0% found this document useful (0 votes)
5 views13 pages

229 LiuSH

Uploaded by

Lakhdar Dehimi
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)
5 views13 pages

229 LiuSH

Uploaded by

Lakhdar Dehimi
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/ 13

Chalcogenide Letters Vol. 21, No. 3, March 2024, p.

229 - 241

Numerical simulation of Sb2Se3-based solar cells

S. H. Liua, J. R. Yuana,*, Y. Wua, X. H. Denga, Q. M. Yub


a
School of Physics and Materials Science, Nanchang University, Nanchang
330031, China
b
School of Chemistry and Chemical Engineering, Nanchang University, Nanchang
330031, China

Antimony selenide (Sb2Se3) has remarkable optoelectronic capabilities that make it a


promising option for the next generation solar cells. In this work, a solar cell with the
structure Al/FTO/CdS/Sb2Se3/Mo is modeled and numerically analyzed using SCAPS-1D
program. Furthermore, a Al/FTO/CdS/Sb2Se3/Sb2S3/Mo solar cell structure that uses Sb2S3
as the back surface field (BSF) layer is proposed. A comprehensive examination of
photovoltaic characteristics for the solar cells was carried out. The optimization process
involved adjusting the operating temperature, series and shunt resistance, doping
concentration, bulk defect density, back contact metal work function, and thickness of the
absorber layer. The optimized Sb2Se3-based solar cell with Sb2S3 material showed a
conversion efficiency of 28.91%, suggesting that Sb2Se3-based solar cells have a great deal
of potential for further development.

(Received December 16, 2023; Accepted March 4, 2024)

Keywords: Antimony selenide, Solar cells, SCAPS-1D, Conversion efficiency

1. Introduction

As fossil energy resources continue to deplete and environmental pollution intensifies,


there is an urgent quest for clean and renewable energy sources. Thin-film solar cells, owing to
their advantages such as minimal raw material consumption, simple preparation processes, and the
material exhibits flexibility, have garnered increasing attention in recent years [1]. Notably, CIGS
and CdTe photovoltaic cells have achieved industrial-scale production, while the laboratory
efficiency of perovskite thin-film solar cells is comparable to that of crystalline silicon solar cells.
However, the presence of rare elements such as indium and gallium in CIGS, the toxicity
associated with cadmium in cadmium telluride, and the imperative need to enhance the stability of
perovskite thin-film solar cells pose constraints on the further development of thin-film solar cell
technologies [2-6]. In recent years, the p-type semiconductor Sb2Se3 has become the focus of
widespread attention in the scientific community. This material boasts numerous advantages,
including an appropriate bandgap, rational charge carrier mobility, high absorption coefficients,
1
abundant elements in the earth's crust, and non-toxicity [7-10]. Consequently, it holds the
potential to serve as an absorber layer material for solar cells. Through the Shockley-Queisser limit
theory, we can calculate that the photovoltaic conversion efficiency of Sb2Se3-based solar cells
could theoretically reach a maximum of approximately 32% [11]. In previous endeavors to
enhance the performance of Sb2Se3-based solar cells, researchers have devised and refined
numerous device structures. The efficiency of cells employing a CdS/Sb2Se3 configuration reached
7.6% [12]. Experimental conversion efficiency for cells adopting the structure
ZnO:Al/ZnO/CdS/TiO2/Sb2Se3 nanorod arrays/MoSe2/Mo amounted to 9.2% [13]. Theoretical
conversion efficiency for solar cells with the FTO/CdS/Sb2Se3/Au configuration reached 16.5%
[14]. Within the structure of Sb2Se3-based thin-film solar cells ITO/ZnO/CdS/Sb2Se3/CNT, with
carbon nanotubes (CNT) serving as the back contact, a theoretical optimum efficiency of 21.67%
was achieved [15]. In the configuration of FTO/CdS/Sb2Se3/HTL/Au solar cells, utilizing CuO as
the hole transport layer (HTL), the theoretical efficiency reached 23.18% [16]. However, when
compared to some other thin-film solar cells, Sb2Se3-based thin-film solar cells lack sufficient
competitive advantages in terms of conversion efficiency and other performance metrics. Hence,
there remains a need to design high-efficiency, low-cost Sb2Se3-based thin-film solar cells.
*
Corresponding author: [email protected]
https://doi.org/10.15251/CL.2024.213.229
230

Numerous factors impact the performance of solar cells, including doping concentration,
absorber layer thickness, work function of the back contact electrode, operating temperature, and
carrier recombination loss. Prior studies have demonstrated that the application of a BSF layer to
the back of the absorber layer can substantially minimize the back surface recombination loss,
consequently improving overall cell performance and conversion efficiency [17-18]. Because of its
high absorption coefficient (>5×104 cm-1) and suitable bandgap (~1.6 eV), Sb2S3 is the ideal
material for the BSF layer in Sb2Se3-based thin-film solar cells. To optimize Sb2Se3-based solar
cells, this study proposes a structure incorporating Sb2S3 as the BSF layer. The SCAPS-1D
software is employed in this work to compare the performance of Sb2Se3-based solar cells with
and without the inclusion of a BSF layer.

