DOI:
10.1039/D4YA00323C
(Paper)
Energy Adv., 2024,
3, 2377-2398
Numerical modeling and extensive analysis of an extremely efficient RbGeI3-based perovskite solar cell by incorporating a variety of ETL and HTL materials to enhance PV performance
Received
22nd May 2024
, Accepted 26th July 2024
First published on 8th August 2024
Abstract
The immense demand for electrical energy motivated us to manipulate solar energy by means of conversion through solar cells (SCs). Advancements in photovoltaic (PV) technology are occurring very rapidly. In recent years, extensive research has been conducted on halide perovskite-based SCs because of their superior optoelectronic properties, enhanced efficiency, lightweight nature, and low cost. However, concerns have arisen regarding their longevity, stability, and commerciality due to the presence of toxic lead (Pb). The most prominent purpose of this investigation is to discover additional efficient, sustainable, and eco-friendly device architectures. In this study, we investigated an all-inorganic, lead-free rubidium germanium iodide (RbGeI3)-based PSC device with the assistance of the SCAPS-1D simulator. Several electron transport layers (ETLs) and hole transport layers (HTLs) were incorporated with the perovskite layer, and an efficient primary structure was discovered. Then, the impact of temperature; back metal work function; series and shunt resistance; surface recombination velocity of carriers; thickness of the perovskite absorber layer, electron transport material (ETM), and hole transport material (HTM); carrier concentration of the perovskite absorber layer, ETM, and HTM; defect density of the perovskite absorber layer, ETM, and HTM; and defect density of the HTL/absorber and absorber/ETL interfaces on the PV performance of the proposed PSC device was analyzed. The optimized device exhibited a power conversion efficiency (PCE) of 30.35%, with superior values for open circuit voltage (Voc), short circuit current density (Jsc), and fill factor (FF) of 1.067 V, 33.15 mA cm−2, and 85.82%, respectively. The investigations in this study may be valuable and impactful to solar cell material researchers and move the research interest forward by one step so that experimental work with non-toxic RbGeI3-based PSC devices will be performed in the future.
1. Introduction
Due to extensive population growth, the technological improvement of civilization, and the progression of developing countries worldwide, the increasing demands for electronic devices as well as electricity are extensively increasing the consumption and utilization of global energy.1–3 It has been predicted that in 2050, the demand for the consumption of energy will reach a level of about 30 terawatts (TW).1,4,5 This substantial need for energy is primarily met by the use of fossil fuels, which continue to account for more than 80% of the global energy system.6–8 Due to the burning of fossil fuels, a tremendous amount of CO2 has been exhausted and has contaminated the ecosystem, causing climate change, global warming, and the greenhouse effect.9–13
The current challenge in photovoltaics is to fabricate eco-friendly solar cells with a higher value of efficiency for the conversion of photon energy, as well as longevity, stability, commerciality, and cost-effectiveness.14 There are several types of solar cells (SCs) that have technologically evolved, including silicon (Si),15 cadmium telluride (CdTe),16 antimony selenide (Sb2Se3),17 molybdenum sulfide (MoS2),1 copper indium gallium selenide (CIGS),18,19 copper zinc tin sulfide (CZTS),20 copper iron tin sulfide (CFTS),21,22 copper indium telluride (CuInTe2),23 polymer, inorganic metal chalcogenide, quantum dot (QD), dye-sensitized solar cells (DSSCs), and perovskite-based solar cells.24–26
Among the different types of PV materials, organic–inorganic (hybrid) halides exhibit various remarkable properties such as a high absorption coefficient, simple synthesis, adjustable band gap, a larger value of diffusion length, efficient solution processing ability, and low cost, and they are desirable materials because of these remarkable properties.24,25 The first innovation of organic–inorganic halide perovskite-based solar cells (PSCs) was in 2009, with a power conversion efficiency (PCE) of 3.8% but only a few minutes of stability.27 Conversely, the presence of Pb (a hazardous element in nature) in perovskite and concern regarding long-term stability are the biggest impediments to its commercialization.
