PCS D 25 02089 Reviewer
PCS D 25 02089 Reviewer
Design and numerical simulation of MASnBr3 (CH3NH3SnBr3) based solar cell using
SCAPS-1D simulation software
--Manuscript Draft--
Abstract: Perovskite solar cells (PSCs) have emerged as a leading research focus in the field of
photovoltaic (PV) technology due to their remarkable power conversion efficiency
(PCE) and low- fabrication cost. However, the commercialization of PSCs has been
hindered by critical challenges such as lead (Pb) toxicity, long-term stability issues, and
environmental concerns. In this study, a lead-free perovskite material (CH3NH3SnBr3)
was employed as an alternative and a novel device structure was simulated using
SCAPS-1D to achieve high PCE. The implementation of Cu2O as the hole transport
layer (HTL), Sulfur-doped tin oxide (STO) as the electron transport layer (ETL),
Fluorine-doped tin oxide (FTO) as the front contact, and Au as back contact defines the
proposed device structure as follows: FTO/STO/CH3NH3SnBr3/Cu2O/Au. Various
device parameters, including layer thickness, defect density, doping concentration, and
series–shunt resistance, were optimized to achieve the best performance. The
optimized structure exhibits an impressive PCE of 35.19%, a high fill factor (FF) of
89.46%, an open circuit voltage (Voc) of 1.1676V, and a short-circuit current density
(Jsc) of 33.689mA/[Link] device exhibits its peak quantum efficiency (QE) of 98.8%
within the wavelength range of 390–400nm. This study, therefore, proposes an eco-
friendly model with remarkable power conversion efficiency, offering valuable insights
for the advancement of clean and sustainable energy technologies.
Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Manuscript File Click here to view linked References
Design and numerical simulation of MASnBr3 (CH3NH3SnBr3) based solar cell using
*Corresponding Author
Email: 06habib05@[Link]
Abstract:
Perovskite solar cells (PSCs) have emerged as a leading research focus in the field of photovoltaic
(PV) technology due to their remarkable power conversion efficiency (PCE) and low- fabrication
cost. However, the commercialization of PSCs has been hindered by critical challenges such as
lead (Pb) toxicity, long-term stability issues, and environmental concerns. In this study, a lead-free
perovskite material (CH3NH3SnBr3) was employed as an alternative and a novel device structure
was simulated using SCAPS-1D to achieve high PCE. The implementation of Cu2O as the hole
transport layer (HTL), Sulfur-doped tin oxide (STO) as the electron transport layer (ETL),
Fluorine-doped tin oxide (FTO) as the front contact, and Au as back contact defines the proposed
including layer thickness, defect density, doping concentration, and series–shunt resistance, were
optimized to achieve the best performance. The optimized structure exhibits an impressive PCE of
35.19%, a high fill factor (FF) of 89.46%, an open circuit voltage (Voc) of 1.1676 V, and a short-
circuit current density (Jsc) of 33.689 mA/[Link] device exhibits its peak quantum efficiency
(QE) of 98.8% within the wavelength range of 390–400 nm. This study, therefore, proposes an
eco-friendly model with remarkable power conversion efficiency, offering valuable insights for
1|Page
1. Introduction:
Perovskite-structured solar cells continue to attract significant interest in the solar energy
community due to their high-power conversion efficiency (PCE) and excellent electrical and
optical properties [1]. Moreover, perovskite solar cells (PSC) offer several advantages, including
low production cost, a simple fabrication process, tunable bandgap, high charge carrier mobility,
and compatibility with flexible substrates [2]. These devices operate based on the photovoltaic
(PV) effect, where absorbed sunlight generates electron-hole pairs that are separated and collected
to produce electrical power. The device employs an absorber layer sandwiched between a hole
transport layer (HTL) and an electron transport layer (ETL) on either side. Additionally, the front
and back contacts serve as electrical terminals that extract and deliver the photogenerated charge
carriers, enabling current flow through the external load. The impressive values of open circuit
voltage (Voc), short circuit current density (Jsc), fill factor (FF) and PCE in PSCs make them highly
attractive for future commercial deployment [9,12]. PSC derive their name from the ABX3 crystal
lattice of the light-absorbing layer, where 'A' and 'B' represent cations and 'X' denotes an anion,
typically a halide. Perovskite formation generally occurs when the A-site cation has an ionic radius
between 1.60 Å and 2.50 Å. Among the various compositions explored, methylammonium lead
halides (CH3NH3PbX3, with X = I⁻, Br⁻, or Cl⁻) have been extensively studied due to their tunable
bandgap ranging from approximately 1.55 to 2.3 eV, depending on the specific halide used.
narrower bandgap spectrum of 1.48–2.2 eV. Its lower minimum bandgap is closer to the theoretical
optimum for single-junction solar cells, making it a strong contender for achieving superior power
conversion efficiencies [7,8]. However, the presence of toxic lead compounds remains a significant
challenge for the widespread adoption of PSCs [10,13]. This concern calls for the development of
2|Page
lead-free perovskite alternatives that maintain high efficiency while ensuring environmental safety
and sustainability [11,14]. Among the promising candidates, tin-based perovskites such as
methylammonium tin halide (MASnX3) have attracted considerable attention as a viable solution,
perovskites, mixing different halide ions can significantly enhance their optoelectronic
performance. Hao and colleagues pioneered the partial substitution of iodide (I) with bromide (Br)
parameters from MASnI3 to MASnBr3, which in turn increases the material’s bandgap [15]. S.