2. Device structure and simulation methodology

2.1. Device structure


The device structure employed in this simulation comprises Al/FTO/CdS/Sb2Se3/Mo, as
illustrated in Fig. 1(a). Herein, the p-type semiconductor Sb2Se3 is used as the absorber layer, the
n-type semiconductor CdS is served as the buffer layer, SnO2:F (FTO) is acted as the window
layer, and molybdenum (Mo) and aluminum (Al) are served as the back contact and front contact,
respectively. Al has a metal work function of 4.08 eV [19], while Mo has a metal work function of
4.95 eV[20]. In the aforementioned configuration, we have introduced Sb2S3 as the BSF layer as
depicted in Figure 1(b). This addition aims to mitigate carrier recombination loss in the solar cell
while enhancing the collection of photogenerated carriers.

(a) (b)
Fig. 1. Schematic structure of the studied Sb2Se3 solar cells: (a) basic Al/FTO/CdS/Sb2Se3/Mo
(without BSF layer) cell and (b) proposed cell Al/FTO/CdS/Sb2Se3/Sb2S3/Mo (with BSF layer).

2.2. Simulation methodology


In the exploration of thin-film solar cell performance, researchers widely employ various
software tools, such as AMPS, COMSOL, wxAMPS and SCAPS-1D. For this study, we have
opted for SCAPS-1D. This software serves as a one-dimensional solar cell simulator, utilized for
modeling the performance and various crucial parameters of solar cells[21]. The program is based
on numerical solutions for the fundamental semiconductor equations, Poisson equation, and
continuity equation (Equations (1)-(3))[22], utilized for the computation of performance
parameters in Sb2Se3-based solar cells.

𝜕𝜕2 Ψ 𝑞𝑞
+ [𝑝𝑝(𝑥𝑥) − 𝑛𝑛(𝑥𝑥) + 𝑁𝑁𝐷𝐷 − 𝑁𝑁𝐴𝐴 + 𝜌𝜌𝑝𝑝 − 𝜌𝜌𝑛𝑛 ] = 0 (1)
𝜕𝜕𝑥𝑥 2 𝜀𝜀

1 𝑑𝑑𝐽𝐽𝑝𝑝
= 𝐺𝐺𝑜𝑜𝑜𝑜 (𝑥𝑥) − 𝑅𝑅(𝑥𝑥) (2)
𝑞𝑞 𝑑𝑑𝑑𝑑
231

1 𝑑𝑑𝐽𝐽𝑛𝑛
= −𝐺𝐺𝑜𝑜𝑜𝑜 (𝑥𝑥) + 𝑅𝑅(𝑥𝑥) (3)
𝑞𝑞 𝑑𝑑𝑑𝑑

In the equations presented, ε is dielectric constant, q represents the charge of an electron,


NA and ND denote the acceptor and donor densities, Ψ signifies the electrostatic potential, p and n
stand for hole and electron concentrations, ρp and ρn represent hole and electron distribution, and
Jp and Jn denote hole current density and electron current density, respectively. Gop(x) is the
optical generation rate, and R(x) net combination rate. All these parameters are functions of the
spatial coordinate x. SCAPS partitions a solar cell structure into slab and main grid points,
discretizing the aforementioned differential equations into sets of algebraic equations [21-22]. The
open-circuit voltage (Voc), short-circuit current (Jsc), fill factor (FF), conversion efficiency (η),
band diagram, I-V characteristics, and other solar cell parameters can all be computed using the
solutions to these equations.

2.3. Simulation parameters


A comprehensive summary of key parameters is provided in Tables 1 and 2, collating data
from literature and experimental results [19-22]. Throughout all simulations, we utilized
illumination with a front light intensity of 1000 W/m2 and AM1.5G spectrum, maintaining a
temperature of 300 K. The absorption coefficient of Sb2Se3 is set to 105 cm−1 [22], In each layer of
materials, the approximate thermal velocities of the holes and electrons are 107 cm/s at room
temperature [19]. The surface recombination velocities of electrons and holes for Mo and Al are
both 107 cm/s. The energy levels of defects are modeled as a neutral Gaussian distribution with a
characteristic energy of 0.1 eV [22].