To overcome these drawbacks and find a replacement for Pb, researchers have explored metal cations such as Sn2+ or Ge2+, which are divalent and have a +2-level of oxidation and a similar configuration of the outermost shell that is comparable with Pb2+.28 The ionic radius of Sn2+ and Ge2+ is 1.35 Å and 0.73 Å, respectively, and that of Pb2+ is 1.49 Å. As a result, when Sn2+ and Ge2+ are used as divalent cations to substitute for Pb2+, the perovskite crystal structure is not affected.24,29 The oxidization of Sn2+ to Sn4+ occurs more easily because the Ge2+ ionic radius is less than that of Sn2+ and Pb2+. Thus, Ge-based perovskite exhibits greater conductivity than Pb-based and Sn-based perovskite. According to the periodic table, a comparison of the alkali metals Rb and Cs indicates that the reactivity of Rb is lower than that of Cs (i.e., Rb < Cs).24
Furthermore, RbGeI3 possesses various promising properties such as a lower band gap of 1.31 eV, an electron affinity of 3.9 eV, a higher dielectric permittivity of 23.1 eV, and a higher electron and hole mobility of 28.6 cm2 V−1 s−1 and 27.3 cm2 V−1 s−1, respectively. Therefore, the choice of RbGeI3 material as an absorber is much preferable.
In 1989, the thermometric characteristics of RbGeI3 were explored via XRD and Raman spectroscopy by Theile et al.30 In 2008, long-range-corrected density functional study for dye-sensitized solar cells was performed by Wong et al.31 In 2019, different characteristics, including magnetic, electrical, and structural, of RbGeI3-based (cubic) and rubidium dysprosium oxide (RbDyO3-based) PSCs were explored by Khursheed et al.32 In 2019, a first-principles study examining various characteristics, such as structural, electrical, and optical, of RbGeI3-based PSCs was conducted, and several characteristics such as relative permittivity at static (ε0), relative permittivity at high frequency (ε∞), hole and electron effective masses energy band gap, lattice parameters, and tolerance factor (0.90) were evaluated via the PBEsol function, as explored by Jong et al.33 In 2021, Jayan et al. investigated the optoelectronic, thermoelectric, thermodynamic, mechanical, and structural properties of RbGeI3-based (cubic Pmm space group, three-dimensional structure) PSCs to determine different exchange correlation functions.34 In 2022, Pindolia et al. researched the impact of various types of hole transport layers (HTLs) and electron transport layers (ETLs) on RbGeI3 as the absorber layer.25 In 2023, a theoretical maximum power conversion efficiency of 23.8% was achieved by Sarkar et al. through RbGeI3 simulation using the solar cell capacitance simulator-one-dimensional (SCAPS-1D) simulator.24
One of the most popular techniques for ab initio calculations of atomic, molecular, crystallographic, and surface structures and their interactions is density functional theory (DFT). Also, the material's band gap, and electronic, magnetic, and optical properties can be estimated using DFT calculations. In 2024, a DFT study of RbGeI3 was performed by Qin et al. to discover its excellent mechanical stability and resilience as well as its performance, which was evaluated using SCAPS-1D simulation software.35 The novelty of this research includes the determination of a new structure with all inorganic materials through many trials with various layer configurations for enhancing the power conversion efficiency, and also the use of low-cost, non-toxic, earth-abundant materials for different layer materials.
In this study, RbGeI3-based solar cells with various HTLs and ETLs were simulated to propose an improved cell structure and discuss the potential of this material. The properties of these materials are derived from previously published articles. The theoretical evolution of 19 different combinations was explored by incorporating ten inorganic HTMs (BaSi2, AgInTe2, Cu2Te, CdSe, SnS, CuO, Sb2Se3, MoS2, CuSCN, and Sb2S3), as well as ten inorganic ETMs (In2S3, In3Se4, CdS, WS2, SnS2, IGZO, ZnSe, In2Se3, Cd0.5Zn0.5S, and TiO2) with the RbGeI3-based absorber. After ascertaining the configurations with the greatest potential from 19 structures, we further investigated to optimize several properties such as temperature; work function of the left metal contact; series and shunt resistance; surface recombination velocity of electrons and holes; thickness of the perovskite-based absorber layer, ETM, and HTM; carrier concentration of the perovskite absorber layer, ETM, and HTM; defect density of the perovskite-based absorber layer, ETM, and HTM; and defect density at the interface of the HTL/absorber and absorber/ETL. Additionally, the C–V attribution and Mott–Schottky plot were also investigated. Finally, a comparison with earlier research was performed using the obtained solar cell parameters.
