a
INL—International Iberian Nanotechnology Laboratory Avenida Mestre José Veiga, 4715-330
Braga, Portugal; bDepartamento de Física, Universidade de Aveiro, Campus Universitário de
Santiago, 3810-193 Aveiro, Portugal; cDivision IMOMEC (Partner in Solliance) Imec
Wetenschapspark 1, 3590 Diepenbeek, Belgium; dInstitute for Material Research (IMO) Hasselt
University (Partner in Solliance) Agoralaangebouw H, 3590 Diepenbeek, Belgium; eEnergyVille,
Thor Park Poort Genk 8310 & 8320, 3600, Belgium; fi3N Departamento de Física da Universidade
de Aveiro Campus Universitário de Santiago, 3810-193 Aveiro, Portugal; gCFisUC, Department of
Physics University of Coimbra R. Larga, P-3004-516 Coimbra, Portugal; hDepartamento de Física,
Instituto Superior de Engenharia do Porto, Instituto Politécnico do Porto, Porto 4200-072, Portugal;
i
CIETI Departamento de Física Instituto Superior de Engenharia do Porto Instituto Politécnico do
Porto 4200-072 Porto, Portugal
*[email protected]

ABSTRACT
One of the trends making its way through the Photovoltaics (PV) industry, is the search for new application possibilities.
Cu(In,Ga)Se2 (CIGS) thin film solar cells stand out due to their class leading power conversion efficiency of 23.35 %,
flexibility, and low cost. The use of sub-µm ultrathin CIGS solar cells has been gaining prevalence, due to the reduction
in material consumption and the manufacturing time. Precise CIGS finite-difference time-domain (FDTD) and 3D-drift
diffusion baseline models were developed for the Lumerical suite and a 1D electrical model for SCAPS, allowing for an
accurate description of the optoelectronic behavior and response of thin and ultrathin CIGS solar cells. As a result, it was
possible to obtain accurate descriptions of the optoelectronic behavior of thin and ultrathin solar cells, and to perform an
optical study and optimization of novel light management approaches, such as, random texturization, photonic
nanostructures, plasmonic nanoparticles, among others. The developed light management architectures enabled to push
the optical performance of an ultrathin solar cell and even surpass the performance of a thin film solar cell, enabling a
short-circuit current enhancement of 6.15 mA/cm2 over an ultrathin reference device, without any light management
integrated.

Keywords: CIGS Technology, Light management, Lumerical, FDTD, SCAPS, Lumerical CHARGE

1. INTRODUCTION
Currently, Cu(In,Ga)Se2 (CIGS) thin film solar cells are a mature photovoltaic technology, with a power conversion
efficiency world record value of 23.35 % 1. In the last decade, CIGS thin film solar cells world record efficiency value was
surpassed 11 times, showing the technology’s notable progress 2. Nonetheless, CIGS is a complex quaternary compound,
being its solar cell architecture no less complicated 3,4. The technology evolution results from a continuous search,
exploration, and implementation of different approaches to improve CIGS optoelectronic properties, such as: out of
stoichiometric growth, alkali doping, bandgap engineering, among others 5,6.
The CIGS exquisite optoelectronic properties and well-established performance raised the interest in an additional branch
to thin film technology, where a drastic decrease of the CIGS layer thickness from a thin (2000 nm) absorber to a sub-
micrometer range, has been developed, welcoming the ultrathin technology. Regardless of the additional fundamental and
technological challenges, which were brought up from a CIGS thickness lower than the necessary for a complete solar
spectrum harvested in the absorber, the transfer to an ultrathin technology requires a deep knowledge on the thin film solar
cell technology.

Physics, Simulation, and Photonic Engineering of Photovoltaic Devices X, edited by Alexandre Freundlich,
Stéphane Collin, Karin Hinzer, Proc. of SPIE Vol. 11681, 1168108 · © 2021 SPIE
CCC code: 0277-786X/21/$21 · doi: 10.1117/12.2577650

Proc. of SPIE Vol. 11681 1168108-1


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Computational simulation, or also called in-silico experiments, has become widely accepted as “Third Pillar of 21st Century
Science”, along with theory and experimentation. Within this framework, the CIGS technology has used accurate and
efficient optical and electrical simulations to better understand and overcome the complexity of the CIGS solar cell system
and for orientation through novel research directions 7–14. To extend the CIGS understanding from the thin to an ultrathin
thickness range, while maintaining the performance of the solar cell, optical and electrical simulations play a key role on
understanding the difference in several phenomena, such as, light propagation, interface scattering, and carrier dynamics.
However, an accurate output from the employed simulations is undoubtedly dependent on the need of the device modelling
to keep pace with the technology evolution.
In this manuscript, a combination of optical and electrical simulations is presented, in order to describe the two main
phenomena governing a solar cell’s working principle: light absorption and charge carrier collection. For that purpose, and
considering that we are modelling light interaction in a medium with a thickness comparable to the light wavelength values,
a finite-difference time-domain (FDTD) method that solves Maxwell’s equations in the time domain was performed with
the Lumerical commercial software suite 15. The optical properties from Lumerical FDTD were used to feed the one
dimensional Solar Cell Capacitance Simulator (SCAPS). This robust software developed at the University of Gent, is freely
available, and used to numerically simulate different types of solar cells 8,16–18. For CIGS technology, SCAPS has been
used to study the influence of physical and electronic parameters from the different layers on the output performance of
the solar cell. A robust CIGS SCAPS model that describes today’s technology for a thin film approach is presented.
Complementing the electrical studies performed with SCAPS, simulations were conducted using Lumerical CHARGE, a
highly versatile three-dimensional (3D) charge transport simulator. It is capable of ingesting the highly precise FDTD
results and simulate accurately complex structures. Furthermore, Lumerical CHARGE allows for the solar cell’s
architecture to be optimized not only for increased optical gains, but for a superior electrical performance as well.
This manuscript presents three sections in addition to the introduction and the conclusion: FDTD Optical Simulation
Model, FDTD Simulations for Solar Cell Light Management Architectures, and Electrical Simulations Support. In the
FDTD Optical Simulations Section a complete overview of the optical simulations model with the demonstration of all the
possible obtainable physical quantities is demonstrated. In the FDTD Simulations for Solar Cell Light Management
Architectures Section, several suitable light management strategies integrated in CIGS ultrathin solar cell are reviewed and
optimized in order to outperform its thin film counter-part. Finally, In the Electrical Simulations Support is shown how
FDTD simulations output are of paramount importance to accurately describe electrically thin and ultrathin CIGS solar
cells with conventional and non-conventional architectures using two electrical simulators, SCAPS and Lumerical
CHARGE. This manuscript shows 3D FDTD simulations potential in the study of charge carrier generation in thin and
ultrathin CIGS solar cells, as well as an exploratory tool to evaluate the impact of CIGS exotic architectures for light
management incorporation. In summary, it was shown that an accurate optical description of the devices is the first step
to more complex simulations to evaluate and understand solar cell principles of operation and respective performance.

2. FDTD OPTICAL SIMULATION MODEL


The optical simulations were performed by a 3D solver of the Maxwell equations, more precisely the FDTD method, which
enables the resolution of the Maxwell equations in complex geometries, providing direct space and time solutions 15,19.
The FDTD solutions package available in the Lumerical software was used to perform the optical simulations 20. The
FDTD solutions method recurs to Fourier transforms to accurately calculate the frequency dependent electromagnetic
fields, returning the complex valued fields and normalized transmissions as a function of wavelength 19. The FDTD method
is used in this work to optically describe CIGS solar cells. The typical reference solar cell architecture consists on the
following structure: Mo (350 nm), MoSe2 (5 nm), CIGS, CdS (50 nm), i-ZnO (100 nm) and Al:ZnO (AZO) (400 nm). In
Figure 1 a schematic representation the described CIGS solar cell stack with a typical thin (2000 nm) and ultrathin (500
nm) absorber is shown in a) and b), respectively. In this Section, the potential of the optical model used is explored through
the demonstration of all the physical parameter values possible to be extracted by the simulation model.
In order to simulate the solar cell stack, the complex refractive index (𝑛(𝜆) = 𝑛 + 𝑖𝑘), the refractive index 𝑛 and the
extinction coefficient 𝑘, of each material needs to be used as input. The CIGS, with [Ga]/[Ga+In] (GGI) = 0.3, optical
properties were obtained through in-house spectroscopic ellipsometry measurements. The AZO, i-ZnO, and CdS
compounds were taken from reference 21, the MoSe2 layer from 22, and the Mo layer from 23. A broadband plane wave
source is used to introduce a uniform electromagnetic field into the solar cell stack in a wavelength range from 300 to 1100
nm. Depending on the simulated solar cell architecture, different Cartesian mesh sizes have been used. However, the mesh
size considered is the smallest possible to respect the memory and time requirements of each simulation. Furthermore,

Proc. of SPIE Vol. 11681 1168108-2


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
override mesh regions are commonly used at critical interfaces or nanostructures to provide the best simulation accuracy.
Typically, the simulated structures present periodicity, and, therefore, to reduce the simulation time, symmetric and anti-
symmetric boundary conditions are also used, enabling to simulate only ¼ of the total space.
The FDTD optical simulation software allows for the calculation of the total light absorbed in each layer. For that purpose,
the power absorbed per unit volume (P) is calculated in each layer, as following 24:
1
P= ωε′′ |E(λ)|2 (1)
2

where, 𝜔 is the angular frequency, 𝜀 ′′ the imaginary part of the dielectric permittivity, and |𝐸(𝜆)|2 is the electric field
intensity as a function of the wavelength value. The light absorbed in each layer per wavelength value (𝐴𝑏𝑠(𝜆)), is then
calculated through the integration of the normalized 𝑃 over the solar cell spectrum. Using 𝐴𝑏𝑠(𝜆), a spectrum containing
the absorption in each layer of the solar cell stack is obtained, as demonstrated in Figure 1 a) and b) for thin and ultrathin
CIGS solar cells, respectively. These plots are essential to accurately evaluate the solar cell optical performance, as it
enables the verification of individual layers’ optical losses at specific wavelength values, as well as, to check improved
optical performances, obtained through the integration of light management structures. For the studied CIGS architectures,
the light collection by the absorber layer is significantly reduced due to parasitic absorption in the solar cell remaining
layers. The transparent conductive oxides (TCO), both in the thin and ultrathin configuration, demonstrate a significant
parasitic absorption in both UV and NIR regions. Through the optimization of the contact layer thicknesses, this parasitic
absorption can be reduced 25. The CdS parasitic absorption stems from the relatively low bandgap energy value of this
buffer layer (~2.4 eV) 26, and is as well thickness dependent. Other buffer materials that have larger bandgap energy values
are currently being investigated in CIGS solar cell devices 27,28. Notably, the solar cell architecture that led to the CIGS
solar cell world record efficiency value (23.35 %) uses an Zn(O,S,OH) x /Zn0.8 Mg0.2 buffer layer 1. An increase in the
detrimental absorption in the Mo layer, as the absorber thickness is reduced from 2000 to 500 nm is observed. Such
parasitic absorption is an indication that ultrathin CIGS absorbers are unable to fully absorb the incoming light in one
single passage. Thus, light management architectures need to be developed to increase the optical path inside the CIGS
absorber, as will be shown in Section 2.