Table 1. Simulation parameters for the Sb2Se3-based solar cell.

Material Properties SnO2 CdS Sb2Se3 Sb2S3


Thickness (μm) 0.05 0.05 1.0* 0.05
Band gap (eV) 3.6 2.4 1.09 1.62
Electron affinity (eV) 4 4.2 3.9 3.7
Dielectric permittivity 9 10 18 9.00
18 18 18
Conductor band effective density 2.2 × 10 2.2 × 10 2.2 × 10 2.2 × 1018
of states (cm-3)
Value band effective density of 1.8 × 1019 1.8 × 1019 1.8 × 1019 1.8 × 1019
states (cm-3 )
Electron thermal velocity (cm/s) 1 × 107 1 × 107 1 × 107 1 × 107
Hole thermal velocity (cm/s) 1 × 107 1 × 107 1 × 107 1 × 107
Electron mobility (cm2 /Vs) 100 100 15 9.8
2
Hole mobility (cm /Vs) 25 25 5.1 10
18 17
Shallow uniform donor density 1 × 10 1 × 10 0 0
-3
ND (cm )
Shallow uniform acceptor density 0 0 1 × 1017* 1 × 1018
−3
NA (cm )
Defect type Single Single Single Single
Acceptor Acceptor Donor Donor
Energetic distribution Gaussian Gaussian Gaussian Gaussian
−3 15 15 14
Defect density (cm ) 1 × 10 1 × 10 1 × 10 1 × 1015
Note: * is a variable field.
232

Table 2. Interface parameters used in the FTO/CdS/Sb2Se3/Sb2S3 heterojunction device Simulation.

Parameters CdS/Sb2Se3 interface Sb2Se3/Sb2S3 interface


Defect type Neutral Neutral
− 19
Capture cross-section of 1 × 10 1 × 10− 19
electrons (cm2 )
Capture cross-section of holes 1 × 10− 19 1 × 10− 19
(cm2 )
Reference for defect energy above the highest Ev above the highest Ev
level Et
Energy with respect to 0.01 0.01
Reference (eV)
Total density (cm− 2 ) 1 × 1010 1 × 1010

3. Results and discussion

In this study, our focal point lies in investigating the impact of various parameters of the
absorber layer on the performance of Sb2Se3-based TFSCs. We employ SCAPS-1D software for
simulation, exploring the effects of Sb2Se3 absorber layer thickness, doping concentration, bulk
defect density, as well as the back contact metal work function, operating temperature, and series
and parallel resistances of the solar cell on the performance of Sb2Se3-based solar cells.
Simultaneously, we conduct a comparative analysis, examining parameters such as Voc, Jsc, FF,
and η for Sb2Se3-based solar cells with and without an Sb2S3 BSF layer under identical conditions.
Furthermore, we optimize various parameters of the solar cell based on this comparison. Utilizing
the optimized data, we delve into the J-V characteristics, band diagrams, built-in electric fields,
and other aspects of the cell, This facilitates the creation of solar cell structures with higher
efficiency.

3.1. Energy band diagram


Fig. 2 illustrates the band structure of Sb2Se3-based solar cells with Sb2S3 as the BSF layer.
At the interface of Sb2Se3/CdS, photo-generated charge carriers are effectively separated and
rapidly accelerated away from the interface of Sb2Se3/CdS under the influence of the electric field.
Holes traverse the Sb2Se3 absorber layer, being collected by the metallic Mo, while electrons flow
into the CdS buffer layer. From the diagram, it is evident that in this simulated study, each layer's
thickness and bandgap are discernible. We observe the formation of a p+-Sb2S3/p-Sb2Se3
heterojunction between the Sb2S3 BSF layer and the Sb2Se3 absorber layer. Consequently, a
potential barrier is established at the interface of the Sb2S3 BSF layer and the Sb2Se3 absorber layer,
aiding in impeding the flow of electrons to the back surface of the solar cell. This potential barrier
formation contributes to mitigating carrier recombination loss in the solar cell[19].
233

Fig. 2. Band diagram of Sb2Se3-based solar cell with BSF layer.