Mushtaq et. al., optimizes a lead-free MASnBr3 perovskite solar cell using numerical simulations,
achieving a high PCE of 34.52%, along with Voc of 1.1214 V, Jsc of 34.8654 mA/cm2, and f of
88.30%. It highlights the potential of MASnBr3 as a stable, efficient, and environmentally friendly
alternative to lead-based PSC [3]. Similarly, A. S. Dungani et. al., optimizes a tin-based MASnBr3
perovskite solar cell using SCAPS-1D simulations, achieving a PCE of 24.98% [4]. M. Shah et.
al., analyzed MASnBr3-based lead-free PSCs, achieving a 22.71% PCE. Results highlight its
strong optoelectronic properties and emphasize the impact of temperature, resistances, and defect
density on device performance and optimization [5]. Additionally, S. Mushtaq et. al., designed a
lead-free, HTL-free MASnBr3-based perovskite solar cell using SCAPS-1D, achieving a PCE of
27.37%. It investigates the effects of thickness, doping, and defect densities, highlighting
MASnBr3/CdS interface properties on overall device performance [6]. Though extensive research
has been conducted on MASnBr3 (CH3NH3SnBr3)-based perovskite solar cells, there remains
significant potential to further enhance their performance through continued optimization and
innovation.
3|Page
This study presents a novel design of a CH3NH3SnBr3-based perovskite solar cell, utilizing
CH3NH3SnBr3 as the primary absorber layer. The choice of CH3NH3SnBr3 as the absorber layer is
motivated by its non-toxic nature, suitable bandgap, excellent optical absorption, and promising
Cu2O was employed as the HTL in this work. The selection of Cu2O over other materials is due to
its relatively low cost, high absorption coefficient, and intrinsic high hole mobility. Additionally,
its energy levels align well with the absorber layer (MASnI3), and it exhibits excellent
photochemical and thermal stability, as well as long-term stability in air [18]. Sulfur-doped tin
oxide (STO) was used as ETL that offers improved electrical conductivity and enhanced charge
extraction, leading to better overall device performance [19]. The proposed device structure is
defined by using FTO as the front contact and gold (Au) as the back contact, arranged as follows:
doping concentration of HTL and ETL, as well as the absorber layer thickness, acceptor and donor
doping concentrations in the absorption layer, defect density of the perovskite layer, interface
defect densities at the HTL/perovskite, perovskite/ETL interfaces, and the series and shunt
resistances. The optimized structure achieved an outstanding PCE of 35.19%, FF of 89.46%, Voc
of 1.1676 V, and Jsc of 33.689 mA/cm2. This comprehensive optimization highlights the potential
The Solar Cell Capacitance Simulator—1D (SCAPS-1D), version 3.3.12, is a widely recognized
simulation tool within the PV research community, and it is utilized in this study to analyze the
4|Page
performance of the proposed solar cell model. The simulation is governed by Poisson’s equation
along with the electron and hole continuity equations, under standard test conditions (STC) with
approach, allowing simulations to be performed repeatedly over a wide range of material and
physical parameters. This capability enhances the reliability and efficiency of the optimization
process. Key factors such as doping concentration profiles, interfacial energy alignments, and
been incorporated to evaluate the overall performance of the designed solar cell.
SCAPS-1D determines key output parameters—such as Voc, Jsc, FF, and PCE—by numerically
solving a set of fundamental differential equations, including Poisson’s equation and the continuity
equations for electrons and holes. Poisson’s equation describes the spatial distribution of
electrostatic potential within the device, which directly affects the internal electric field and charge
distribution. This field plays a crucial role in governing charge carrier transport, separation, and
recombination dynamics. Accurately modeling the potential profile is essential for optimizing
charge collection efficiency and enhancing the overall performance of the solar cell.
𝑑2 Ψ(𝑥) q
=ϵ [p(x) – n(x) + ND – NA + ρp – ρn ] (1)
𝑑𝑥 2 o ϵr
The continuity equations for holes (Eq. 2) and electrons (Eq. 3) govern the dynamic behavior of
charge carriers in semiconductors by expressing the balance between their generation, transport,
and recombination processes. These equations are crucial for simulating the internal physics of a
solar cell, as they track how carrier densities evolve over time and space under steady-state or
illuminated conditions. Accurate resolution of these continuity relations allows for precise
5|Page
estimation of photocurrent, voltage characteristics, and the influence of material properties on
device behavior. Incorporating these equations into numerical models, such as those used in
SCAPS-1D, is vital for predicting performance metrics like Voc, Jsc, FF, and PCE, and for guiding
the design of high-efficiency solar cells through material and structural optimization.
𝛿𝑝 1 𝛿𝐽𝑝
=𝑞 + Gp - Rp (2)
𝛿𝑡 𝛿𝑥
𝛿𝑛 1 𝛿𝐽𝑛
= + Gn – Rn (3)
𝛿𝑡 𝑞 𝛿𝑥
Equations (4) and (5) represent the drift-diffusion model, which describes the transport
accounts for two fundamental driving forces: drift, due to the influence of internal electric fields,
and diffusion, arising from spatial gradients in carrier concentration. It plays a pivotal role in
simulating the behavior of solar cells, as it directly impacts charge separation, recombination rates,
and the generation of photocurrent. By incorporating both electric field effects and carrier mobility,
the drift-diffusion framework provides a comprehensive basis for evaluating device performance,
predicting efficiency outcomes, and informing design strategies for improved PV operation.