100
a) TCO - Jsc = 2.72 mA/cm2
80 CdS - Jsc = 1.92 mA/cm2
CIGS - Jsc = 32.70 mA/cm2
Absorption (%)

Mo+MoSe2 - Jsc = 1.53 mA/cm2


60

40

20

0
400 600 800 1000
Incoming light wavelength (nm)
100

b) TCO - Jsc = 2.75 mA/cm2


80 CdS - Jsc = 1.92 mA/cm2
CIGS - Jsc = 28.81 mA/cm2
Absorption (%)

Mo+MoSe2 - Jsc = 4.36 mA/cm2


60

40

20

0
400 600 800 1000
Incoming light wavelength (nm)

Figure 1. Absorption in the solar stack layers and schematic CIGS solar stack representation for a) thin (2000 nm) and b) ultrathin
(500 nm) absorber.

Proc. of SPIE Vol. 11681 1168108-3


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
From the absorbed light, quantities like the charge carrier generation rate (𝐺) and short-circuit current density (𝐽𝑠𝑐 ) can be
obtained, through
λ λ
G = ∫λ max Abs(λ) IAM1.5G (λ)dλ (2)
hc min

Jsc = qG (3)

where, h is the Planck’s constant, c is the speed of light, IAM1.5G (λ) is the AM1.5G solar spectrum and 𝑞 is the elementary
charge. A quantification of the solar cell’s optical performance is defined by its 𝐽𝑠𝑐 value. This value will be the maximum
current density that the CIGS solar cell will achieve considering only the optical properties of the solar cell architecture,
disregarding non ideal effects on the transport and extraction of charge carriers. The 𝐺 parameter may be represented in
1D or 2D profiles as a function of position and/or incoming light wavelength values, as represented in Figures 2 for CIGS
thin a) / c), and ultrathin b) / d), respectively.

1.6x1028 1.6x1028

a) Generation Rate 2000 nm b) Generation Rate 500 nm

Generation Rate (m-3 s-1)


Generation Rate (m-3 s-1)

1.2x1028 1.2x1028

8.0x1027 8.0x1027

4.0x1027 4.0x1027

0.0 0.0
2000 1800 1600 1400 1200 1000 800 600 400 200 0 500 400 300 200 100 0
CdS CIGS Thickness (nm) CdS CIGS Thickness (nm) Mo
Mo

2000
500
c) G -2.000E+23 d) G -2.000E+23
CIGS Thickness (nm)

1600 5.850E+24 400 5.850E+24


CIGS Thickness (nm)

1.190E+25
1.190E+25
1200 1.795E+25
300 1.795E+25
-3 -1
2.400E+25 (m s )
2.400E+25
(m-3 s-1)
3.005E+25
800 3.005E+25
3.610E+25
200
3.610E+25
4.215E+25
4.215E+25
400 4.820E+25 100 4.820E+25

CIGS (2000 nm) CIGS (500 nm)


0
0
400 600 800 1000
400 600 800 1000
Incoming light wavelength (nm) Incoming light wavelength (nm)

Figure 2. 1D Generation rate profile as a function of CIGS thickness in, a) a thin (2000 nm) and b) ultrathin (500 nm)
absorber; 2D Generation rate profile as a function of the incoming light wavelength in: c) a thin and d) ultrathin absorber.
A transmission monitor is always placed above the solar cell stack, to quantify the total solar cell reflection (Figure 3 a)),
which can be used to evaluate anti-reflection properties as well as light trapping effects of different light management
architectures. In this case, in the ultrathin absorber (500 nm) there is a higher reflectance at the NIR wavelength range,
stemming from light leaving the solar cell after being reflected from the back contact, that otherwise would be absorbed
in the thicker 2000 nm absorber. The electric field intensity profile in the absorber layer (Figure 3 b)) or at any other
layer/interface, giving insight into the effect of light concentration or scattering by different nanostructures used for light
management, can also be obtained. A Transfer Matrix analytical formalism is used to validate the reference solar stack
simulations accuracy, through the comparison of the simulated light absorption in the solar cell stack with the calculated
absorption from the Transfer Matrix Analytical formalism (Figure 3 c)). The almost perfect overlap between the simulated
and analytical absorptions, demonstrate the accuracy and viability of these simulations and opens the door for more
complex geometries and exotic light management architectures. The optical performance of a solar cell stack with complex
geometries can be fully characterized through all the simulated quantities presented in this Section. Furthermore, various

Proc. of SPIE Vol. 11681 1168108-4


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
parameters obtained from the optical simulations can be used to improve the accuracy of electric simulations performed
either with SCAPS or Lumerical CHARGE, as it will be shown in Section 3.
100 500

a) b) |E|2
Thin (2000 nm) 0.1940

Total Solar Cell Reflectance (%)


400
75 Ultrathin (500 nm) 0.2565

CIGS Thickness (nm)


0.3190
300 0.3815
(V/m)2
50 0.4440
0.5065
200
0.5690
25 0.6315
100
0.6940
CIGS (500 nm)
0 0
400 600 800 1000 -1000 -500 0 500 1000
Incoming light wavelength (nm)
Lateral size (nm)
100
c)
Total Solar Cell Absorption (%)

80

60

40
Transfer matrix analytical formalism (500 nm)
FDTD simulated absorbance (500 nm)
20 Transfer matrix analytical formalism (2000 nm)
FDTD simulated absorbance (2000 nm)

0
400 600 800 1000
Incoming light wavelength (nm)

Figure 3.a) Total solar cell reflection of both the thin and ultrathin devices; b) Electric field intensity profile in an ultrathin
absorber at a specific y plane in the CIGS absorber; and c) Total solar cell absorption from the transfer matrix formalism and
from the FDTD simulation in an thin and ultrathin solar cell devices.

3. FDTD SIMULATIONS FOR SOLAR CELL LIGHT MANAGEMENT ARCHITECTURES


One of the biggest challenges to attain conversion efficiency values that match the Shockley-Queisser limit is the optical
losses present in solar cell devices 29. Light reflection, parasitic absorption and absorption losses that either prevent the
entrance of light in the absorber material or hinder an efficient light absorption, need to be minimized 25. Light trapping
schemes are then essential to enhance the efficiency of solar cell devices. Their integration may be done through many
ways, with the intent of increasing the solar cells efficiency value, by reducing the external reflection of light and increasing
the internal optical path length. Over the past years, light trapping architectures have been thoroughly studied for
application in thin film solar cells, since it allows them to be viable competitors to their bulk counter-parts 30–32. Applying
light trapping structures in a solar cell allows for the absorption of the same quantity of sunlight in a much thinner absorber
material, which leads to a reduction of material and production costs on photovoltaics (PV) devices, helping the mass
production of solar cells composed by scarce materials, such as, CIGS and CdTe solar cells 33. In this Section, a review on
light management schemes that can be exploited to boost the optical performance of an ultrathin CIGS solar cell stack is
performed, recurring to highly precise 3D optical FDTD simulations. Currently, thin film CIGS solar cell modules have
an absorber thickness of about 2000 nm. Economical and sustainable industry imposed goals drive the reduction of the
absorber thickness to sub-micrometre thicknesses, in what is called the ultrathin regime 33. In Figure 4, the light absorbed
in two typical CIGS solar cell stacks (reference), with an absorber thickness of 2000 nm (thin) and 500 nm (ultrathin)
without any light management architectures is shown. At longer wavelengths (600 to 1100 nm), the optical performance
of the thinner absorber starts to degrade in comparison to the 2000 nm thick absorber. Such effect is related to the
incomplete light absorption by the solar cell absorber, as the CIGS layer is not sufficiently thick to accommodate the
penetration depth of NIR light in one single pass. By reducing the absorber thickness from 2000 nm down to 500 nm an
almost 4 mA/cm2 loss on the 𝐽𝑆𝐶 value is observed. In order to achieve or surpass the optical performance of conventional
2000 nm CIGS solar cell stacks two main issues may be tackled: (1) reflection at the front surface and (2) poor absorption
at the NIR range.

Proc. of SPIE Vol. 11681 1168108-5


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
100
Simulated CIGS Absorption (%)

80

60

40
Ultrathin reference - Jsc = 28.81 mA/cm2
Thin reference - Jsc = 32.70 mA/cm2
20

0
400 600 800 1000
Incoming light wavelength (nm)
Figure 4. Simulated CIGS absorption of two solar cell stacks with 2000 and 500 nm absorber layer thickness and the obtained JSC
value of 32.70 and 28.81 mA/cm2, respectively. Schematic representation of the CIGS solar cells stack architectures.

Despite of the typical CIGS architecture having an AZO/i-ZnO/CdS top layer stack, that already offers a good 𝑛 match
with the absorber, various strategies can be implemented to achieve a better anti-reflection performance 25. The most simple
and common approach consists of using interference type Anti-Reflection (AR) layers at the top of the solar cell device
34,35
. These thin dielectric layers can successfully minimize reflections at specific wavelength values by allowing a
destructive interference of the light reflected from the front and rear sides of the AR layer 35. In order to do so, the thickness
of the AR layer (𝑑𝐴𝑅 ) should correspond to one quarter of the wavelength range to be minimized:
λ0
dAR = (4)
4n1

where 𝜆0 is the wavelength value for optimization and 𝑛1 is the AR layer refractive index. Normally an MgF2 anti-reflection
layer (𝑛 = 1.4 at 600 nm) is used in laboratory CIGS solar cells, as it allows for a good refractive index match between the
air (𝑛 = 1) and the AZO contact (𝑛 = 1.9 at 600 nm). To optimize the performance of an MgF2 AR layer, a particle swarm
optimization algorithm included in the FDTD solutions software was used 35. With this approach, the AR layer thickness
was optimized to attain the maximum JSC value in a CIGS ultrathin solar cell. In Figure 5, simulated CIGS absorption in
an ultrathin device with and without the optimized MgF2 reflection layer is shown, alongside the simulated reflectance.
The optimized MgF2 thickness corresponds to 114 nm. With the implementation of a 114 nm MgF 2 layer, an overall
broadband anti-reflection improvement is achieved, as shown by the reduced reflection throughout the simulated spectrum
in Figure 5, leading to a JSC improvement of 2.59 mA/cm2.