3.2. Effect of absorber layer thickness on the device


The performance of the solar cells is influenced by the thickness of the absorption layer,
encompassing the fill factor and conversion efficiency. When the absorption layer is excessively
thin, the solar cell fails to fully capture all incident light, resulting in a decline in cell efficiency, as
explicitly depicted in Formula (4):

𝑎𝑎(𝜆𝜆, 𝑊𝑊) = 1 − 𝑒𝑒 (−2𝑎𝑎(𝜆𝜆)𝑊𝑊) (4)

Here, W represents the thickness of the solar cell, λ is the wavelength of absorbed light,
and 𝑎𝑎(λ) denotes the absorption of light with wavelength λ [22]. Insufficient thickness of the
absorption layer results in an inability to capture all incident light, consequently leading to a
decrease in efficiency [23]. Fig. 3 illustrates the impact of varying the thickness of the Sb2Se3
absorber layer, ranging from 0.1 μm to 2 μm, on the performance parameters of Sb2Se3-based
TFSCs. Additionally, we conducted a comparative analysis between solar cells with and without a
Sb2S3 BSF layer to examine its influence.

Fig. 3. Effect of the absorber layer thickness on the parameters of the solar cell (Voc, Jsc, η, FF).
234

Throughout this process, we maintained constant values for other parameters. Through
simulation, we derived the solar cell's Voc, Jsc, FF, and η to explore the effects of absorber layer
thickness on solar cell performance.
In thin-film solar cells (TFSCs) without the Sb2S3 BSF layer, we observe an ascending
trend in the Voc, Jsc, FF, and η as the absorber layer thickness increases. However, beyond a
thickness of 1 μm, the upward trajectory of these parameters begins to plateau. At an absorber
layer thickness of 0.1 μm, Voc is 0.63V, Jsc is 21.25 mA/cm², FF is 73.56%, and η is 9.83%. As the
thickness increases to 2 μm, Voc becomes 0.7V, Jsc reaches 41.46 mA/cm², and FF and η attain
values of 82.21% and 23.99%, respectively. In Sb2Se3-based solar cells with the inclusion of the
Sb2S3 BSF layer, at an absorber layer thickness of 0.1 μm, the FF is 80.18%, and η is 16.8%. With
increasing thickness, FF and η continue to rise. After reaching a thickness of 0.5 μm, the upward
trend in FF and η begins to decelerate. At an absorber layer thickness of 1 μm, the solar cell
achieves maximum FF and η values, reaching 85.4% and 28.91%, respectively. The impact of
absorber layer thickness on solar cell performance primarily manifests in two aspects: the
collection of photons and the efficiency of charge carrier transmission. With a thinner absorber
layer, the impediment to charge carrier transmission within the cell is minimal, and photon
absorption plays a dominant role in solar cell performance. As the absorber layer thickness
increases, the number of absorbed photons also increases, thereby enhancing the solar cell's
performance. However, as the absorber layer thickness continues to rise, the efficiency of charge
carrier transmission becomes pivotal in determining the FF and conversion efficiency of the solar
cell. The increase in absorber layer thickness results in a longer path for charge carrier
transmission, leading to partial recombination losses during the transmission process, thereby
affecting solar cell performance. Under the combined influence of these factors, selecting an
appropriate absorber layer thickness becomes crucial for optimizing solar cell performance. As the
absorber layer thickness increases, the recombination rate of charge carriers exceeds the generation
rate, causing Jsc to approach saturation.
Through comparative analysis, we discern that, during the transition from an absorber
layer thickness of 0.1 μm to 2.0 μm, the performance of Sb2Se3-based solar cells with the inclusion
of the Sb2S3 BSF layer consistently surpasses that of solar cells lacking the Sb2S3 buffer layer. At a
thickness of 1.0 μm, the solar cell with the BSF layer exhibits a Voc of 0.78V, Jsc of 43.22 mA/cm²,
FF of 85.4%, and an efficiency of 28.91%. In contrast, the Sb2Se3-based TFSCs without the BSF
layer displays a Voc of 0.68V, Jsc of 39.28 mA/cm², FF of 80.34%, and an efficiency of 21.66%.
The introduction of the Sb2S3 BSF layer results in the formation of a p+-Sb2S3/p-Sb2Se3
heterojunction, creating a distinctive electric field at the interface between the absorber layer and
the BSF layer, as depicted in Fig. 4.This electric field hinders the passage of minority carriers from
the backside of the solar cell, effectively reducing recombination losses and diminishing dark
current within the cell. Additionally, the reduction in dark current within the cell contributes to the
elevation of Jsc, thereby enhancing the performance of the Sb2Se3-based thin-film solar cells
[19,25]. Comparative studies reveal that the incorporation of the Sb2S3 BSF layer not only
significantly improves the overall efficiency of the cell but also enables the attainment of higher
Voc, Jsc, FF, and η, even as the thickness of the absorber layer is minimized.