𝑑𝑛(𝑥)
Jn(x) = qµnn(𝑥)𝜀(𝑥) + qDn (4)
𝑑𝑥
𝑑𝑝(𝑥)
Jn(x) = qµpp(𝑥)𝜀(𝑥) + qDp (5)
𝑑𝑥
models this phenomenon using the Shockley-Read-Hall (SRH) recombination formalism, which
is integrated into the continuity equations alongside the drift-diffusion model. The SRH approach
provides a realistic representation of non-radiative carrier losses by accounting for trap states
6|Page
within the bandgap that capture and release charge carriers. This detailed treatment allows for in-
factors. By accurately simulating these recombination processes, the SRH model supports the
targeted optimization of material quality and device structure to enhance overall efficiency.
PV performance is evaluated using key metrics including the Voc, Jsc, FF, and PCE. Together, these
parameters reflect the balance between charge generation, recombination losses, and the device’s
ability to convert sunlight into electrical energy. Analyzing these factors offers valuable insights
into the limitations and optimization potential of solar cells, guiding improvements in material
Voc represents the highest voltage a solar cell can produce when no external current flows. It is
primarily influenced by the absorber layer’s bandgap energy and losses due to recombination at
interfaces. The magnitude of Voc can be calculated using the following equation.
mkB T Jsc
Voc = ln(1+ ) (6)
q Jo
A higher Voc value typically indicates reduced non-radiative recombination within the device.
Jsc represents the maximum photocurrent generated per unit area (mA/cm²) when the solar cell
terminals are shorted (zero applied voltage). It depends on the efficiency of light absorption and
the effectiveness of charge carrier collection. Jsc can be calculated using Equation (7).
A higher Jsc value signifies better photon harvesting and more efficient extraction of charge
carriers.
7|Page
FF quantifies the squareness of the current-voltage (J-V) curve, reflecting the solar cell’s ability to
efficiently convert generated charge carriers into usable electrical power by balancing voltage and
current output. Physically, FF is constrained by factors such as series resistance and recombination
losses within the device. It is defined as the ratio of the maximum achievable power output to the
product of Voc and Jsc, serving as a key indicator of charge extraction efficiency.
A higher FF —typically above 80%—signals reduced resistive losses and well-optimized material
interfaces.
PCE measures the proportion of incident solar energy that a solar cell successfully converts into
usable electrical power, serving as the primary metric of the device’s energy conversion
effectiveness.
The theoretical maximum efficiency is constrained by the Shockley-Queisser limit, which defines
the upper bound for a single-junction cell based on its bandgap energy. Achieving PCE values
close to this limit requires minimizing losses due to recombination, resistive effects, and
The schematic layout of the proposed perovskite solar cell (PSC) is depicted in Fig. 1(a). To avoid
the toxicity concerns associated with lead-based materials, CH3NH3SnBr3 was chosen as a non-
8|Page
toxic, lead-free absorber layer. Upon exposure to sunlight, the perovskite layer absorbs incoming
photons, resulting in the generation of electron–hole pairs. These photogenerated carriers are then
separated and guided by adjacent transport layers—electrons move through the ETL, and holes
through the HTL—toward their respective electrodes, facilitating effective charge extraction. The
success of this charge transport process largely depends on the precise alignment of energy levels
among the absorber, ETL, and HTL, which helps reduce recombination losses. In this design,
sulfur-doped tin oxide (STO) functions as the ETL, and Cu2O serves as the HTL; both materials
exhibit energy band positions well-matched with CH3NH3SnBr3, as illustrated in Fig. 1(b). The
structure incorporates FTO coated on a glass substrate as the transparent front contact, while a gold
Fig. 1. Solar cell (a)Schematic diagram, and (b) Energy band diagram
The performance of a PSC is largely determined by the physical dimensions and optoelectronic
characteristics of its individual layers—including the absorber, charge transport layers, and
electrodes. Key factors such as bandgap energy, absorption coefficient, charge carrier mobility,
exciton diffusion length, and defect density play a direct role in shaping the PCE, Voc, Jsc, and FF.
9|Page
Achieving optimal layer thickness is crucial; while a thicker absorber enhances photon absorption,
overly thick transport layers can lead to increased recombination losses. Conversely, layers that
are too thin may limit light absorption or hinder charge collection, thereby reducing J sc or Voc.
Among the optoelectronic parameters, the bandgap is particularly influential in determining the
spectral range of light absorption. A narrower bandgap (<1.7 eV) typically improves J sc due to
better infrared absorption but compromises Voc. In contrast, a wider bandgap (>1.7 eV) enhances
Voc but limits Jsc by reducing the amount of absorbed light. Therefore, fine-tuning the bandgap is
affinity between the perovskite absorber and the ETL reduces potential energy barriers, facilitating
smooth electron flow, lowering interfacial recombination, and improving both Voc and Jsc.
Additional optoelectronic properties also influence PSC operation. The relative dielectric constant
(𝜖r) affects exciton binding energy and charge screening; higher 𝜖r values help suppress
recombination. The effective densities of states in the conduction (Nc) and valence (Nv) bands
impact carrier concentration, which influences Fermi level positioning and charge extraction.
Electron and hole thermal velocities affect injection dynamics at interfaces, while charge carrier
mobilities are central to transport efficiency—higher mobilities reduce resistive losses and boost
FF. Donor (ND) and acceptor (NA) doping concentrations shape conductivity, whereas elevated
defect density (Nt) introduces trap-assisted recombination pathways, degrading both Voc and FF.