100 100
Simulated CIGS Absorption (%)

Simulated Solar cell Reflection (%)

80 80

60 60

Ultrathin reference - Jsc = 28.81 mA/cm2


40 40
AR - Jsc = 31.40 mA/cm2
Reflection ultrathin reference
Reflection AR
20 20

0 0
400 600 800 1000
Incoming light wavelength (nm)

Figure 5. Simulated CIGS absorption and total solar cell reflection of an ultrathin reference and ultrathin solar cell with an optimized
AR layer. Schematic representation of the CIGS solar cells stack architecture with an MgF2 AR layer implemented.

Proc. of SPIE Vol. 11681 1168108-6


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Another anti-reflection approach consists in introducing texturization on the solar cell top layers. To study the optical
benefits of such approach, two different architectures were implemented and studied by FDTD simulations: i) one
consisting on a random texturized AZO surface, that can practically be achieved in wet-etching procedures through acidic
solutions, enabling an anisotropic etching 36–39; ii) an approach with higher fabrication complexity, consisting on a
nanostructured AZO layer with a periodic 2D triangular grating.
The random texturization was implemented by using the surface roughness structure present in the FDTD Lumerical
software object library. This object allows to modulate the surface roughness through the control of two parameters, the
specified RMS roughness (𝜎𝑅𝑀𝑆 ) and correlation length (𝐿𝑐 ), which determines the overall size of the craters at the surface
- a surface with a large 𝐿𝐶 value will be smother and with larger craters, than a surface with a small 𝐿𝐶 value. In initial
studies, it was verified that the AR performance is enhanced both by increasing the 𝜎𝑅𝑀𝑆 and decreasing the 𝐿𝑐 values,
therefore the optimization goes on establishing a lower surface roughness limit. For these simulations, a 𝜎𝑅𝑀𝑆 value of 100
nm and a 𝐿𝑐 of 100 nm, that is in accordance with the structural features verified in wet-etch procedures of ZnO and AZO
layers in acidic solutions, was applied to the 400 nm AZO layer 40–42. The simulated absorption in the CIGS layer, alongside
the simulated reflection, for an ultrathin reference and an ultrathin device with a random textured AZO layer device are
shown in Figure 6 a). With random texturization, a 𝐽𝑆𝐶 value of 32.09 mA/cm2, which corresponds to a 3.28 mA/cm2
increase over the ultrathin reference device, was obtained. The 𝐽𝑆𝐶 enhancement is due to the improvement of the AR
behavior with an increasingly rougher surface, as seen by the broadband decrease of the measured reflectance. This is
shown in Figures 6 b) and c), where the electric field intensity at the top surface of the ultrathin reference solar cell and
the random texturized surface at 600 nm, respectively, is shown. In Figure 6 b) interference fringes stemming from the
interaction between the incident light with the light reflected from a perfectly flat surface are present, while in Figure 6 c)
the reflection is significantly decreased with the textured surface. Thus, having a rougher surface gives rise to an increased
number of multiple reflections in the created slopes leading to the improvement of the AR properties of the solar cell.

100 100
Simulated CIGS Absorption (%)

Simulated Solar cell Reflection (%)

a)
80 80

60 60
2
Ultrathin reference - Jsc= 28.81 mA/cm
Random Texturization - Jsc= 32.09 mA/cm2
40 40
Reflection ultrathin reference
Reflection Random texturization

20 20

0 0
400 600 800 1000
Incoming light wavelength (nm)
800 800

700 b) Air 700 c) Air


0.4550 0.07000
|E|2
600 0.6050
600
|E|2 0.3550
Height (nm)

0.7550 0.6400
Height (nm)

500 500
0.9050 0.9250

400 1.055
(V/m)2 400 1.210 (V/m)2
1.205 1.495
300 300
1.355 1.780
200 1.505 200 2.065

100 AZO 1.655


100
AZO
2.350

0
-400 -200 0 200 400 -400 -200 0 200 400

Lateral size (nm) Lateral size (nm)

Figure 6. a) Simulated CIGS absorption and total solar cell reflection of an ultrathin reference and ultrathin solar cell with an
optimized random texturized AZO layer; and electric field intensity at 600 nm on the air/AZO interface for b) an ultrathin
reference and c) an ultrathin solar cell with an optimized random texturized AZO layer. Schematic representation of the
CIGS solar cell stack architecture with a random texturized AZO layer.

Proc. of SPIE Vol. 11681 1168108-7


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
An alternative approach is the use of nanoscale periodic structures. With such periodic structures, a more precise control
over the light management mechanisms can be achieved, just through the variation of some morphological parameters 43.
When the texture periodicity (𝛬) is much larger than the incident wavelength (𝛬 >> 𝜆), the light will be scattered into
various diffraction orders 44. The intensity of such diffraction orders depends on the texture’s shape and height. This
contributes to an enhancement of the optical path length at the absorber layer, as the light enters the solar cell in oblique
directions 44. However, such scattering effects also enable a larger optical path in the solar cell window and buffer layers,
increasing the parasitic absorption in these layers 45,46. Furthermore, such large periodic structures enable multiple
reflections within the large textures allowing for a reflection decrease 44. When the structural features present critical
dimensions much smaller than the incident wavelength (𝛬 << 𝜆), the light is not scattered into different diffraction orders,
but rather achieves the best anti-reflection performance. This happens because with such small dimensions, the effective
medium theory predicts that the nanostructures behave as a single layer with a refractive index optimized to achieve a
broadband AR behavior 44. In order to demonstrate this light management concept, optical simulations of a solar cell stack
with a 2D triangular trench pattern on the AZO were performed. For this approach, different period values were used in
the trench pattern with a height of 300 nm, to demonstrate effect of this morphological parameter on light guidance. The
first approach was to simulate different period values to demonstrate the two extreme cases (𝛬 >> 𝜆 and 𝛬 << 𝜆). For
this purpose, a period of 50 and 1000 nm were chosen. For the latter value, the simulation memory and time requirements
were considered, as increasing the period involved a larger simulation volume that would be unfeasible. In Figure 7 a) and
b), the electric field intensity inside the absorber layer is represented at a wavelength value of 1000 nm, for both the 50
and 1000 nm period. Both representations demonstrate interference effects, that happen because the simulated absorber
only has 500 nm, so inevitably, a portion of the light will be reflected from the Mo rear contact back to the absorber creating
the observed interference fringes. The most important observation from the electric field profiles, is that the 50 nm period
structures behave collectively as a flat surface, as predicted by the effective medium theory 44, while in the structure with
a period of 1000 nm a clear diffraction effect from the triangular grating can be verified. The absorption and reflection of
both structures were simulated and compared with the ultrathin reference device, as shown on Figure 7 c). From the
simulated reflection values, it is clear the better AR response arising from the nanoscale texturization, as a reflection close
to 0 % is achieved for the majority of the region of interest of the solar spectrum. As explained before, this broadband AR
effect arises from the small dimensions of the nanoscale texturization in comparison with the wavelength value. The
nanoscale features behave as if light interacts with a homogeneous medium with a gradual refractive index, more precisely
as if multiple thin ARC layers are stacked on top of each other with an optimal refractive index profile 47. The reflection
increase over the NIR wavelength range can be justified from the light reflected in the back contact and exits the solar cell
through the top surface. The slight absorption increases in the NIR region, as well as the lower reflection values at this
range, demonstrate the better scattering performance of the texturized structure with 1000 nm period. The better AR
performance of the solar cell device with a 50 nm period led to a 3.6 mA/cm2 increase of the JSC value, over the ultrathin
reference solar cell. However, the complexity of a periodic structure with 50 nm period may hinder the fabrication of such
texturized surface, as it is hard to achieve this dense array of nanostructures with a high aspect ratio over large areas.

Proc. of SPIE Vol. 11681 1168108-8


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
500

a) |E|2 0.2480
400 0.3087
CIGS Thickness (nm)

0.3695

300 0.4302
(V/m)2
0.4910
0.5518
200
0.6125
0.6733
100 0.7340

CIGS
0
-1000 -500 0 500 1000
Lateral size (nm)
100 100
500 c)
|E|2

Simulated CIGS absorption (%)

Simulated solar cell Reflection (%)


b) 80 80
0.03400
400
0.1172
CIGS Thickness (nm)

0.2005
60 60
300 0.2838
(V/m)2
0.3670
Ultrathin reference - Jsc = 28.81 mA/cm2
200 0.4503 40 1000 nm pith - Jsc = 30.87 mA/cm2 40
50 nm pitch - Jsc = 32.41 mA/cm2
0.5335 Reflection ultrathin reference
Reflection 1000 nm pitch
0.6168
100 Reflection 50 nm
0.7000 20 20
CIGS
0
-1000 -500 0 500 1000 0 0
Lateral size (nm) 400 600 800 1000
Incoming light wavelength (nm)
Figure 7. Electric field intensity at 1000 nm in the CIGS layer for a) periodic structure with a 50 nm period and b) periodic
structure with a 1000 nm period; c) simulated absorption in the CIGS layer and total solar cell reflection in the three
different studied architectures: 2D triangular grating with a 50 nm period, 2D triangular grating with a 1000 nm period and a
reference solar cell with no texturization. Schematic representation of the CIGS solar cells stack architecture with a 2D
triangular array pattern AZO layer.
Another critical optical loss is the poor light absorption in the NIR region. To tackle such issue, light management strategies
are employed to enhance the optical path length inside the absorber layer. Light management architectures can be employed
at the back interface and enhance this interface reflection/scattering capabilities. The simpler approach consists of adding
a highly reflective metallic layer as the solar cell back contact, enabling to double the light optical path inside the solar cell
25
. However, in CIGS solar cell devices, employing highly reflective layers, such as Al, Ag or Au can be difficult. During
the absorber deposition, the metals are submitted to high temperatures and may diffuse to the absorber layer, or react with
Se, degrading the solar cell electrical performance 14,48. Commonly a Mo layer is used as back contact, which is a more
chemically and thermally stable metal. Nevertheless, the Mo reflection lacks in comparison with other metals, such as the
one aforementioned 49–52. To study the optical benefits of a metallic layer on a CIGS ultrathin solar cell device it is
imperative that a dielectric layer is added on top of the optical mirror. The following rear contact architecture was
employed: Mo/Ag (25 nm)/SiO2 (20 nm). It is important to note that in this approach, the influence of a contact architecture
is not studied, as a purely optical study is being performed in this Section. Furthermore, here the metal of choice is Ag,
since this structure will be integrated in an architecture with Ag plasmonic nanoparticles (NPs) further down in this section.
In Figure 8, the simulated CIGS absorption, alongside the total solar cell reflectance of the solar cell with the metallic layer
and the ultrathin reference is presented. With the introduction of the metallic layer, there is a 1.49 mA/cm2 improvement
over the reference JSC value since it allows an additional pass of the light that wasn’t absorbed in the CIGS layer during
the first pass. Furthermore, there is more light leaving the solar cell, as shown by the increased reflection in the NIR.
However, adding the metallic layer only doubles the optical path, as it works as an optical mirror, only allowing an increase
in the light’s specular reflection 53,54.