Fig. 4. The built-in electric field at the p+-Sb2S3 / p-Sb2Se3 heterojunction


235

The reduction in absorber layer thickness not only enhances performance but also proves
instrumental in diminishing material costs required for production. Consequently, the inclusion of
the Sb2S3 BSF layer plays a pivotal role in achieving a thinner, more efficient Sb2Se3-based
thin-film solar cells.

3.3. The impact of series resistance and parallel resistance on solar cell performance.
The complexity of series resistance, which is made up of several resistances, must be
taken into account in order to fully comprehend the effects of shunt resistance (Rsh) and series
resistance (Rs).For example, each layer of material within the cell contributes to resistance, there is
resistance between semiconductor materials and the contact points with front and rear metals, and
the metal electrode itself possesses resistance. These resistances collectively form the series
resistance Rs. Simultaneously, the occurrence of shunt resistance is associated with reverse
saturation current [26].To investigate the effects of Rs and Rsh, while keeping other parameters at
their optimized values, we varied Rs and Rsh in the ranges of 0 to 21Ω-cm2 and 10 to 300Ω-cm2,
respectively.
The effect of Rs and Rsh on the solar cell's performance is shown in Fig. 5. From Fig. 5, it
can be observed that as Rs increases and Rsh decreases, the cell efficiency decreases from 27.31%
to 1.51%. Jsc and FF gradually decrease, while the variation in Voc is significantly influenced by
Rsh, increasing from 0.43V to 0.78V. These observations indicate that high series resistance and
low shunt resistance can significantly degrade the solar cell performance. Therefore, achieving low
series resistance and high shunt resistance is imperative for realizing high efficiency and
outstanding performance in solar cells.

(a) (b)

(c) (d)
Fig. 5. The influence of series resistance and parallel resistance on the parameters of
the solar cell (Voc, Jsc, FF, η).
236

Furthermore, this study reveals that lower Rsh results in a reduction in Voc and Jsc,
consequently lowering the maximum power of the cell and ultimately affecting efficiency and fill
factor. Thus, this research underscores the crucial impact of series resistance and shunt resistance
on the cell performance.

3.4. Current density-voltage characteristics


Fig. 6(a) illustrates the I-V characteristics of two solar cell structures. In the absence of
illumination, the energy supply to the cell emanates from the externally applied voltage. The
diffusion current generated by the majority charge carriers exhibits exponential growth under the
influence of external voltage, displaying characteristics akin to a diode. Under illuminated
conditions, the energy supply of the cell originates from internally generated photo-generated
carriers. The increased abundance of minority charge carriers, facilitated by their drift, results in a
substantial current density. Hence, we consider that the fundamental mechanisms governing
current transmission in solar cells involve the diffusion and drift of carriers [19,27].
Fig. 6(b) depicts the C-V characteristics of the two solar cell structures. Beyond a voltage
threshold of 0.5V, the capacitance experiences rapid ascent. Notably, solar cells incorporating an
Sb2S3 BSF layer exhibit a swifter rise in capacitance. A greater rate of change in capacitance with
voltage amplifies the differential capacitance, promoting faster drift and response of
photogenerated carriers, thereby contributing to the enhancement of solar cell performance.

Fig. 6. (a)J-V characteristic, (b)C-V characteristic, (c) Mott - Schottky plot, (d) Conductance-Voltage
characteristic.

Figure 6(c) depicts the Mott-Schottky curves of the solar cells, from which we can deduce
that the built-in potential (Vbi) for both types of cells is 0.72V.A higher Vbi accentuates the
depletion layer in the cell, facilitating the absorption of photons and the generation of electron-hole
pairs[28-29]. Moreover, Vbi directly impacts the Voc, Jsc, FF, thereby exerting a significant
237

influence on the conversion efficiency of solar cells. Fig. 6(d) illustrates the variation of
conductance with voltage. Post 0.6V, conductance undergoes rapid augmentation. Comparative
analysis indicates that the introduction of an Sb2S3 BSF layer results in a notable enhancement in
both the I-V and C-V characteristics of solar cells, significantly contributing to improvements in
Voc, Jsc, FF, and η.

3.5. Effect of the variation in the concentration of shallow acceptor density


Fig. 7 illustrates the impact of the carrier concentration in the absorber layer, ranging from
1×1012 to 1×1017 cm−3, on the Voc, Jsc, FF, and η of Sb2Se3-based TFSCs with an Sb2S3 BSF layer.
It is observed that Jsc remains nearly constant throughout this process. However, as the carrier
concentration increases, the TSA conversion process in the BSF layer is affected, leading to a
slight decrease in Jsc [30-31]. With the rise in carrier concentration, Vbi also increases, resulting in
a reduction in the internal recombination current within the cell [32]. Consequently, the Voc, FF,
and η of the solar cell are influenced. As the carrier concentration in the absorber layer increases
from 1×1012 to 1×1017 cm−3, Voc rises from 0.72 to 0.78V. The fill factor and conversion efficiency
also increase from 76.3% and 24.08% to 85.4% and 28.91%, respectively.