Proper tuning and optimization of these parameters are vital for enhancing charge collection and
10 | P a g e
Table 1. Input parameters of FTO, ETL, absorber, HTL
carrier collection and leading to a drop in both Voc and FF. Accounting for these imperfections in
effects. This approach also supports the development of effective passivation techniques and fine-
tunes interfacial design to boost device efficiency. The parameters used for simulating interface-
11 | P a g e
Table 2. Input parameters of interface defect layers
The PV performance of a solar cell is strongly influenced by the thickness and doping
concentration of its constituent layers, particularly the absorber, ETL, and HTL. In this study, the
proposed perovskite solar cell architecture was optimized by systematically adjusting these
parameters to achieve maximum efficiency. Extensive simulations were performed to assess the
individual and combined effects of layer thickness and doping density on device behavior. The
resulting optimized design achieves an effective balance between efficient charge carrier transport
The thickness of the absorber layer plays a pivotal role in determining the PV performance of a
solar cell, as it directly impacts light absorption, charge carrier generation, and overall energy
capture without compromising carrier extraction. Similarly, the thicknesses of the ETL and HTL
are equally critical. Properly optimized ETL and HTL dimensions facilitate efficient charge
12 | P a g e
separation and extraction, minimize recombination losses, and improve carrier mobility. In
contrast, suboptimal layer thickness can lead to increased series resistance, reduced carrier
collection efficiency, and performance degradation, adversely affecting key metrics such as PCE,
FF, Voc and Jsc. Thus, careful calibration of these layers is vital for achieving peak solar cell
performance.
The combined influence of absorber layer thickness (ranging from 0.7 μm to 1.1 μm) and HTL
thickness (ranging from 0.03 μm to 0.07 μm) on the performance of the proposed PSC is illustrated
in Fig. 2. The variation of Voc with absorber and HTL thickness is presented in Fig. 2(a). Voc
exhibits a gradual decline from 1.0104 V at 0.7 μm absorber thickness to 0.9951 V at 1.1 μm. This
reduction is primarily due to increased bulk recombination that outweighs the benefit of enhanced
In terms of Jsc, shown in Fig. 2(b), a steady rise is observed with increasing absorber thickness—
from 33.8500 mA/cm² at 0.7 μm to 34.6540 mA/cm² at 1.1 μm—attributed to improved photon
absorption and carrier generation. Additionally, Jsc increases very slowly as HTL thickness grows
from 0.3 μm to 0.7 μm, owing to better hole extraction and reduced interface recombination.
However, beyond 0.5 μm, the improvement becomes marginal, as further increases contribute little
The FF trends are depicted in Fig. 2(c). FF decreases gradually from 81.94% to 81.49% with
increasing absorber thickness, primarily due to elevated series resistance and recombination within
the bulk, which hinder efficient carrier extraction. On the other hand, FF improves with HTL
thickness up to 0.5 μm, driven by enhanced hole transport and lower interfacial recombination
13 | P a g e
losses. Beyond this point, FF stabilizes, as additional HTL thickness offers no substantial electrical
The PCE of the PSC is highly sensitive to variations in both absorber and HTL thicknesses, as
shown in Fig. 2(d). Initially, increasing the absorber thickness enhances PCE, reaching a peak of
28.17% at 0.9 μm. This improvement is attributed to greater photon absorption and increased
carrier generation. However, further increases beyond 0.9 μm lead to a decline in efficiency due to
intensified bulk recombination and diminished carrier extraction, which offset the optical gains.
In contrast, the HTL thickness exhibits a more limited but still notable effect on PCE. A minimum
HTL thickness of 0.05 μm is necessary to maintain peak device performance. Below this threshold,
14 | P a g e
PCE drops sharply due to insufficient hole extraction and increased interface recombination.
Beyond 0.05 μm, PCE remains nearly constant, as further increases do not contribute meaningfully
to performance and may introduce resistive losses. Additional analysis reveals that above this
threshold, Voc, FF, and PCE stabilize, while Jsc gains become negligible—establishing 0.05 μm as
With these optimized values—0.9 μm for the absorber and 0.05 μm for the HTL—the simulated
PSC achieves a Voc of 1.0016 V, a Jsc of 34.4093 mA/cm², a FF of 81.73%, and a maximum PCE
of 28.17%.
The combined influence of the ETL and front contact- FTO thicknesses on the PV performance of
the proposed device is analyzed in Fig. 3. During this investigation, the absorber and HTL
thicknesses were held constant at their optimized values of 0.9 μm and 0.05 μm, respectively. The
ETL thickness was systematically varied from 0.18 μm to 0.22 μm, while the FTO layer thickness
ranged from 0.03 μm to 0.06 μm to evaluate and optimize their combined effects.
Fig. 3(a) depicts the effect of ETL and FTO thickness on the Voc. Initially, Voc is 1.0002 V at an
ETL thickness of 0.18 μm but slightly increase with thickness, reaching a stable value beyond
0.2 μm. This trend results from initial changes in the internal electric field and carrier transport
efficiency, while further increases in thickness no longer impact charge dynamics significantly. In
contrast, the thickness of the FTO layer has no noticeable effect on Voc.
Jsc increases from 34.3825 mA/cm2 at 0.18 μm and increases with ETL thickness, shown in Fig.
3(b), due to enhanced electron extraction and reduced interfacial recombination, which improve
15 | P a g e
carrier collection efficiency. In contrast, Jsc initially increases with FTO thickness—from
34.4231 mA/cm² at 0.03 μm to a peak of 34.4058 mA/cm² at 0.04 μm. Beyond this, Jsc starts to
decrease due to increased optical absorption and resistive losses in the thicker FTO layer.
Fig. 3(c) illustrates the FF response. FF increases with ETL thickness, reaching a peak of 84.24%
at 0.2 μm due to reduced recombination and enhanced electron mobility. Beyond this thickness,
FF becomes saturated, indicating that further increases in ETL thickness offer no additional
thickness, rising from 84.12% at 0.03 μm, due to enhanced electrical conductivity and improved
16 | P a g e
Fig. 3(d) illustrates the variation of PCE as a function of ETL and FTO layer thicknesses. As the
28.98% at 0.2 μm. Beyond this point, the efficiency plateaus, suggesting that the charge transport
and recombination processes are optimized, and further thickness increases no longer contribute
to performance enhancement.