Proc. of SPIE Vol. 11681 1168108-9


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
100 100

Simulated CIGS Absorption (%)

Simulated Solar cell Reflection (%)


80 80

60 60

Ultrathin reference - Jsc= 28.81 mA/cm2


40 Metal - Jsc= 30.30 mA/cm2
40
Reflection Ultrathin reference
Reflection Metal
20 20

0 0
400 600 800 1000
Incoming light wavelength (nm)
Figure 8. Simulated absorption in the CIGS layer and total solar cell reflection of the ultrathin reference and the solar cell
device with the metallic layer at the rear contact. Schematic representation of the CIGS solar cells stack architecture with an
Ag metallic/dielectric based substrate.
Other light trapping concepts need to be explored to further enhance the light path inside the absorber 25. The surface
texture of the back contact is also a potential strategy to increase light absorption in thin film solar cells. For example,
Huang et al. produced in 55 a highly scattering rear Si surface texture by plasma etching of poly-Si thin film solar cells.
Using the plasma-etching technique, sub-micron textures at the rear surface were produced and values of haze reflection
higher than 95 % at the Si-air interface were achieved 55. However, in thin film solar cells such texturization can degrade
the overall cell performance by creating defects in the absorber material 56. Another approach consists in exploring the
plasmonic resonances of metallic nanoparticles. When light reaches the metallic nanoparticle, the electric field will lead
the metal’s free electrons to the nanoparticle surface. Since these entities are confined, positive charges will be accumulated
at the opposite surface creating an electric dipole. This charge displacement generates an electric field that induces an
oscillatory behavior of the electron cloud 57. At certain wavelength values, the incident light can be in resonance with the
metals free electrons oscillations, leading to resonant oscillations of the electron cloud that can generate high scattering or
absorption cross-sections, depending on the NPs size, shape and dielectric medium 57,58. The nanoparticle resonance can
be tuned to enhance the scattering effect at the NIR region, where the thin film solar cells absorber thickness is not sufficient
to accommodate the long wavelength’s penetration depth. To redshift the nanoparticle resonant wavelength value, large
sizes (around 50 nm radius) and elongated shapes are preferred 57,58. Usually these plasmonic entities are integrated in the
solar cell structure accompanied by a metallic layer, creating a so-called plasmonic back contact 59. This way, the light
scattered backwards by the NPs can be reemitted to the solar cell’s absorber.
For a viable integration of the Ag plasmonic nanomaterials, one must study the scattering cross-sections at the localized
surface plasmon resonance (LSPR). In order to do so, a total-field scattered-field (TFSF) source with a wavelength range
between 300 to 1100 nm was used to illuminate Ag NPs with a radius of 50 nm. The TFSF source allows to study the
scattering and absorption behavior of the spherical NPs, as it can separate the NP scattered field from the incident
electromagnetic field 19. The scattering and absorption efficiency of a NP can be calculated by normalizing the scattering
(CS) and absorption cross-section (CA) to the NP volume, using the following equation:
PA,S
CA,S = (5)
IAM1.5

where PA,S is the net power absorbed/scattered by the NP. It is important to keep in mind that the refractive index of the
medium where the Ag NPs will be inserted is different than 1 (air). The CIGS absorber has a higher refractive index close
to 3. Furthermore, the NPs will be encapsulated by a dielectric material. Therefore, for a proper study of the effective
scattering behavior of the Ag NP (Figure 9) located inside the solar cell architecture, a refractive index of 3 was used and
an oxide shell of different dielectric materials (SiO2, Al2O3, and TiO2) was placed surrounding the NP. The resonant
wavelengths of the stand-alone Ag NP are within the desired NIR wavelength range. However, one must consider that the
NPs cannot be implemented directly in contact with the CIGS layer, since Ag would not withstand the harsh growth
conditions of the absorber. Therefore, the NPs need to be encapsulated with a dielectric layer. On the other hand, this will
affect the resonant properties of this nanostructures, as the local refractive index of the medium is different. The oxide
shells have a lower refractive index than the CIGS, so a blue-shift and dampening of the scattering cross-section is verified

Proc. of SPIE Vol. 11681 1168108-10


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
with the implementation of every oxide shell 60. SiO2 has the lowest refractive index, leading to the largest blue-shift and
dampening effect, while the opposed is verified in TiO 2 that has the lowest dampening and blue-shift. As such, for an
improved optical performance, a TiO2 shell holds more promise, as it is the configuration with the highest scattering
efficiency matching the wavelength range where most optical losses will occur due to incomplete light absorption in an
ultrathin solar cell device.
6 r = 50 nm No shell n=3
Ag NP
r = 50 nm TiO2 shell
r = 50 nm Al2O3 shell

Scattering efficiency (arb. un.)


5 r = 50 nm SiO2 shell

3
Ag NP/dielectric
2

0
400 600 800 1000
Incoming light wavelength (nm)

Figure 9. Scattering efficiency of 50 nm radius Ag NPs incorporated in a medium with 𝑛 = 3: without an oxide shell, with a
10 nm SiO2, Al2O3 and TiO2 shell.
A periodic structure of Ag NPs displayed in a rectangular lattice was integrated in an ultrathin CIGS solar cell. A metallic
layer was placed below the NPs, to collect the light that is scattered backwards and would be absorbed by the Mo layer.
The NPs were integrated into the solar cell through a volume compensation approach, i.e. the occupied volume by the NPs
was added to the thickness of the CIGS layer without any texturization transfer. This method was used to decouple the AR
effects that would occur from the front-surface, if the NPs texturization was translated to the upper-most layers. Previous
to the solar cell stack simulation, an optimization sweep was performed to study the best distance between the metallic
NPs, in order to obtain the NP surface coverage presenting the highest scattering efficiency 61. This optimization was
achieved once again through a particle swarm optimization algorithm included in the FDTD solutions software 19. The
CIGS simulated absorption, of three solar cell devices with this plasmonic configuration is represented in Figure 10 a).
When comparing the performance of the different dielectric layers, the configuration with the TiO2 shell, with the highest
dielectric constant, presents the highest simulated 𝐽𝑆𝐶 value (31.04 mA/cm2). As demonstrated before, in Figure 9 when
the NPs are encapsulated in a TiO2 shell, the best resonant matching and the best scattering performance for a single NP is
achieved. However, it is important to consider the effect of this dielectric layer in the solar cell electrical performance.
While Al2O3 and SiO2 dielectric layers have been successfully employed and extensively studied in CIGS solar cells as
surface passivation layers 48,54,62–66, TiO2 was not. The employment of TiO2 in this optical study served, however, to
demonstrate the benefits of using oxide layers with a higher refractive index. In Figure 10 b), the CIGS simulated
absorption of the ultrathin reference, a solar cell stack with a metallic layer and the best performing plasmonic configuration
are presented. With the plasmonic configuration, a 2.23 mA/cm2 increase of the 𝐽𝑆𝐶 value can be achieved over a reference
solar cell, and around 0.74 mA/cm2 when compared with a structure with only a metallic layer. Such enhanced optical
performance is a result of the increased scattering at the NIR wavelength range. Comparing the absorption spectrum of the
solar cell with the metallic layer and the plasmonic solar cell, there is an absorption dip over the 900 nm wavelength region
in the solar cell with nanoparticles. This effect is related to a fundamental limit of the plasmonic approach, since at the
resonant wavelength values, the Ag NPs also present a strong absorption cross-section, leading to parasitic absorption that
it is not used to generate photocurrent 25,56. This effect is demonstrated on the higher Ag parasitic absorption in the
plasmonic configuration as opposed to the metal one, also presented in Figure 10 b). Furthermore, the integration of
metallic NPs can also be complex in substrate architectures, as the metal may diffuse into the absorber during its high
temperature deposition 25,56. Therefore, chemically stable and absorption-free dielectric NPs have also been a topic of
interest for scattering entities in solar cell devices at the rear interface 67,68. Besides an optical benefit, such dielectric
nanostructures may provide a passivation effect beneficial for the solar cell optoelectronic properties 25.

Proc. of SPIE Vol. 11681 1168108-11


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
100 100 100
CIGS Simulated Absorption (%)
a) b)

CIGS Simulated Absorption (%)


80 80

Ag parasitic absorption (%)


80

60 60 60

40 40 Ultrathin reference - Jsc = 28.81 mA/cm2 40


Plasmonic with TiO2 - Jsc = 31.04 mA/cm 2 Metal - Jsc = 30.30 mA/cm2
2
Plasmonic - Jsc = 31.04 mA/cm
Plasmonic with Al2O3 - Jsc = 30.68 mA/cm2
20
Plasmonic with SiO2 - Jsc = 30.44 mA/cm2
20 Ag absorption - Metal configuration 20
Ag absorption - Plasmonicconfiguration

0 0 0
400 600 800 1000 400 600 800 1000
Incoming light wavelength (nm) Incoming light wavelength (nm)

Figure 10. a) Simulated absorption in the CIGS layer on solar cell stacks with the plasmonic architecture using three
different oxide layers (SiO2, Al2O3 and TiO2); b) Simulated absorption in the CIGS layer and Ag parasitic absorption in the
three different studied architectures: configuration with only a metallic layer, plasmonic configuration and a reference solar
cell with no light trapping architecture. Schematic representation of the CIGS solar cells stack architecture with an Ag
metallic/dielectric/Ag NPs/ dielectric based substrate.
The most effective light management approaches both to reduce the front surface reflection and to increase the optical path
length value were incorporated in a solar cell stack. In Figure 11 a), the simulated absorption in the CIGS layer of the
optimized architecture (best texturization plus plasmonic) is compared with the reference solar cell devices (ultrathin and
thin ). The broadband AR performance, as well as the increased optical path length, led to a 𝐽𝑠𝑐 value of 34.96 mA/cm2,
2.26 mA/cm2 higher than the thin reference with 2000 nm thick CIGS. The optical path length through the employment of
the best light management architectures is compared with the one of the ultrathin reference (Figure 11 b)). To calculate the
optical path length, a method developed by Hegedus and Shafarmann was followed 69. With the implementation of the
plasmonic back reflector as well as from the texturized AZO layer an optical path length as high as 4 is achieved in the
NIR region. The increase of the optical path length over the ultrathin reference leads to an optical absorption performance
equivalent to the one verified in the NIR range in the thick 2000 nm absorber, as it is seen in Figure 11 a). Therefore, such
light management architectures enable a reduction of the solar cell production costs, since a 4 times lower absorber
thickness may be used without significant optical losses. Furthermore, the developed nanostructures can be produced by
low-cost, industrially friendly approaches, as the trench design can be implemented through nano-imprint lithography and
solution-based depositions can be used to deposit the metallic plasmonic nanoparticles.