2. Simulation process, material specifications, and device-designing methodology
2.1 Simulation process and material specifications
The numerical simulation method is crucial for quickly comprehending the physical characteristics and functionality of PV devices as well as the behavior of each device parameter without expending a great deal of money and time. There are numerous simulation tools currently available, including PC1D, AFORS-HET, ATLAS, Sentaurus TCAD, AMPS-1D, wxAMPS, and SCAPS-1D, which are frequently utilized to build and assess a photovoltaic device's performance.36–39 Notably, SCPAS-1D demonstrated significant potential for modeling and simulating various PV device structures with flexible parameters involving spectral response, carrier generation and recombination mechanism, capacitance and voltage relation, capacitance and frequency relation, working temperature, series resistance and shunt resistance, metal contacts, and precisely calculated Voc, Jsc, fill factor (FF), and PCE through numerical convergence and rapid single and batch calculations with an intuitive user interface.24,40 Thus, the present investigation was conducted using the SCAPS-1D simulation tool.
Conversely, there are several restrictions on this tool. One of SCAPS-1D's restrictions is that it can only operate in one dimension. Additional drawbacks of the SCAPS-1D simulator are that it has a layer limit of seven, there is a volatile interpretation for a secondary barrier or n–p (rather than p–n) junction, and error in divergence occurs when the number of steps in simulation is limitless.24,40 SCAPS-1D is a useful tool that was created by the Department of electronics and information systems (EIS) at the University of Gent in Belgium. Alex Niemegeers, Marc Burgelman, Koen Decock, Stefaan Degrave, and Johan Verschraegen are the researchers that have been involved in its development. The application was initially created for the CuInSe2 and CdTe families of cell architectures.
At present, this program is also practicable for amorphous-type silicon (a-Si) cells, crystalline-type silicon (c-Si) SCs, and the GaAs family.24,41–43 Additionally, SCAPS is a powerful tool for analyzing semiconductor equations (eqn (1)–(3)). By accounting for the boundary conditions, SCAPS-1D can solve Poisson's equation (eqn (1)), as well as the electron and hole continuity equations given in eqn (2) and (3), respectively, at every location within the device.44–47 SCAPS-1D analyzes the operation of the photovoltaic devices by taking into account the Shockley–Read–Hall (SRH) recombination statistics:42
| | (1) |
where
E(
x) denotes electrostatic potential;
e denotes the charge of an electron;
ε0 and
εr denote vacuum and relative permittivity, respectively;
n(
x) and
p(
x) denote the concentration of electrons and holes, respectively;
NA and
ND denote the charge density of acceptor and donor, respectively; and
ρn and
ρp denote the distribution of electrons and holes, respectively.
| | (2) |
| | (3) |
where
Jn and
Jp denote the electron and hole current density, respectively, and
G and
R denote the generation rate and recombination rate, respectively.
In semiconductors, carrier transport originates from drift and diffusion, and can be described as eqn (4) and (5).
| | (4) |
| | (5) |
where
Dn and
Dp denote the electron and hole diffusion coefficients, respectively, and
μn and
μp denote the electron and hole mobility, respectively.
Table 1 shows the values of several specifications for separate layers of the proposed PSC device.
Table 2 shows the value of defect density for different interfaces.
Table 3 shows the values of various electrical characteristics for the front metal contact, as well as the back metal contact.
Table 4 shows the values of various specifications for several HTL materials.
Table 5 shows the values of various specifications for several ETL materials.