For the FTO front contact, the PCE also peaks at 28.98% when the thickness is set to 0.04 μm.
This thickness provides an effective compromise between sufficient electrical conductivity and
adequate optical transparency, ensuring optimal light transmission and minimal resistive losses.
Based on these findings, the optimal thicknesses are determined to be 0.2 μm for the ETL and
0.04 μm for the FTO layer, as these configurations yield the highest overall PCE for the proposed
device structure.
The doping concentrations within each layer of CH3NH3SnBr3-based PSCs critically influence
their PV performance. These doping levels directly affect key processes such as charge carrier
generation, transport dynamics, and recombination rates. To further optimize device efficiency, a
Fig. 4 presents the effect of varying doping densities in both the perovskite absorber HTL on
essential PV output parameters, including Voc, Jsc, FF, and PCE. In parallel, Fig. 5 illustrates the
17 | P a g e
influence of doping concentration variations in the ETL and the FTO front contact on the same
performance metrics.
Variations in the doping concentration of the HTL exhibit minimal influence on Voc, as shown in
Fig. 4(a), indicating that the Voc is predominantly governed by the electronic properties of the
CH3NH3SnBr3 absorber layer in this device architecture. Similarly, changes in HTL doping levels
have a negligible effect on Jsc, suggesting that photocurrent generation is largely controlled by the
2×10¹⁵ cm-3, as depicted in Fig. 4(a). This enhancement is attributed to improved hole transport
and reduced interfacial recombination losses. However, beyond this threshold, FF becomes
Likewise, PCE increases with HTL doping up to 2×1015 cm-3 due to improved carrier extraction
and reduced recombination, but shows negligible change thereafter, reflecting a saturation effect.
Based on these observations, the optimal doping concentrations are identified as 1019 cm-3 for the
absorber and 2×1015 cm-3 for the HTL, which collectively deliver the highest device efficiency.
18 | P a g e
Fig. 4. Effect acceptor density of HTL and Absorber on PV performance.
Fig. 4(b) shows that Voc exhibits a positive correlation with increasing acceptor doping
concentration in the CH3NH3SnBr3 absorber layer, reaching a peak value of 1.1685 V at 1019 cm-
3
. This enhancement is attributed to a stronger built-in electric field and improved separation of
The same figure also illustrates the variation in Jsc with absorber doping. Jsc shows a non-
monotonic trend—initially peaking at 33.8601 mA/cm2 for 1017 cm-3 but slightly decreasing
thereafter to 33.5896 mA/cm2 at 1019 cm-3. This decline at higher doping levels is likely due to
19 | P a g e
increased recombination losses and reduced minority carrier diffusion length, which limits charge
collection efficiency.
FF increases progressively with the acceptor doping density in the CH3NH3SnBr3 absorber layer,
rising from 84.21% at 1016 cm-3 to a peak of 88.36% at 1019 cm-3. This improvement is attributed
to reduced series resistance and enhanced charge extraction efficiency, facilitated by stronger built-
in electric fields and lower recombination rates. Similarly, PCE exhibits a significant rise from
28.98% to 34.68% across the same doping range, as shown in Fig. 4(d). The concurrent
improvement in FF and PCE underscores the critical role of optimized absorber doping in
maximizing solar cell performance through improved charge transport and reduced non-radiative
losses.
The PV performance parameters (Voc, Jsc, FF, and PCE) - exhibit significant sensitivity to
variations in the donor doping concentrations of ETL and front contact layer (FTO), underscoring
the importance of doping optimization in high-performance PSCs. To explore this impact, the
donor concentrations were systematically varied from 2×1016 cm-3 to 2×1019 cm-3 for the ETL and
from 1013 cm-3 to 1016 cm-3 for the FTO layer, with the resulting device characteristics presented
As shown in Fig. 5(a), Voc shows a slight decline from 1.1685 V to 1.1665 V as the ETL donor
density increases from 2×1016 to 2×1019 cm-3, indicating that the Voc is relatively insensitive to
ETL doping. This minimal change occurs because Voc is predominantly determined by the
absorber layer’s energy levels rather than the transport layer properties. In contrast, J sc increases
20 | P a g e
steadily from 33.5896 mA/cm2 to 33.7118 mA/cm2 due to enhanced electron mobility and
improved conductivity in the ETL, which reduce series resistance and facilitate more efficient
charge transport. The FF rises from 88.36% to a peak of 89.46% at 2×1018 cm-3, reflecting reduced
energy losses and improved charge collection efficiency. However, a slight drop to 88.97% at
2×1019 cm-3 suggests that excessive doping may increase charge carrier scattering or promote
recombination. Consequently, PCE reaches a maximum of 35.19% at 2×1018 cm-3 before declining
slightly, highlighting that over-doping can negatively affect overall performance. Therefore, an
ETL donor density of 2×1018 cm-3 is identified as optimal for achieving the best balance between
21 | P a g e
The effect of FTO donor doping on PSC performance is summarized in Fig. 5(b). As the donor
concentration increases from 2×1013 to 2×1016 cm-3, Voc remains relatively stable, showing only a
slight decrease from 1.1680 V to 1.1676 V. This indicates that the built-in electric field and junction
quality are not significantly affected within this doping range. Jsc shows a marginal yet consistent
increase, from 33.6878 mA/cm2 to 33.6911 mA/cm2, primarily due to enhanced electron transport
and reduced resistive losses with higher donor density in the ETL. FF also improves noticeably,
from 89.00% at 2×1013 cm-3 to a peak value of 89.46% at 2×1015 cm-3 and remains constant
Consequently, PCE rises from 35.02% to a maximum of 35.19%, after which further doping offers
no significant gain. These results suggest that increasing ETL donor density enhances device
concentration of ≥2×1015 cm-3 is identified as optimal for maximizing PCE without risking
performance degradation.