100 4.0
Simulated CIGS Absorption (%)

3.5 Ultrathin reference


80 Best Texturization + Plasmonic
Optical Path Length (arb. un.)

3.0

2.5
60

2.0

40
1.5

Ultrathin reference - Jsc = 28.81 mA/cm2 1.0


20 Thin reference - Jsc = 32.70 mA/cm2
2
Best Texturization + Plasmonic - Jsc = 34.96 mA/cm 0.5

0 0.0
400 600 800 1000 400 600 800 1000
Incoming light wavelength (nm) Incoming light wavelength (nm)

Figure 11. a) Simulated absorption in the CIGS layer on reference solar stacks with 500 and 2000 nm absorbers and in a solar cell
stack where the best light management approaches were integrated b) Estimated optical path length in a solar cell stack with the best
light management approaches and a reference solar cell with an absorber thickness of 500 nm. Schematic representation of the CIGS
solar cells stack architecture with full optimized light management integration.

Proc. of SPIE Vol. 11681 1168108-12


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
4. ELECTRICAL SIMULATION SUPPORT
The optical parameters, extracted from the analysis of a solar cell stack with the FDTD software, can be integrated into
electrical simulations to obtain a fully viable optoelectronic model. In this Section, two different examples where FDTD
was used to enhance electrical simulations, will be described. First, the validation of the SCAPS electrical model was
performed through the comparison of the figures of merit of a simulated 2000 nm CIGS solar cell with the current
conversion efficiency record holder with equivalent architecture 70. Second, the FDTD method was used to improve the
electrical simulations of a more complex solar cell architecture in the Lumerical CHARGE software. Here, an
optoelectronic model was developed to study an ultrathin architecture with a patterned metallic/dielectric interlayer
integrated at the Mo rear contact.
At first, the use of two electrical simulators, SCAPS and Lumerical CHARGE, to study CIGS solar cells might seem as a
waste of resources. However, even though each program has its own strong points, both exhibit certain weaknesses that
end up being complemented by the other software solution.
SCAPS software is a 1D solar cell and capacitor simulator, tailored for the simulation of thin film solar cells 18. Therefore,
it incorporates in its models various phenomena typically found in thin film solar cells, such as various recombination
mechanisms located throughout the cell, a multitude of defects levels, and many other parameters 71,72. The end result is
the possibility to develop electrical numerical models with a high degree of correlation with real samples, which in turn,
can be used to better understand the behavior of said samples, and thus, provide a pathway for improvement. Moreover,
not only is this tool freely available to the PV scientific community, as it only requires modest computational resources.
The consequence of these advantages is a multitude of available literature that can act as a reference and as a springboard,
for one’s work into highly advanced electrical studies of thin film solar cell’s architectures. Nonetheless, even though
SCAPS is a powerful tool, it has its own shortcomings as well. The most significant one is the fact that it can only simulate
1D architectures. If the object of study is a line contact or a point contact passivation layer for example, 2D and 3D
structures respectively, SCAPS cannot accurately model said structures, requiring thus, a considerable amount of
assumptions and approximations during the creation of the electrical model of the design in study. Another aspect that can
be lacking, is the optical aspect of the solar cell simulations, as it only accounts for simpler optical models and phenomena
when compared with more specialized optical simulation software, whereas the outsourcing of reliable optical data is
necessary for such SCAPS models.
The Lumerical CHARGE electrical simulator is built on the finite element drift-diffusion method and it is capable of
performing 2D and 3D simulations of custom and highly complex structures. In the case of solar cells, it takes into
consideration the complex optical response of the entire solar cell architecture, by importing the charge carrier generation
rate distribution calculated with Lumerical FDTD. The end result is that CHARGE cannot only calculate the conventional
figures of merit, but it also provides a multi-dimensional distribution of the variation of multiple parameters throughout
the solar cell, such as the band structure, carrier current density, free charge density, electric fields, recombination, carrier’s
lifetime and so on. The ability to incorporate the results of complex thermal simulations, performed with Lumerical HEAT
is available as well, thus allowing the study of the impact of the thermal response of a solar cell in its own behavior.
Additionally, the Lumerical simulation suite offers the possibility to use Particle Swarm optimization algorithms, or other
custom algorithms, to better exploit the available computational resources, and aid the architectural design. However,
once again, this software tool has its own constraints. Unlike SCAPS, Lumerical CHARGE requires the purchase of
software licenses, thus restricting those who can access it. Furthermore, the ability to perform multidimensional simulations
comes at a cost of high computational requirements. As a consequence, this tool is only infrequently found in the thin film
PV literature. Since Lumerical CHARGE was designed with a more universal application range in scope, even though this
approach has several advantages, it also means that, for example, it is not possible to set up semiconductor’s electronic
properties as well tailored towards the intricacies of thin film solar cells, as SCAPS can. One other example is that CIGS,
does not exist in an undoped state, meaning that the ability of CHARGE to automatically calculate the effects of a certain
doping concentration on the characteristics of the intrinsic material, and, thus, the effects on the simulated device, acts as
a constraint in the case of CIGS solar cells. The simulations cannot accurately be run without the doping concentration
simulation object present, thus a work-around needs to be put in place, in order to more accurately simulate the self-doping
CIGS absorber. To overcome this obstacle, the CIGS semiconductor model must initially be defined in an “undoped” state,
with its parameters taking into account that a certain dopant level will be added afterwards, bringing the semiconductor
properties into the desired level. Most existing CIGS models found in the literature for SCAPS, can be used, but require
certain parameters adjustments, as for example, CHARGE requires the work function of a material, instead of the electron
affinity required by SCAPS.

Proc. of SPIE Vol. 11681 1168108-13


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
4.1 SCAPS electrical simulations
The model simulated by SCAPS aims to describe the best performance of the 2000 nm CIGS solar cell proposed
architecture, through the validation of the figures of merit values comparing with the ones experimentally reported for the
champion cell (22.6 %) with equivalent architecture 70. To develop a robust SCAPS model to describe the current CIGS
technology, several parameters have been updated, namely optical properties that were obtained through 3D FDTD optical
simulations. An important accuracy boost in the model was the update of the rear and front optical reflection values through
FDTD simulation reflection data. It is important to note that in SCAPS simulations, a bandgap grading profile in CIGS
was defined according to the experimental champion solar cell. However, in the FDTD simulations it was only considered
the minimum bandgap value. In order to replicate the champion device bandgap profile in the Lumerical FDTD software,
several finite CIGS layers with different optical properties need to be used. The presence of several CIGS layers with
different refractive index values produced additional interference fringes in the optical reflection data at the NIR
wavelength range. Therefore, the replication of the bandgap profile was discarded due to unrealistic interference fringes at
the optical front reflection together with an increased simulation complexity. For this Section, the CIGS optical constants
were taken from 73, where a complete database of the optical properties of CIGS absorber layers with different GGI values
is presented, allowing to simulate the gradient profile existent in the champion solar cell architecture.
The reflection of the Mo substrate was simulated with the FDTD simulation region inside a medium with 𝑛 of 3. This
value corresponds to the mean value of the CIGS material. The fixed refractive index is an approximation of the reality
due to the variation of this index with wavelength (index variation of 2.8 to 3.2). The rear optical reflection obtained
through the 3D FDTD optical simulations and used in the SCAPS-1D model is presented in Figure 12 a), together with the
impact in the external quantum efficiency (EQE) curve, which may be seen by comparing both EQE curves of equivalent
solar cells with implemented rear optical reflection from Lumerical and set at 0 %. With Lumerical’s rear optical reflection,
there is an improvement in the light absorption for the IR region, describing in a more suitable way the behavior of a real
device. Hence, a certain percentage of light in this region of the spectrum reaches the Mo rear contact and is reflected back
to the absorber, enabling a second pass in the CIGS layer, increasing its probability to be absorbed.
The front optical reflection defined in the SCAPS model is also provided by 3D FDTD optical simulations. To obtain an
experimentally based model that can be compared with high-efficiency CIGS solar cells, an AR layer at the front contact
was considered. Furthermore, the thickness of such layer was optimized to achieve the highest 𝐽𝑠𝑐 value. Additionally,
atomic force microscopy experimental measurements provided the CIGS roughness from a complete solar cell, which was
implemented in the FDTD optical simulations. The front optical reflection with the optimized AR layer with a thickness
of 113 nm coupled with CIGS roughness is presented in Figure 12 b). It is important to refer the 1 nm difference between
the thickness attained with this optimization and the optimized AR layer thickness in Section 3. This difference stems from
slightly different thicknesses of the TCO layers and optical properties of the CIGS absorber, as the SCAPS model was
validated through the employment of a highly efficient thin film architecture 70. Figure 12 b) also presents the EQE response
curve obtained with the SCAPS model with fully optimized rear and front reflections, and the EQE curve obtained from
just the optimization of the rear reflection. For comparison purposes, the EQE of the 22.6 % efficient solar cell champion
device is included in the Figure. The SCAPS model with rear and front optical reflections from FDTD simulations produces
an overall good fit with the experimental high-efficiency device regarding the EQE response curves. However, the
interference fringes are different between both curves, which is explained by the different CIGS roughness that is rather
variable and process-dependent. Finally, in Figure 12 c) the comparison between the current density vs voltage (J-V)
characteristic curves obtained by the SCAPS electrical model and the champion cell is shown. The SCAPS model was able
to describe the behavior of an experimental high-efficiency CIGS solar cell device by incorporating the FDTD simulation
results. The observation of the results in Figure 12 c) shows the accuracy of the SCAPS model describing the experimental
high-efficiency device. Hence, the open circuit voltage (𝑉𝑜𝑐 ) and fill factor (𝐹𝐹) values are well modelled, mainly due to
experimentally updated parameters and some included post deposition treatments effects, which will be reported in future
contributions. Moreover, the optical reflection data from FDTD simulations was essential to provide a solid model
regarding the optical performance and depicted by the similar 𝐽𝑠𝑐 values between both J-V characteristic curves. Therefore,
bringing together the electrical and optical simulation tools, FDTD and SCAPS respectively, allows for the creation of a
robust model that can effectively describe experimental champion solar cells.