Table 1 Input specifications for the simulation of the proposed PSC devicea
Specifications |
ITO42 |
TiO224 |
RbGeI324 |
Sb2S344–46 |
N.B.: CB = conduction band, VB = valence band.
|
Thickness, L (nm) |
50 |
10–100 |
100–1000 |
10–200 |
Bandgap, Eg (eV) |
3.65 |
3.2 |
1.31 |
1.62 |
Electron affinity, χ (eV) |
4 |
4 |
3.9 |
3.7 |
Dielectric permittivity (relative), εr (eV) |
8.9 |
9 |
23.1 |
7.08 |
CB effective density of states, NC (1 cm−3) |
5.2 × 1018 |
2.0 × 1018 |
2.8 × 1019 |
2.0 × 1019 |
VB effective density of states, NV (1 cm−3) |
1 × 1018 |
1.8 × 1019 |
1.4 × 1019 |
1 × 1019 |
Electron thermal velocity, vT,n (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1.7 × 107 |
Hole thermal velocity, vT,p (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1.4 × 107 |
Electron mobility, μn (cm2 V−1 s−1) |
10 |
20 |
28.6 |
9.8 |
Hole mobility, μp (cm2 V−1 s−1) |
10 |
10 |
27.3 |
10 |
Shallow uniform donor density, ND (1 cm−3) |
1 × 1018 |
1012–1019 |
— |
— |
Shallow uniform acceptor density, NA (1 cm−3) |
— |
— |
1012–1021 |
1012–1021 |
Defect type |
Single acceptor |
Single acceptor |
Single donor |
Single donor |
Capture cross section (electrons), σn (cm2) |
1 × 10−15 |
1 × 10−15 |
1 × 10−15 |
1 × 10−15 |
Capture cross section (holes), σp (cm2) |
1 × 10−15 |
1 × 10−15 |
1 × 10−15 |
1 × 10−15 |
Energetic distribution |
Uniform |
Uniform |
Uniform |
Uniform |
Energy level with respect to reference (eV) |
0.600 |
0.600 |
0.600 |
0.600 |
Characteristic energy (eV) |
0.100 |
0.100 |
0.100 |
0.100 |
Total defect density, Nt (1 cm−3) |
1 × 1014 |
1012–1018 |
1012–1018 |
1012–1021 |
Series resistance, Rs (Ω cm2) |
1 |
Shunt resistance, Rsh (Ω cm2) |
1 × 105 |
Table 2 Input specifications for the defect density of different interfaces
Specifications |
RbGeI3/TiO2 interface |
Sb2S3/RbGeI3 interface |
Defect type |
Neutral |
Neutral |
Capture cross section (electrons), σn (cm2) |
1 × 10−19 |
1 × 10−19 |
Capture cross section (holes), σp (cm2) |
1 × 10−19 |
1 × 10−19 |
Energetic distribution |
Single |
Single |
Energy level with respect to reference (eV) |
0.600 |
0.600 |
Characteristic energy (eV) |
0.100 |
0.100 |
Total defect density, nt (1 cm−2) |
1012–1019 |
1012–1021 |
Table 3 Input specifications for the front metal and back metal
Specifications |
Front/right contact48 |
Back/left contact49 |
Material |
Aluminum (Al) |
Platinum (Pt) |
Crystal lattice and orientation |
FCC and 110 |
331 |
Surface recombination velocity of electrons (cm s−1) |
1 × 107 |
1 × 105 |
Surface recombination velocity of holes (cm s−1) |
1 × 105 |
1 × 107 |
Work function, Φ (eV) |
4.06 |
5.12 |
Table 4 Input specifications for the simulation of several HTL materials
Specifications |
BaSi250 |
AgInTe217 |
Cu2Te51 |
CdSe16 |
SnS52 |
CuO53 |
Sb2Se316 |
MoS214,54 |
CuSCN25 |
L (nm) |
50 |
50 |
50 |
50 |
50 |
50 |
50 |
50 |
50 |
E
g (eV) |
1.3 |
1.16 |
1.19 |
1.7 |
1.6 |
1.51 |
1.53 |
1.7 |
3.4 |
χ (eV) |
3.3 |
3.6 |
4.1 |
3.93 |
3.6 |
4.07 |
4.04 |
3.8 |