Fig. 7. The influence of carrier concentration in Sb2Se3 absorber layer on various parameters
of solar cell (Voc, Jsc, FF, η).

3.6. Effect of variation in back contact work function


The back contact work function has a significantly impact on solar cells' efficiency. In
order to avoid the formation of a Schottky junction at the interface, the back contact metals for
Sb2Se3-based solar devices, such as Au, Ag, Mo, and Pt, among others, are expected to have work
functions greater than 4.8 eV [33].Fig. 8. depicts the influence of the back contact work function
on the conversion efficiency of solar cells. In our experiments, we adjusted the back contact work
function between 4.7 and 5.1 eV in order to evaluate the performance of the solar cell. However,
the efficiency of the cell stopped showing appreciable growth once the work function reached 4.95
eV. We chose Mo, whose work function is 4.95 eV, as the back contact in order to optimize the
solar cell's structure and performance. This decision enhances the performance of the cell.
238

Fig. 8. Effect of back contact work function on the conversion efficiency of Sb2Se3-based solar cells
containing a BSF layer.

3.7 Effect of variation in the concentration of defect density


Fig. 9 demonstrates the impact of the defect density in the absorber layer on the output
parameters of Sb2Se3-based solar cells with a BSF layer. Throughout the simulation process, other
parameters were maintained at their optimized values, while the defect density varied from 1×1013
to 1×1017 cm−3. It is observed that with the increase in defect density, both Voc and Jsc gradually
decrease. Conversely, FF initially increases, reaching a maximum of 85.19% at a defect density of
1×1014 cm−3, and then steadily decreases. At defect densities of 1×1014 cm−3 and 1×1017 cm−3, the
efficiency of the solar cell is 28.91% and 10.69%, respectively. As the photo-generated current is
primarily produced by the Sb2Se3 absorber layer, an increase in defect density leads to a reduction
in carrier lifetime and diffusion length, resulting in carrier recombination losses and a substantial
decrease in solar cell efficiency [34]. Through our investigation, we ascertain that a high defect
density in the absorber layer significantly diminishes the output parameters of the solar cell. To
optimize the Sb2Se3-based solar cell, we choose a defect density of 1×1014 cm−3, at which point the
cell achieves a conversion efficiency of 28.91%.

Fig. 9. Effect of defect density of absorber layer on various parameters of solar cell (Voc, Js c, FF, η).
239

3.8. Effect of absorber layer temperature on the device


To validate the stability of the proposed Sb2Se3-based solar cell, we conducted a detailed
analysis of the influence of operating temperature in the range of 250 to 450 K on the cell
performance. In this numerical study, all other parameters remained constant, as specified in Table
1. Figure 10 illustrates the performance parameters of Sb2Se3-based solar cells, both without a BSF
and with an Sb2S3 BSF layer, at different operating temperatures, including Voc, Jsc, FF, and
efficiency.

Fig. 10. The effect of temperature on the parameters of the solar cell (Voc, Jsc, FF, η).

It is observed from Fig. 10 that the Voc, Jsc, FF, and efficiency exhibit a decreasing trend ,
as the temperature increases. The results indicate that, compared to solar cells without a BSF layer,
those with an Sb2S3 BSF layer demonstrate superior performance. The Voc of the cell gradually
decreases with rising temperature. It is attributed to enhance reverse saturation current caused by
temperature elevation. Alongside an increase in series resistance and a reduction in carrier
diffusion length, the cell generates more interface defects. With an increase in temperature, the
material bandgap decreases, resulting in a minor upward trend in Jsc for the cell [27].At 250 K, the
conversion efficiency of the solar cell with a BSF layer is 31.48%; at 450 K, the efficiency
decreases to 18.96%. This decrease is attributed to the impact of elevated temperature on the
mobility of holes and electrons, as well as the concentration of carriers, resulting in a overall
decrease in the solar cell efficiency [26].