Proc. of SPIE Vol. 11681 1168108-14


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
100 100 100 100
a)
External Quantum Efficiency (%)
Reflection b)

External Quantum Efficiency (%)


EQE

80 80 80 80
EQE Reflection
+ 0.9 mA/cm2

Front reflection (%)


Reflection (%)
60 60 60 60

40 40 40 EQE - champion cell (22.6%) 40


EQE - Rear reflection = 0 % EQE - Optimized Rear Reflection
EQE - Rear reflection update EQE - Optimized Front and Rear Reflection
Rear reflection from 3D FDTD simulation Front reflection from optimized scaps model

20 20 20 20

0 0 0 0
400 600 800 1000 1200 400 600 800 1000 1200
Incoming light wavelength (nm) Incoming light wavelength (nm)
50
ZSW experimental result
40 c) SCAPS model
30
Current Density (mA/cm2)

20

10

0
Voc FF
Jsc (mA . cm-2) Efficiency (%)
(mV) (%)
-10
SCAPS model 747 37.0 80.3 22.2
-20 ZSW champion cell 741 37.8 80.6 22.6
-30

-40

-50
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Voltage (V)
Figure 12. a) EQE response curves comparing the rear optical reflection of 0 % with the presented Mo rear optical reflection (dashed
line) obtained with the 3D FDTD optical simulations in a 2000 nm thin film CIGS solar cell. Note that the front optical reflection is
defined to 0 %; b) EQE response curve considering: the optimized rear reflection with front reflection equal to 0 %, the optimized AR
layer and additional CIGS roughness used in the SCAPS model at the front contact and the experimental record cell; c) J-V curve
comparison between the SCAPS model and the experimental result of the champion cell.

4.2 3D electrical Simulations Charge


Lumerical FDTD studies rely on the determination of the optical absorption of each layer, calculated through the FDTD
method, and its dependence on the properties of the incident spectrum. Through the optical absorption rate, it is possible
to calculate the number of generated charge carriers, the 2D and/or 3D distribution of said 𝐺 per each of the incident
wavelengths. To know how the performance of the solar cell will be affected, when the charge transport mechanics are
considered, the architecture needs to be studied under a charge transport simulator 74. However, a purely electronic study
cannot accurately describe a CIGS solar cell on its own. As previously shown by the FDTD analysis, the charge carrier
generation is not homogenous in the CIGS absorber, and it will vary depending on the optical architecture in place. Thus,
the charge carrier generation distribution needs to be transported from the FDTD analysis, into the Lumerical CHARGE,
such that the electronic analysis can take into consideration the optical properties in the density of charge carrier generation.
However, prior to the electrical study with Lumerical CHARGE, the charge carrier generation distribution can provide
information beyond the total solar cell 𝐽𝑆𝐶 value. By integrating the charge carrier distribution through the length and width
of the CIGS absorber, multiplying it by 𝑞 and plotting the resulting 𝐽𝑆𝐶 value in relation to the thickness of the absorber,
the relationship between the generated current and the thickness depth at the absorber can be seen. As an example, a study
was performed about a metallic rear reflector, deposited on the Mo rear contact. In CIGS solar cells, patterned dielectric
layers can be used to encapsulate a highly reflective metal, allowing for both an interfacial passivation effect, as well as
an enhanced optical reflection effect, as explained in Section 3 14,48. Since the 25 nm rear reflector’s metal (Al) had to be
shielded from the CIGS layer, in order to avoid unwanted metallic diffusion into the absorber, it was encapsulated with a
20 nm SiOx film, with the dielectric not only acting as a diffusion barrier, but as a passivation layer as well 7. The previously
described current generation thickness dependence was calculated for the referred architecture, and the results are shown
in Figure 13 a), as well as an illustration of the design in question. In this example, there are two noticeably different

Proc. of SPIE Vol. 11681 1168108-15


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
curves, with the encapsulated sample showing higher total current, in comparison to a reference ultrathin CIGS solar cell.
Furthermore, the current augments are located close to the Mo rear contact, with the current increase due to the optically
enhanced rear reflector and dielectric layer. With this data, the importance of the passivation layer is brought to the
forefront. It shows that the optical gains are located close to the highly recombinative rear interface, meaning that if the
current gains are to be taken advantage off to their fullest potential, the rear contact interface has to be passivated, and
thus, the recombination velocity reduced 74. The Lumerical CHARGE simulations can provide more detail about the solar
cell architectures. For instance, the extraction method of the EQE of a solar cell device can be mimicked and extrapolated
into the Lumerical CHARGE simulations, enabling to attain the EQE plot and charge extraction ratios. The charge carrier
generation distribution can be exported with either the wavelength information integrated into a single distribution, or it
can be exported with it containing the information for every single simulated wavelength value. To calculate the EQE
through Lumerical CHARGE simulations, the latter needs to be performed, and the electrical simulations need to be carried
out for every single wavelength value at short circuit conditions. Afterwards, the current attained through the electrical
simulations is compared with the current values calculated through the FDTD simulations, and the ratio of how successfully
the charge carriers are extracted is calculated. For instance, if through FDTD a current of 0.5 mA/cm 2 is calculated for a
given wavelength value, but through CHARGE the resulting current for the same wavelength is 0.3 mA/cm2, the ratio of
the successfully extracted charge carrier is 0.6. By repeating this process for the entire spectrum, the complete charge
extraction ratio profile can be calculated. Afterwards, the resulting optical absorption calculated through FDTD can be
multiplied by said extraction ratio, and the resulting data is the solar cell’s EQE. An example of the results achieved by the
process described can be seen in Figure 13 b), where an ultrathin reference and a solar cell with back reflector and a
dielectric encapsulation are compared. Through the resulting EQE’s and charge extraction ratios, the effectiveness of the
passivation layer and optical reflector is demonstrated. The former in the higher extraction ratio of the passivated solar
cell, while the latter shows itself in the higher resulting photocurrent of the encapsulated solar cell once again.

0.35 35 100 1.0


Current Density per Height (mA cm-2 nm-1)

a)
Sum of Current Density (mA/cm2)

b)
External Quantum Efficiency (%)

0.30 30
80 0.8
0.25 25

Extraction Ratio
0.20 20 60 0.6
Ultrathin reference
0.15 Encapsulated 15
40 0.4
0.10 10 Ultrathin reference Optical
Ultrathin reference Optical + Electrical
Encapsulated Optical
0.05 5 20 0.2
Encapsulated Optical+Electrical
Ultrathin reference Charge Extraction Ratio
0.00 0 Encapsulated Charge Extraction Ratio
500 400 300 200 100 0
0 0.0
CdS CIGS Height (nm) Mo 400 600 800 1000
Wavelength (nm)

Figure 13. a) Current density generated throughout the thickness of the CIGS layer (dashed lines represent the current density sum);
b) Simulated EQE with and without charge transport mechanics of passivated and reference solar cells. The charge carrier extraction
ratio per wavelength is shown with the respective axis on the right side. Schematic representation of the CIGS solar cells stack
architecture with an encapsulated Al layer.

5. CONCLUSIONS
A showcase of the potential of combining optical and electrical models to efficiently characterize a solar cell stack with
the complex quaternary compound, that is CIGS, is demonstrated. The FDTD Lumerical software enables the realization
of accurate simulations over a great variety of complex and exotic geometries that can provide a detailed insight over light
management architectures in ultrathin CIGS solar cells. A detailed review of several light management approaches capable
of improving the optical performance of ultrathin CIGS solar cell devices, through FDTD optical simulations is also
presented, demonstrating the capability of this approach to accurately simulate and describe the optical benefits of exotic
photonic or plasmonic light management architectures. The realistic light management architectures were simulated in a

Proc. of SPIE Vol. 11681 1168108-16


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
ultrathin solar cell device capable of surpassing the optical performance of a typical thin film (2000 nm) CIGS solar cell,
translated in a 6.15 mA/cm2 enhancement of the short-circuit current over an ultrathin reference device, and 2.26 mA/cm2
over the thin one. Furthermore, the FDTD software can also support and complement electrical simulations performed
through the 1D SCAPS and 3D CHARGE software. The 1D SCAPS simulations supported through optical data extracted
from FDTD optical simulations were validated by the comparison with the champion solar cell holding the record
efficiency for the studied architecture. The potential of the FDTD and 3D CHARGE simulations was also demonstrated,
where a novel approach of simulating the EQE of a solar cell device with a complex rear architecture is demonstrated. This
study shows, through optical and electrical simulations, the potential of the CIGS ultrathin technology to meet the global
electrical demand, as a “green” and sustainable competitor to the current PV market players.

ACKNOWLEDGMENTS

This work was funded in part by the Fundação para a Ciência e a Tecnologia (FCT) under Grants 2020.04564.BD,
IF/00133/2015, PD/BD/142780/2018, SFRH/BD/ 146776/2019, UIDB/04564/2020 and UIDP/04564/2020, and through
the projects NovaCell (PTDC/CTM-CTM/28075/2017), and InovSolarCells (PTDC/FISMAC/29696/2017) co-funded by
FCT and the ERDF through COMPETE2020.