1.7 |
ε
r (eV) |
11.17 |
8.9 |
10 |
9.5 |
12.5 |
18.1 |
18 |
13.6 |
10 |
N
C (1 cm−3) |
2.6 × 1019 |
3.66 × 1019 |
7.8 × 1017 |
2.8 × 1019 |
7.5 × 1018 |
2.2 × 1019 |
2.2 × 1018 |
2.8 × 1019 |
2.2 × 1019 |
N
V (1 cm−3) |
2 × 1019 |
1.35 × 1019 |
1.8 × 1019 |
1.2 × 1019 |
1.0 × 1019 |
5.5 × 1020 |
1.8 × 1019 |
1 × 1019 |
1.8 × 1018 |
v
T,n (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
v
T,p (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
μ
n (cm2 V−1 s−1) |
820 |
1011 |
500 |
5.93 |
100 |
100 |
15 |
12 |
100 |
μ
p (cm2 V−1 s−1) |
100 |
887 |
100 |
25 |
4 |
0.1 |
5.1 |
2.8 |
25 |
N
A (1 cm−3) |
1 × 1019 |
1 × 1019 |
1 × 1019 |
1 × 1019 |
1 × 1019 |
1 × 1019 |
1 × 1019 |
1 × 1019 |
1 × 1019 |
Table 5 Input specifications for the simulation for several ETL materials
Specifications |
In2S355 |
In3Se456 |
CdS50 |
WS225 |
SnS257 |
IGZO25 |
ZnSe17 |
In2Se358 |
Cd0.5Zn0.5S54 |
L (nm) |
50 |
50 |
50 |
50 |
50 |
50 |
50 |
50 |
50 |
E
g (eV) |
2.8 |
1.8 |
2.42 |
1.87 |
2.24 |
3.05 |
2.7 |
2.4 |
2.8 |
χ (eV) |
4.6 |
4.55 |
4.4 |
4.3 |
4.24 |
4.16 |
4.09 |
3.8 |
4 |
ε
r (eV) |
13.5 |
5.54 |
10 |
11.9 |
10 |
10 |
10 |
10 |
10 |
N
C (1 cm−3) |
2.2 × 1017 |
1.0 × 1018 |
2.2 × 1018 |
2.4 × 1019 |
2.2 × 1017 |
5.0 × 1018 |
1.5 × 1018 |
2.2 × 1018 |
1.0 × 1018 |
N
V (1 cm−3) |
1.8 × 1019 |
1.0 × 1018 |
1.8 × 1019 |
1.0 × 1019 |
1.8 × 1019 |
5.0 × 1018 |
1.8 × 1019 |
1.8 × 1019 |
1.0 × 1018 |
v
T,n (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
v
T,p (cm s−1) |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
1 × 107 |
μ
n (cm2 V−1 s−1) |
50 |
100 |
100 |
260 |
50 |
15 |
50 |
50 |
100 |
μ
p (cm2 V−1 s−1) |
20 |
50 |
25 |
51 |
50 |
0.1 |
20 |
12 |
25 |
N
A (1 cm−3) |
1 × 1015 |
1 × 1015 |
1 × 1015 |
1 × 1015 |
1 × 1015 |
1 × 1015 |
1 × 1015 |
1 × 1015 |
1 × 1015 |
2.2 Structure and design process for the device
PV device performance is generally evaluated by four performance specifications: Jsc, Voc, FF, and PCE. The spectrum of the incident light (the global air mass ‘AM1.5G, 1 Sun’ spectrum at the temperature of 300 K was used in this investigation) and the optical characteristics (recombination, reflection, and absorption) affect the short circuit current density (Jsc). It can be evaluated using the formula in eqn (6):59 | | (6) |
where q denotes the charge of an electron, T(λ) indicates optical transmission, ϕi indicates spectral power density, and Δλi indicates the interval between the two subsequent values of wavelength.
When there is no net current flowing through the solar cell, open circuit voltage (Voc) is created and can be calculated using the formula in eqn (7):60
| | (7) |
where
n,
k,
T,
Jsc, and
Jo denote the diode ideality factor, Boltzmann constant, ambient temperature, short circuit current density, and reverse bias saturation current density, respectively.
The solar cell's maximum power is determined by the fill factor (FF) and is stated in eqn (8):61
| | (8) |
where
is denoted as normalized
Voc.