Defect densities in the absorber, HTL, ETL, FTO layers, and their interfaces critically impact PSC
performance. These defects serve as non-radiative recombination centers, reducing charge carrier
lifetime and mobility. In the absorber, high defect levels increase bulk recombination, lowering J sc.
They also reduce quasi-Fermi level splitting, decreasing Voc, and raise series resistance, which
degrades FF. Interface defects are particularly harmful, hindering charge extraction and enhancing
recombination at key junctions like ETL/absorber. Together, these effects significantly reduce
22 | P a g e
PCE. Minimizing both bulk and interfacial defects is essential for high performance and long-term
device stability.
Defect states in different layers of PSCs significantly impact device performance by acting as
recombination centers that limit carrier lifetimes and extraction efficiency. To systematically
evaluate this influence, the defect density (Nt) was varied across a wide range (from 1013 to
1018 cm-3) in the absorber (CH3NH3SnBr3), HTL (Cu2O), ETL (STO), and front contact (FTO),
Among all layers, the CH3NH3SnBr3 absorber exhibited the most pronounced sensitivity to defect
density. As Nt increased from 1×1013 to 1×1018 cm-3, a substantial degradation was observed in all
key performance metrics. The Voc dropped from 1.2262 V to 0.9565 V, while the Jsc declined
sharply from 34.50 mA/cm2 to 5.22 mA/cm2. FF also decreased from 89.88% to 78.62%, leading
to a dramatic reduction in PCE from 38.02% to 3.93%. This deterioration is attributed to enhanced
Shockley–Read–Hall (SRH) recombination within the absorber, which reduces carrier lifetimes
and increases non-radiative losses. Since the absorber is the primary region for light absorption
and charge generation, defect-induced recombination here has a disproportionately large impact
In contrast, the Cu2O-HTL showed exceptional tolerance to increased defect densities. Across the
entire Nt range from 1013to 1018 cm-3, there were no observable changes in Voc, Jsc, FF, or PCE, all
of which remained stable at 1.1676 V, 33.689 mA/cm2, 89.46%, and 35.19%, respectively. This
23 | P a g e
insensitivity suggests that the Cu2O HTL is less prone to defect-induced recombination or that
The STO, ETL exhibited a mild response to defect density variation. While Voc and FF remained
nearly constant, Jsc decreased slightly from 33.69 mA/cm2 to 33.51 mA/cm2, and the PCE dropped
marginally from 35.19% to 34.99% at 1018 cm-3. This trend indicates that although defects in the
ETL have a comparatively weaker influence than those in the absorber, they can still contribute to
Similarly, defect density variation in the FTO front contact had negligible effects. Across all tested
Nt values, the PV parameters remained nearly unchanged, highlighting the electrical robustness of
Overall, the simulation results strongly emphasize the critical role of the absorber layer’s quality
in determining PSC performance. Minimizing bulk and interfacial defects in the CH3NH3SnBr3
absorber is vital for suppressing recombination, enhancing charge collection, and achieving high
24 | P a g e
Fig. 6. Impact of (a) defect density (Nt) and (b) interface defect density (Nint) on PV performance.
STO/FTO junctions on device performance is shown in Fig. 6(b). The defect density variation
ranges from 1×1010 to 1×1018 cm-3, revealing distinct sensitivity levels across the interfaces.
defect density. Voc decreases from 1.1676 V at 1010 cm-3 to 1.1182 V at 1018 cm-3, while Jsc drops
from 33.69 mA/cm2 to approximately 28.25 mA/cm2. FF declines slightly from 89.46% to 88.81%,
resulting in a PCE reduction from 35.19% to about 28.05%. This moderate degradation indicates
25 | P a g e
that the HTL (Cu2O) interface is somewhat tolerant to higher defect concentrations but still
densities.
The CH3NH3SnBr3/STO interface shows more pronounced sensitivity to defects. Voc declines
sharply from 1.1676 V to 0.8130 V as defect density increases from 1010 to 1018 cm-3.
Correspondingly, Jsc exhibits a steep drop from 33.69 mA/cm2 to nearly zero (0.81 mA/cm2), while
FF decreases from 89.46% to 84.33%. These combined losses cause PCE to fall drastically from
35.19% to 0.55%. This severe performance deterioration highlights the critical role of the
ETL/absorber interface in maintaining efficient charge separation and transport. High defect
densities at this junction significantly increase interface recombination, severely limiting device
efficiency.
In contrast, the STO/FTO interface demonstrates remarkable stability even at elevated defect
densities. Voc remains nearly constant around 1.1675 V, Jsc only slightly decreases from 33.69
mA/cm2 to 33.51 mA/cm2, and FF maintains a stable value close to 89.43%, resulting in minimal
change in PCE (~35.19% to 34.99%). This stability suggests that the back-contact interface is less
Overall, these results emphasize the need for stringent control of interface defects, especially at
the CH3NH3SnBr3/STO ETL/absorber interface, to achieve optimal solar cell performance. While
the Cu2O/absorber interface can tolerate moderate defects, passivation and interface engineering
remain essential. The STO/FTO interface is comparatively robust, but minimizing defects
throughout the device stack will collectively enhance PCE and device longevity.