REFERENCES

[1] Nakamura, M., Yamaguchi, K., Kimoto, Y., Yasaki, Y., Kato, T. and Sugimoto, H., "Cd-Free Cu (In,Ga)(Se,S)2
Thin-Film Solar Cell With Record Efficiency of 23.35 %," IEEE J. of Photovoltaics, 9 (6), 1863–1867 (2019).
[2] Wilson, G.M., Al-Jassim, M., Metzger, W.K., Glunz, S.W., Verlinden, P., Xiong, G., Mansfield, L.M., Stanbery,
B.J., Zhu, K., Yan, Y., Berry, J.J., Ptak, A.J., Dimroth, F., Kayes, B.M., Tamboli, A.C., Peibst, R., Catchpole, K.,
Reese, M.O., Klinga, C.S., Denholm, P., Morjaria, M., Deceglie, M.G., Freeman, J.M., Mikofski, M.A., Jordan,
D.C., Tamizhmani, G., and Sulas-Kern, D.B., "The 2020 photovoltaic technologies roadmap. J. Phys. D. Appl.
Phys., 53 (49), 493001 (2020).
[3] Teixeira, J.P., Salomé, P.M.P., Alves, B., Edoff, M., and Leitão, J.P., "Evidence of Limiting Effects of Fluctuating
Potentials on Voc of Cu (In,Ga)Se2 Thin-Film Solar Cells," Phys. Rev. Appl., 11 (5), 054013 (2019).
[4] Lindahl, J., Zimmermann, U., Szaniawski, P., Torndahl, T., Hultqvist, A., Salomé, P., Platzer-Björkman, C., and
Edoff, M., "Inline Cu(In,Ga)Se2 co-evaporation for high-efficiency solar cells and modules," IEEE J.
Photovoltaics, 3 (3), 1100–1105 (2013).
[5] Salomé, P.M.P., Rodriguez-alvarez, H., and Sadewasser, S., " Incorporation of alkali metals in chalcogenide solar
cells," Sol. Energy Mater. Sol. Cells, 143, 9–20 (2015).
[6] Schleussner, S.M., Törndahl, T., Linnarsson, M., Zimmermann, U., Wätjen, T., and Edoff, M., "Development of
gallium gradients in three‐stage Cu (In,Ga)Se2 co‐evaporation processes," Progress in Photovoltaics, 20 (3), 284–
293 (2012).
[7] Cunha, J.M.V., Oliveira, K., Lontchi, J., Lopes, T.S., Curado, M.A., Barbosa, J.R.S., Vinhais, C., Chen, W.-C.,
Borme, J., Fonseca, H., Gaspar, J., Flandre, D., Edoff, M., Silva, A.G., Teixeira, J.P., Fernandes, P.A., and
Salomé, P.M.P., " High‐Performance and Industrially Viable Nanostructured SiOx Layers for Interface
Passivation in Thin Film Solar Cells", Sol. RRL, 2000534 (2021).
[8] Gloeckler, M., Fahrenbruch, A.L., and Sites, J.R. " Numerical modeling of CIGS and CdTe solar cells: setting the
baseline", Proc. 3rd World Conference on Photovoltaic Energy Conversion, 491–494 (2003).
[9]Hasheminassab, S.M.S., and Imanieh, M., "Influence of the Shape and Size of Ag Nanoparticles on the Performance
Enhancement of CIGS Solar Cells: the Role of Surface Plasmons," Plasmonics, 16, 273–282 (2021).
[10]Kovačič, M., Krč, J., Lipovšek, B., Chen, W., Edoff, M., Bolt, P.J., Deelen, J. Van, Zhukova, M., Lontchi, J.,
Flandre, D., Salomé, P., and Topic, M. "Modelling Supported Design of Light Management Structures in Ultra-
thin CIGS Photovoltaic Devices," Journal of Microelectronics, 49 (3), 183–190 (2019).
[11] Lontchi, J., Zhukova, M., Kovacic, M., Krc, J., Chen, W., Edoff, M., Bose, S., Salomé, P.M.P., Goffard, J., Cattoni,
A., Gouillart, L., Collin, S., Gusak, V., and Flandre, D. "Optimization of Back Contact Grid Size in Al2O3-Rear-
Passivated Ultrathin CIGS PV Cells by 2-D Simulations", IEEE J. of Photovoltaics, 10 (6), 1908–1917 (2020).
[12] Wang, Y., Chen, C., Su, T., Yang, T., Liu, W., Cheng, F., Wang, Z.M., and Chueh, Y. "Design of suppressing
optical and recombination losses in ultrathin CuInGaSe2 solar cells by Voronoi nanocavity arrays,". Nano

Proc. of SPIE Vol. 11681 1168108-17


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Energy, 78, 105225 (2020).
[13] Salomé, P.M.P., Vermang, B., Ribeiro-Andrade, R., Teixeira, J.P., Cunha, J.M.V., Mendes, M.J., Haque, S., Borme,
J., Águas, H., Fortunato, E., Martins, R., González, J.C., Leitão, J.P., Fernandes, P.A., Edoff, M., and
Sadewasser, S. "Passivation of Interfaces in Thin Film Solar Cells: Understanding the Effects of a
Nanostructured Rear Point Contact Layer," Adv. Mater. Interfaces, 5 (2), 1–10 (2018).
[14] Lopes, T.S., Cunha, J.M.V., Bose, S., Barbosa, J.R.S., Borme, J., Donzel-Gargand, O., Rocha, C., Silva, R.,
Hultqvist, A., Chen, W.C., Silva, A.G., Edoff, M., Fernandes, P.A., and Salome, P.M.P."Rear Optical Reflection
and Passivation Using a Nanopatterned Metal/Dielectric Structure in Thin-Film Solar Cells". IEEE J.
Photovoltaics, 9 (5), 1421-1427 (2019).

[15] Rahman, T., Fobelets, K., and Fobelets, K. "Efficient tool flow for 3D Photovoltaic modelling", Computer Physics
Communications, 193, 124-130 (2015).
[16] Sayeed, A., and Rouf, H.K. "Numerical Simulation of Thin Film Solar Cell Using SCAPS-1D : ZnSe as Window
Layer". Proc. of 2019 22nd Int. Conf. Comput. Inf. Technol., (1–5 (2019).
[17] Pettersson, J., Zimmermann, U., and Edoff, M. "Baseline model of graded-absorber Cu(In,Ga)Se2 solar cells applied
to cells with Zn1−xMgxO buffer layers," Thin Solid Films, 519 (21), 7476–7480 (2011).
[18] Burgelman, M., Nollet, P., and Degrave, S., "Modelling polycrystalline semiconductor solar cells,", Thin Solid
Films, 361-362, 527–532 (2000).
[19] Ansys/Lumerical, [Lumerical FDTD Solutions Reference Manual] (2014).

[20] "Finite Difference Time Domain (FDTD) solver introduction", https://support.lumerical.com/hc/en-


us/articles/360034914633-Finite-Difference-Time-Domain-FDTD-solver-introduction (06 February 2021).
[21] Carron, R., Avancini, E., Feurer, T., Bissig, B., Losio, P.A., Figi, R., Schreiner, C., Bürki, M., Bourgeois, E.,
Remes, Z., Nesladek, M., Buecheler, S., and Tiwari, A.N. "Refractive indices of layers and optical simulations of
Cu(In,Ga)Se2 solar cells". Sci. Technol. Adv. Mater., 19 (1), 396–410 (2018).
[22] Beal, A. R. and Hughes, H. P. "Kramers-Kronig analysis of the reflectivity spectra of 2H-MoS2, 2H-MoSe2 and 2H-
MoTe2,"J. Phys. C Solid State Phys., 12, 881–890 (1979).
[23] Werner, W.S.M., Glantschnig, K., and Ambrosch-Draxl, C. "Optical constants and inelastic electron-scattering data
for 17 elemental metals," J. Phys. Chem. Ref. Data, 38 (4), 1013–1092 (2009).
[24] Mendes, M.J., Haque, S., Sanchez-Sobrado, O., Araújo, A., Águas, H., Fortunato, E., and Martins, R. "Optimal-
Enhanced Solar Cell Ultra-thinning with Broadband Nanophotonic Light Capture," iScience, 3, 238–254 (2018).
[25] Schmid, M. "Review on light management by nanostructures in chalcopyrite solar cells," Semicond. Sci. Technol.,
32 (4), 043003 (2017).
[26] Oliva, A. I., Solís-Canto, O., Castro-Rodríguez, R. and Quintana, P. "Formation of the band gap energy on CdS thin
films growth by two different techniques," Thin Solid Films, 391 (1), 28–35 (2001).
[27] Kapilashrami, M., Kronawitter, C.X., Törndahl, T., Lindahl, J., Hultqvist, A., Wang, W.C., Chang, C.L., Mao, S.S.,
and Guo, J. "Soft X-ray characterization of Zn1-xSnxOy electronic structure for thin film photovoltaics," Phys.
Chem. Chem. Phys., 14 (29), 10154–10159 (2012).
[28] Witte, W., Spiering, S., and Hariskos, D. "Substitution of the CdS buffer layer in CIGS thin-film solar cells: Status
of current research and record cell efficiencies," Vak. Forsch. und Prax., 26 (1), 23–27 (2014).
[29] Shockley, W. and Queisser, H. J. "Detailed balance limit of efficiency of p-n junction solar cells," J. Appl. Phys.,
32, 510–519 (1961).
[30] Sun, C., and Wang, X. "Efficient Light Trapping Structures of Thin Film Silicon Solar Cells Based on Silver
Nanoparticle Arrays," Plasmonics, 10 (6), 1307–1314 (2015).
[31] Lee, S., Park, J., Yi, J., and Jeong, C. "Improvement of a-Si:H Thin-film Solar Cells by Employing Large Ag
Nanoparticles" Isr. J. Chem., 55(10), 1103–1108 (2015).
[32] Liu, D., and Wang, Q., "Light-trapping surface coating with concave arrays for efficiency enhancement in
amorphous silicon thin-film solar cells," Opt. Commun., 420, 84–89 (2018).
[33] Yu, Z., Raman, A., and Fan, S. "Fundamental limit of nanophotonic light trapping in solar cells," Proc. Natl. Acad.
Sci. U. S. A., 107 (41), 17491–17496 (2010).
[34] Deka, N., and Changmai, P. "A Review on Light Trapping Capacities of Different Photovoltaic Cells" ABDU
Journal of Electrical and Electronical Engineering, 1 (2), 18–28 (2017).
[35] Kang, G., Yoo, J., Ahn, J., and Kim, K. "Transparent dielectric nanostructures for efficient light management in
optoelectronic applications," Nano Today, 10 (1), 22–47 (2015).