4. Conclusion

In this study, we utilized the SCAPS-1D software to investigate the performance of


Sb2Se3-based thin-film solar cells. We proposed the Sb2S3 as a back surface field (BSF) layer for
traditional Sb2Se3-based solar cells. We conducted a comparative analysis, studying the impact of
absorber layer thickness, bulk defect density in the absorber layer, doping concentration,
environmental temperature, back contact work function, and series-parallel resistance on the
performance of Sb2Se3-based TFSCs with and without the Sb2S3 BSF layer. The results indicate
that incorporating the Sb2S3 BSF layer in traditional Sb2Se3-based solar cells effectively reduces
internal carrier recombination losses, thereby enhancing open-circuit voltage, current density, fill
factor, and conversion efficiency. Maintaining other parameters constant, the addition of an Sb2S3
BSF layer to traditional Sb2Se3-based solar cells increases the conversion efficiency from 21.66%
240

to 28.91%. In addition, the addition of Sb2S3 can improve the solar cell performance in a number
of ways while reducing the thickness of the absorber layer and optimizing the cell structure. The
optimized solar cell parameters include a 1.0 μm thick Sb2Se3 absorber layer, a 0.05μm thick Sb2S3
BSF layer, and a carrier concentration of 1×1017 cm−3 in the Sb2Se3 absorber layer. When the bulk
defect density in the absorber layer decreases to 1×1014 cm−3, the Sb2Se3-based solar cell achieves
an optimal conversion efficiency of 28.91%, with Voc of 0.78 V, Jsc of 43.22 mA/cm2, and FF of
85.4%. The introduction of the Sb2S3 BSF layer also improves the cell performance and stability of
Sb2Se3-based thin-film solar cells. These results suggest significant potential for the development
of Sb2Se3-based solar cells, and numerical simulations can contribute to the creation of low-cost,
high-efficiency thin-film solar cells based on Sb2Se3.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No.
11964018) and the Natural Science Foundation of Jiangxi Province of China (Grant No.
20224BAB202032). The authors acknowledge the use of the SCAPS-1D program developed by
Prof. Burgelman's group of the University of Gent, Belgium.

References

[1] H. Jalali, A. A. Orouji, I. Gharibshahian, Solar Energy Materials and Solar Cells 260, 112492
(2023); https://doi.org/10.1016/j.solmat.2023.112492
[2] S. Abbas, S. Bajgai, S. Chowdhury, A. S. Najm, M. S. Jamal, K. Techato, S. Channumsin, S.
Sreesawet, M. Channumsin, A. Laref, Materials 15(18), 6272 (2022);
https://doi.org/10.3390/ma15186272
[3] K. K. Maurya, V. N. Singh, Heliyon 8(10), e10925 (2022);
https://doi.org/10.1016/j.heliyon.2022.e10925
[4] Y. Singh, K. K. Maurya, V. N. Singh, Mater. Today Sustainability 18, 100148 (2022);
https://doi.org/10.1016/j.mtsust.2022.100148
[5] M. D. Chatzisideris, N. Espinosa, A. Laurent, F. C. Krebs, Solar Energy Materials and Solar
Cells 156, 2 (2016); https://doi.org/10.1016/j.solmat.2016.05.048
[6] K. K. Maurya, V. N. Singh, J. Sci.:-Adv. Mater. Device. 7(2), 100445 (2022);
https://doi.org/10.1016/j.jsamd.2022.100445
[7] S. Chowdhury, A. S. Najm, M. Luengchavanon, A. M. Holi, C. H. Chia, K. Techato, S.
Channumsin, I. K. Salih, Energ Fuel 37(9), 6722 (2023);
https://doi.org/10.1021/acs.energyfuels.2c03593
[8] Y. Zhou, L. Wang, S. Chen, S. Qin, X. Liu, J. Chen, D.-J. Xue, M. Luo, Y. Cao, Y. Cheng, E.
H. Sargent, J. Tang, Nat. Photonics 9(6), 409 (2015); https://doi.org/10.1038/nphoton.2015.78
[9] C. Chen, W. Li, Y. Zhou, C. Chen, M. Luo, X. Liu, K. Zeng, B. Yang, C. Zhang, J. Han, Appl.
Phys. Lett. 107(4), 043905 (2015); https://doi.org/10.1063/1.4927741
[10] A. Ait Abdelkadir, M. Sahal, E. Oublal, N. Kumar, A. Benami, Opt. Quantum Electron. 55(6),
514 (2023); https://doi.org/10.1007/s11082-023-04797-7
[11] S. Rühle, Sol. Energy 130, 139 (2016); https://doi.org/10.1016/j.solener.2016.02.015
[12] X. Wen, C. Chen, S. Lu, K. Li, R. Kondrotas, Y. Zhao, W. Chen, L. Gao, C. Wang, J. Zhang,
G. Niu, J. Tang, Nat. Commun. 9(1), 2179 (2018); https://doi.org/10.1038/s41467-018-04634-6
[13] Z. Li, X. Liang, G. Li, H. Liu, H. Zhang, J. Guo, J. Chen, K. Shen, X. San, W. Yu, R. E. I.
Schropp, Y. Mai, Nat. Commun. 10(1), 125 (2019); https://doi.org/10.1038/s41467-018-07903-6
[14] L.-y. Lin, L.-q. Jiang, Y. Qiu, B.-d. Fan, J. Phys. Chem. Solids 122, 19 (2018);
https://doi.org/10.1016/j.jpcs.2018.05.045
[15] F. Baig, Y. H. Khattak, B. M. Soucase, S. Beg, S. R. Gillani, S. Ahmed, J. Nanoelectron.
Optoelectron. 14(1), 72 (2019); https://doi.org/10.1166/jno.2019.2451
[16] Z.-Q. Li, M. Ni, X.-D. Feng, Mater. Res. Express 7(1), 016416 (2020);
https://doi.org/10.1088/2053-1591/ab5fa7
[17] M. S. Rana, M. M. Islam, M. Julkarnain, Sol. Energy 226, 272 (2021);
https://doi.org/10.1016/j.solener.2021.08.035
241