26 | P a g e
3.4. Series and Shunt Resistance and Operating Temperature Impacts
Analyzing series resistance (Rs), shunt resistance (Rsh), and operating temperature is essential for
optimizing solar cell efficiency, ensuring thermal stability, and maintaining long-term device
reliability. High series resistance reduces power output by limiting current flow, while low shunt
resistance causes leakage losses that degrade performance. Additionally, elevated operating
temperatures negatively impact the open-circuit voltage, accelerate material degradation, and
The impact of series resistance (Rs) on the PV performance of the perovskite solar cell is shown
in Fig. 7(a). As Rs increases from 0 to 5 Ω·cm2, the Voc remains relatively stable, showing a slight
increase from 1.1676 V to 1.1693 V, indicating minimal sensitivity of Voc to Rs within this range.
However, FF decreases significantly from 89.46% to 75.75%, due to increased resistive losses that
impede current flow and reduce charge extraction efficiency. This reduction in FF directly leads
to a notable decline in PCE, which drops from 35.19% at zero series resistance to 29.84% at 5
Ω·cm². Jsc remains nearly unchanged, suggesting that Rs predominantly affects the device’s charge
transport rather than light absorption or charge generation. Overall, minimizing series resistance
The effect of shunt resistance (Rsh) on the PV parameters of the PSC is presented shown in Fig.
7(b). As Rsh increases from 102 Ω·cm2 to 108 Ω·cm2, there is a substantial improvement in FF,
27 | P a g e
which rises sharply from 61.44% at low Rsh (102 Ω·cm2) to approximately 89.46% at high Rsh
values (≥107 Ω·cm2). This improvement in FF reflects the reduction of leakage currents and
recombination losses associated with higher shunt resistance. Correspondingly, the PCE increases
from 23.95% to about 35.19%, indicating that minimizing shunt pathways is critical for optimizing
device performance. Voc and Jsc remain relatively stable across the Rsh range, highlighting that
shunt resistance mainly affects charge extraction and recombination rather than carrier generation.
These results emphasize the importance of maintaining high shunt resistance to ensure maximum
The influence of operating temperature on the PV performance of the perovskite solar cell is
summarized in Fig. 7(b). As the temperature increases from 290 K to 350 K, a gradual decline in
Voc is observed, decreasing from 1.1751 V to 1.1294 V. This reduction is primarily due to enhanced
contrast, Jsc exhibits a slight increase, rising from 33.66 mA/cm2 to 33.82 mA/cm2, which can be
moderately from 89.75% to 87.86%, reflecting increased resistive losses and reduced charge
extraction efficiency with temperature. Consequently, PCE declines from 35.50% to 33.56%.
These results indicate that while higher temperatures slightly boost current generation, the adverse
effects on voltage and FF dominate, leading to overall efficiency loss. Therefore, maintaining
lower operating temperatures is essential to preserve the optimal performance and longevity of
28 | P a g e
Fig. 7. Impact of (a) series resistance (Rs), (b) shunt resistance (Rsh) and (c) temperature on PV
performance.
The above analysis indicates that achieving higher PSC performance requires minimizing Rs while
optimizing shunt resistance Rsh to values of at least 107 Ω·cm2. Furthermore, maintaining the
increased Jsc and decreased Voc and FF, ultimately maximizing the overall power conversion
efficiency.
Fig. 8(a) displays the AM 1.5G illuminated J–V characteristics, yielding a Voc of 1.1676 V, Jsc of
33.67 mA/cm2, FF of 89.46%, and an impressive PCE of 36.19%. These performance indicators
reflect enhanced charge separation and reduced carrier recombination, facilitated by optimized
29 | P a g e
interface. The favorable heterojunction design ensures strong built-in electric fields, improving
carrier mobility and minimizing resistive losses, which collectively contribute to the superior
Fig. 8(b) shows the quantum efficiency (QE) spectrum over the 300-900 nm range, peaking at
98.8% around 390-400 nm. This high QE reflects efficient light absorption and charge collection,
particularly in the near-UV and visible regions. As the wavelength increases beyond 360 nm, QE
gradually decreases to about 85.1% at 900 nm, due to reduced absorption near the band edge of
CH3NH3SnBr3. The strong response at shorter wavelengths is attributed to the perovskite’s high
absorption coefficient, enhancing carrier generation and contributing to the elevated Jsc and overall
device efficiency.
30 | P a g e
4. Conclusion
A lead-free and environmentally friendly CH3NH3SnBr3-based perovskite solar cell has been
designed and numerically simulated using SCAPS-1D. The proposed device structure—
concentrations of each layer to achieve maximum PCE. A comprehensive analysis of the absorber
layer's defect density was also conducted, highlighting its critical influence on key PV parameters
such as Voc, Jsc, FF, and overall PCE. The results suggest that minimizing bulk and interfacial
defects in the CH3NH3SnBr3 absorber is essential for enhancing the performance of lead-free
perovskite solar cells. Optimizing absorber, HTL, ETL, and FTO layer thicknesses is crucial for
peak PSC performance. The absorber at 0.9 μm, HTL at 0.05 μm, ETL at 0.2 μm, and FTO at
0.04 μm balance light absorption, charge extraction, and resistive losses, maximizing efficiency.