Proc. of SPIE Vol. 11681 1168108-18


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
[36] Hu, Y.-H., Chen, Y.-C., Xu, H.-J., Gao, H., Jiang, W.-H., Hu, F., and Wang, Y.-X. "Texture ZnO Thin-Films and
their Application as Front Electrode in Solar Cells," Engineering, 02 (12), 973–978 (2010).
[37] Owen, J.I., Hüpkes, J., Zhu, H., Bunte, E., and Pust, S.E. "Novel etch process to tune crater size on magnetron
sputtered ZnO:Al," Phys. Status Solidi Appl. Mater. Sci., 208 (1), 109–113 (2011).
[38] Seo, B.H., Lee, S.H., Seo, J.H., Jeon, J.H., and Choe, H.H. "Study on the wet etch behavior of a zinc-oxide
semiconductor in acid solutions," J. Korean Phys. Soc., 53 (1), 402–405 (2008).
[39] Hüpkes, J., Owen, J.I., Pust, S.E., and Bunteâ "Chemical etching of zinc oxide for thin-film silicon solar cells,"
ChemPhysChem, 13 (1), 66–73 (2012).
[40] Payne, D.N.R., Boden, S.A., Clark, O.D., Delahoy, A.E., and Bagnall, D.M. "Characterization of experimental
textured ZnO:Al films for thin film solar cell applications and comparison with commercial and plasmonic
alternativesc," Proc. of Conf. Rec. IEEE Photovolt. Spec. Conf., 1560–1564 (2010).
[41] Yan, X., Venkataraj, S., and Aberle, A.G., "Wet-chemical surface texturing of sputter-deposited ZnO:Al films as
front electrode for thin-film silicon solar cells." Int. J. Photoenergy, 2015, 548984 (2015).
[42] Müller, J., Schöpe, G., Kluth, O., Rech, B., Ruske, M., Trube, J., Szyszka, B., Höing, T., Jiang, X., and Brauer, G.
"Texture-etched zinc oxide substrates for silicon thin film solar cells-from laboratory size to large areas," Proc of
Conf. Rec. IEEE Photovolt. Spec. Conf., 758–761 (2000).
[43] Cho, C., Kim, H., Jeong, S., Baek, S.W., Seo, J.W., Han, D., Kim, K., Park, Y., Yoo, S., and Lee, J.Y. "Random and
V-groove texturing for efficient light trapping in organic photovoltaic cell,". Sol. Energy Mater. Sol. Cells, 115,
36–41 (2013).
[44] Amalathas, A.P., and Alkaisi, M.M. "Nanostructures for light trapping in thin film solar cells," Micromachines, 10
(9), 619 (2019).
[45] Čampa, A., Krč, J., Malmström, J., Edoff, M., Smole, F., and Topič, M. "The potential of textured front ZnO and
flat TCO/metal back contact to improve optical absorption in thin Cu(In,Ga)Se2 solar cells," Thin Solid Films,
515, 5968–5972 (2007).
[46] Dahan, N., Jehl, Z., Hildebrandt, T., Greffet, J.J., Guillemoles, J.F., Lincot, D., and Naghavi, N. "Optical approaches
to improve the photocurrent generation in Cu(In,Ga)Se2 solar cells with absorber thicknesses down to 0.5 μm," J.
Appl. Phys., 112 (9), 094902 (2012).
[47] Bharat Bhushan, [Encyclopedia of Nanotechnology], Springer, Netherlands (2007).

[48] Oliveira, A.J.N., de Wild, J., Oliveira, K., Valença, B.A., Teixeira, J.P., Guerreiro, J.R.L., Abalde-Cela, S., Lopes,
T.S., Ribeiro, R.M., Cunha, J.M.V., Curado, M.A., Monteiro, M., Violas, A., Silva, A.G., Prado, M., Fernandes,
P.A., Vermang, B., and Salomé, P.M.P. "Encapsulation of Nanostructures in a Dielectric Matrix Providing
Optical Enhancement in Ultrathin Solar Cells," Sol. RRL, 4 (11), 2000310..
[49] Ong, K.H., Agileswari, R., Maniscalco, B., Arnou, P., Kumar, C.C., Bowers, J.W., and Marsadek, M.B. "Review on
Substrate and Molybdenum Back Contact in CIGS Thin Film Solar Cell," International Journal of Photoenergy,
2018, 9106269 (2018).
[50] Feurer, T., Reinhard, P., Avancini, E., Bissig, B., Löckinger, J., Fuchs, P., Carron, R., Weiss, T.P., Perrenoud, J.,
Stutterheim, S., Buecheler, S., and Tiwari, A.N. "Progress in thin film CIGS photovoltaics – Research and
development, manufacturing, and applications," Prog. Photovoltaics Res. Appl., 25 (7), 645–667 (2018).
[51] Scheer, R., and Schock, H.-W., [Chalcogenide Photovoltaics Physics, Technologies, and Thin Film Devices] Wiley-
VCH, Berlin, 2011.
[52] P. M. P. Salomé, "Chalcogenide Thin Films for Solar Cells: Growth and Properties", Doctoral Dissertation,
Universidade de Aveiro, 2011.
[53] Mendes, M.J., Morawiec, S., and Simone, F. "Colloidal plasmonic back reflectors for light trapping in solar cells,"
Nanoscale, 2014, 6, 4796-4805 (2014).
[54] Bose, S., Cunha, J.M. V., Suresh, S., De Wild, J., Lopes, T.S., Barbosa, J.R.S., Silva, R., Borme, J., Fernandes,
P.A., Vermang, B., and Salomé, P.M.P. "Optical Lithography Patterning of SiO2 Layers for Interface Passivation
of Thin Film Solar Cells," Sol. RRL, 2 (12), 1800212 (2018).
[55] Huang, Y., Sahraei, N., Widenborg, P.I., Marius, I., Kumar, G., Iskander, A., and Aberle, A.G. "Enhanced light
trapping in polycrystalline silicon thin-film solar cells using plasma-etched submicron textures," Sol. Energy
Mater. Sol. Cells, 122, 146–151 (2014).

[56] Morawiec, S., Mendes, M.J., Filonovich, S.A., Mateus, T., Mirabella, S., Águas, H., Ferreira, I., Simone, F.,
Fortunato, E., Martins, R., Priolo, F., and Crupi, I. "Broadband photocurrent enhancement in a-Si:H solar cells

Proc. of SPIE Vol. 11681 1168108-19


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
with plasmonic back reflectors," Opt. Express, 22 (S4), A1059 (2014).
[57] Garcia, M.A. "Surface plasmons in metallic nanoparticles : fundamentals," Journal of Physics D: Applied Physics,
Volume 44 (28), 283001 (2011).
[58] Shafiqaa, A.R., Azizb, A.A., and Mehrdel, B. "Nanoparticle Optical Properties : Size Dependence of a Single Gold
Spherical," J. Phys. Conf. Ser., 1083, 012040 (2018).
[59] Morawiec, S., Mendes, M.J., Filonovich, S.A., Mateus, T., Mirabella, S., Águas, H., Ferreira, I., Simone, F.,
Fortunato, E., Martins, R., Priolo, F., and Crupi, I. " Broadband photocurrent enhancement in a-Si:H solar cells
with plasmonic back reflectors," Opt. Express, 22 (S4), A1059 (2014).
[60] Harrison, R. "Mechanisms and applications of near-field and far-field enhancement using plasmonic nanoparticles",
Doctoral Dissetation, University of Texas, 2012.
[61] Akimov, Y.A., Koh, W.S., Sian, S.Y., Ren, S., Akimov, Y.A., Koh, W.S., Sian, S.Y., and Ren, S. "Nanoparticle-
enhanced thin film solar cells : Metallic or dielectric nanoparticles?" Appl. Phys. Lett. 96 (7), 073111 (2010).
[62] Bose, S., Cunha, J.M.V., Borme, J., Chen, W.C., Nilsson, N.S., Teixeira, J.P., Gaspar, J., Leitão, J.P., Edoff, M.,
Fernandes, P.A., and Salomé, P.M.P. "A morphological and electronic study of ultrathin rear passivated
Cu(In,Ga)Se2 solar cells," Thin Solid Films, 671 (1), 77–84 (2017).
[63] Ledinek, D., Salomé, P., Hägglund, C., Zimmermann, U., and Edoff, M. "Rear Contact Passivation for High
Bandgap Cu(In, Ga)Se2 Solar Cells with a Flat Ga profile," IEEE J. Photovoltaics, 8 (3), 864–870 (2018).
[64] Curado, M.A., Teixeira, J.P., Monteiro, M., Ribeiro, E.F.M., Vilão, R.C., Alberto, H. V., Cunha, J.M.V., Lopes,
T.S., Oliveira, K., Donzel-Gargand, O., Hultqvist, A., Calderon, S., Barreiros, M.A., Chiappim, W., Leitão, J.P.,
Silva, A.G., Prokscha, T., Vinhais, C., Fernandes, P.A., and Salomé, P.M.P. "Front passivation of Cu(In,Ga)Se2
solar cells using Al2O3: Culprits and benefits," Appl. Mater. Today, 21, 100867 (2020).
[65] Cunha, J.M. V., Lopes, T.S., Bose, S., Hultqvist, A., Chen, W.-C., Donzel-Gargand, O., Ribeiro, R.M., Oliveira,
A.J.N., Edoff, M., Fernandes, P.A., and Salome, P.M.P., "Decoupling of Optical and Electrical Properties of
Rear Contact CIGS Solar Cells," IEEE J. Photovoltaics, 9 (6), 1857–1862 (2019)
[66] Cunha, J.M.V., Fernandes, P.A., Hultqvist, A., Teixeira, J.P., Bose, S., Vermang, B., Garud, S., Buldu, D., Gaspar,
J., Edoff, M., Leitao, J.P., and Salome, P.M.P., "Insulator Materials for Interface Passivation of Cu(In,Ga)Se2
Thin Films," IEEE J. Photovoltaics, 8 (5), 1313–1319 (2018).
[67] Lare, C. Van, Yin, G., Polman, A., and Schmid, M. "Light Coupling and Trapping in Ultrathin Cu(In,Ga)Se2 Solar
Cells Using Dielectric Scattering Patterns," ACS Nano, 9, 10, 9603–9613 (2015).
[68] Yin, G., Manley, P., and Schmid, M. "Light trapping in ultrathin CuIn1-xGaxSe2 solar cells by dielectric
nanoparticles," 163 (15), 443–452 (2018).
[69] Hegedus, S.S., and Shafarman, W.N. "Thin-film solar cells: device measurements and analysis," Prog. Photovoltaics
Res. Appl., 12 (23), 155–176 (2004).

[70] Jackson, P., Wuers, R., Hariskos, D., Lotter, E., Witte, W. and Powalla M., "Effects of heavy alkali elements in
Cu(In,Ga)Se2 solar cells with efficiencies up to 22.6%," Phys. Status Solidi - Rapid Res. Lett., 10 (8), 583–586
(2016).
[71] Decock, K., Khelifi, S., and Burgelman, M. "Modelling multivalent defects in thin film solar cells," Thin Solid
Films, 519 (21), 7481–7484 (2011).
[72] Burgelman, M., Decock, K., Khelifi, S., and Abass, A. "Advanced electrical simulation of thin film solar cells,"
Thin Solid Films, 535 (1), 296–301 (2013).
[73] Minoura, S., Kodera, K., Maekawa, T., Miyazaki, K., Niki, S., and Fujiwara, H."Dielectric function of Cu(In,
Ga)Se2-based polycrystalline materials," J. Appl. Phys., 113 (6), 063505 (2013).
[74] "CHARGE solver introduction – Lumerical Support", https://support.lumerical.com/hc/en-
us/articles/360034917693-CHARGE-solver-introduction, (06 February 2021).

Proc. of SPIE Vol. 11681 1168108-20


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 12 Jan 2024
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use

You might also like