[18] S. Souri, M. Marandi, Opt. Quantum Electron. 55(5), 397 (2023);


https://doi.org/10.1007/s11082-023-04563-9
[19] S. R. Al Ahmed, A. Sunny, S. Rahman, Solar Energy Materials and Solar Cells 221, 110919
(2021); https://doi.org/10.1016/j.solmat.2020.110919
[20] K. K. Maurya, V. N. Singh, Sol. Energy 228, 540 (2021);
https://doi.org/10.1016/j.solener.2021.09.080
[21] A. Basak, U. P. Singh, Solar Energy Materials and Solar Cells 230, 111184 (2021);
https://doi.org/10.1016/j.solmat.2021.111184
[22] K. K. Maurya, V. N. Singh, Sol. Energy 230, 803 (2021);
https://doi.org/10.1016/j.solener.2021.11.002
[23] R. T. Mouchou, T. C. Jen, O. T. Laseinde, K. O. Ukoba, Mater. Today-Proc. 38, 835 (2021);
https://doi.org/10.1016/j.matpr.2020.04.880
[24] R. Kumari, Mamta, A. K. Chaudhary, V. N. Singh, Adv. Theor. Simul., 2300322 (2023);
https://doi.org/10.1002/adts.202300322
[25] I. Gharibshahian, A. A. Orouji, S. Sharbati, Solar Energy Materials and Solar Cells 212,
110581 (2020); https://doi.org/10.1016/j.solmat.2020.110581
[26] H. Heriche, Z. Rouabah, N. Bouarissa, Int. J. Hydrogen Energy 42(15), 9524 (2017);
https://doi.org/10.1016/j.ijhydene.2017.02.099
[27] A. Laidouci, A. Aissat, J. P. Vilcot, Sol. Energy 211, 237 (2020);
https://doi.org/10.1016/j.solener.2020.09.025
[28] B. K. Mondal, S. K. Mostaque, M. A. Rashid, A. Kuddus, H. Shirai, J. Hossain, Superlattices
Microstruct. 152, 106853 (2021); https://doi.org/10.1016/j.spmi.2021.106853
[29] P. Kumari, U. Punia, D. Sharma, A. Srivastava, S. K. Srivastava, Silicon 15(5), 2099 (2023);
https://doi.org/10.1007/s12633-022-02163-y
[30] B. K. Mondal, S. K. Mostaque, J. Hossain, Heliyon 8(3), e09120 (2022);
https://doi.org/10.1016/j.heliyon.2022.e09120
[31] A. Kuddus, A. B. M. Ismail, J. Hossain, Sol. Energy 221, 488 (2021); https://doi.org/
10.1016/j.solener.2021.04.062
[32] J. Hossain, M. Rahman, M. M. A. Moon, B. K. Mondal, M. F. Rahman, M. H. K. Rubel, Eng.
Res. Express 2(4), 045019 (2020); https://doi.org/10.1088/2631-8695/abc56c
[33] Mamta, R. Kumari, R. Kumar, K. K. Maurya, V. N. Singh, Sustainability 15(13), 10465
(2023); https://doi.org/10.3390/su151310465
[34] S. R. I. Biplab, M. H. Ali, M. M. A. Moon, M. F. Pervez, M. F. Rahman, J. Hossain, J.
Comput. Electron. 19(1), 342 (2020); https://doi.org/10.1007/s10825-019-01433-0

You might also like