Optimizing doping concentrations in absorber, HTL, ETL, and FTO layers is crucial for
PSC performance. Ideal doping improves charge transport, reduces recombination, and
enhances PCE, with optimal values of 1019 cm-3 (absorber), 2×1015 cm-3 (HTL), and
Defect densities in PSC layers and interfaces significantly affect performance by increasing
recombination. The absorber’s defects drastically reduce Voc, Jsc, FF, and PCE, while HTL
and FTO are more tolerant. The absorber/ETL interface is highly sensitive, requiring
Series resistance (Rs) reduces FF and efficiency by hindering charge transport, while shunt
temperature lowers Voc and FF despite slight Jsc gain, causing overall efficiency loss.
Controlling Rs, Rsh, and temperature is vital for stable PSC operation.
31 | P a g e
The quantum efficiency (QE) ranges from 300 to 900 nm, peaking at 98.8% near 390–400
nm, then gradually decreasing to 85.1% at 900 nm due to reduced absorption near the
These findings offer a comprehensive roadmap for improving the efficiency and reliability of lead-
free perovskite solar cells. Ongoing advancements in material and interface engineering will be
crucial to bridging the gap between simulation results and experimental implementation.
References:
[1] Pandey, R. et al. Halide composition engineered a non-toxic perovskite−silicon tandem solar
cell with 30.7% conversion efficiency. ACS Appl. Electron. Mater. 5, 5303–5315.
[Link] (2023).
[2] Bhattarai, S. et al. Comparative study of distinct halide composites for highly efficient cesium-
[Link] (2023).
[3] Mushtaq, Shammas, et al. "Performance optimization of lead-free MASnBr3 based perovskite
solar cells by SCAPS-1D device simulation." Solar Energy 249 (2023): 401-413.
[4] Dungani, Anshi S., et al. "Designing a stable lead-free MASnBr3 perovskite solar cell using
numerical simulation for efficient energy harvesting." International Journal of Modern Physics
B (2025): 2540059.
32 | P a g e
[5] Shah, Masood, et al. "First-Principles insights and SCAPS-1D simulations for optimizing
MASnBr3-based perovskite solar cells." Computational Materials Science 250 (2025): 113699.
[6] Mushtaq, Shammas, et al. "Performance Optimization of Cost-Effective Hole Transport Layer
Free Perovskite Solar Cell by Numerical Simulation." Silicon 17.6 (2025): 1449-1463.
[7] Park, Dong‐Am, and Nam‐Gyu Park. "Perovskite‐Inspired Materials (PIMs): Exploring Their
[8] Eperon, Giles E., et al. “Formamidinium Lead Trihalide: A Broadly Tunable Perovskite for
Efficient Planar Heterojunction Solar Cells.” Energy & Environmental Science, vol. 7, no. 3, 2014,
[9] Green, M. A., et al. “Solar Cell Efficiency Tables (Version 57).” Progress in Photovoltaics:
Research and Applications, vol. 29, no. 1, 2021, pp. 3–15, [Link]
[10] Li, Wenguang, et al. "Mitigating Lead Toxicity in Halide Perovskite Solar Cells: Strategies
[11] Al-Aaraji, Mohammed N., Wisam N. Hasan, and Kutaiba Al-Marzoki. "Progress In Lead Free-
Relaxor Ferroelectrics For Energy Storage Applications." Journal of Physics: Conference Series.
[12] Jeon, N. J., et al. “Compositional Engineering of Perovskite Materials for High-Performance
Solar Cells.” Nature, vol. 517, no. 7535, 2015, pp. 476–80, [Link]
[13] Noel, Nakita K., et al. “Lead-Free Organic–Inorganic Tin Halide Perovskites for Photovoltaic
Applications.” Energy & Environmental Science, vol. 7, no. 9, 2014, pp. 3061–3068,
[Link]
33 | P a g e
[14] Tsai, Hsing-Yin, et al. “A Lead-Free Halide Double Perovskite Cs2AgBiBr6 for Photovoltaic
[Link]
[15] Hao, Feng, et al. "Lead-free solid-state organic–inorganic halide perovskite solar
[16] Bhattarai, Sagar, and T. D. Das. "Optimization of carrier transport materials for the
performance enhancement of the MAGeI3 based perovskite solar cell." Solar Energy 217 (2021):
200-207.
[17] Saha, Prithick, Sangeeta Singh, and Sanjib Bhattacharya. "FASnI3-based eco-friendly
heterojunction perovskite solar cell with high efficiency." Micro and Nanostructures 186 (2024):
207739.
[18] Li, Song, et al. "A brief review of hole transporting materials commonly used in perovskite
[19] Zhang, Wei, et al. “Enhanced Electron Transport in Perovskite Solar Cells Using Sulfur-
Doped Tin Oxide.” Journal of Materials Chemistry A, vol. 11, no. 15, 2023, pp. 7845–7853.
[20] Alam, Intekhab, Rahat Mollick, and Md Ali Ashraf. "Numerical simulation of Cs2AgBiBr6-
based perovskite solar cell with ZnO nanorod and P3HT as the charge transport layers." Physica
[21] Jayan, K. Deepthi. "Enhancement of efficiency of (FA) 2BiCuI6 based perovskite solar cells
34 | P a g e
[22] Kavithaa, M. V., C. K. Anjalia, and K. S. Sudheerb. "Device simulation and optimization of
HTL-free perovskite solar cell with CH3NH3SnBr3 as the absorber layer using solar cell
[23] Nyiekaa, Emmanuel A., et al. "Simulation and optimization of 30.17% high performance N-
type TCO-free inverted perovskite solar cell using inorganic transport materials." Scientific
Highlights
Optimized the effect of the different hole and ETLs on the cell performance.
35 | P a g e
The thicknesses, doping density, defect density of HTL/ETL/absorber layer also
optimized.
efficiency 35.19%.
36 | P a g e