The efficiency for the photon's power conversion (PCE) is determined by eqn (9):60
| | (9) |
Herein,
Pin denotes the input photon's power from the sun (1000 W m
−2 was used in this investigation).
To finalize the structure of a PV device, a study was performed to obtain additional information regarding the performance of the device with a combination of several HTLs and several ETLs. Unfortunately, charge transport materials (CTMs) consist of organic compounds that are very unstable.62–65 Additionally, because of their simplest manufacturing technique, inorganic CTMs are inexpensive. Inorganic CTMs exhibit greater thermodynamical and chemical durability than organic CTMs. In comparison with organic CTMs, inorganic CTMs exhibit a wider band gap, greater mobility of the charge carrier, and superior transparency to the different radiations of visible, infrared, and ultraviolet.66–69
Therefore, in this investigation, to ascertain the primary structure of the PSC device, a PSC device based on inorganic RbGeI3 was investigated by incorporating ten inorganic HTMs (BaSi2, AgInTe2, Cu2Te, CdSe, SnS, CuO, Sb2Se3, MoS2, CuSCN, Sb2S3) as well as ten inorganic ETMs (In2S3, In3Se4, CdS, WS2, SnS2, IGZO, ZnSe, In2Se3, Cd0.5Zn0.5S, TiO2). The values of thickness and concentration of the charge carrier of the RbGeI3-based absorber layer were initially set as 800 nm and 1016 cm−3 respectively, and that of each HTL as 50 nm and 1 × 1019 cm−3, and each ETL as 50 nm and 1 × 1015 cm−3, respectively, while other specifications remained the same as in Tables 1–3. Simulation was performed for different configurations with the combinations of various HTLs and ETLs, and all the primary data on performance parameters (such as Voc, Jsc, FF, and PCE) were recorded.
Table 6 shows the data for four performance parameters, and the maximum device performance was obtained when Sb2S3 and TiO2 were individually applied as HTM and ETM. Also, Sb2S3 possesses superior optoelectronic properties such as an appropriate band gap (∼1.62 eV), a superior absorption coefficient (∼105 cm−1), ingredient abundance on earth, minimal toxicity, and a low-cost production procedure.52 Additionally, the valence band offset (VBO) of Sb2S3 and the conduction band offset (CBO) of TiO2 are much lower in comparison with other HTLs and ETLs.
Table 6 Recorded performance of the PSC device for different configurations through numerical simulation
Configuration |
V
oc (V) |
J
sc (mA cm−2) |
FF (%) |
PCE (%) |
Al/ITO/TiO2/RbGeI3/BaSi2/Pt |
0.5915 |
32.115889 |
66.48 |
12.63 |
Al/ITO/TiO2/RbGeI3/AgInTe2/Pt |
0.7508 |
33.586818 |
71.52 |
18.04 |
Al/ITO/TiO2/RbGeI3/Cu2Te/Pt |
0.8468 |
34.061598 |
82.04 |
23.66 |
Al/ITO/TiO2/RbGeI3/CdSe/Pt |
0.8976 |
34.091434 |
79.61 |
24.36 |
Al/ITO/TiO2/RbGeI3/SnS/Pt |
0.8974 |
34.096428 |
81.29 |
24.87 |
Al/ITO/TiO2/RbGeI3/CuO/Pt |
0.8918 |
34.105198 |
81.89 |
24.91 |
Al/ITO/TiO2/RbGeI3/Sb2Se3/Pt |
0.9003 |
34.105396 |
81.78 |
25.11 |
Al/ITO/TiO2/RbGeI3/MoS2/Pt |
0.9003 |
34.097394 |
82.04 |
25.18 |
Al/ITO/TiO2/RbGeI3/CuSCN/Pt |
0.8998 |
34.092787 |
82.10 |
25.19 |
Al/ITO/TiO
2
/RbGeI
3
/Sb
2
S
3
/Pt
|
0.9002
|
34.100013
|
82.10
|
25.20
|
Al/ITO/Cd0.5Zn0.5S/RbGeI3/Sb2S3/Pt |
0.8970 |
34.085439 |
81.21 |
24.83 |
Al/ITO/In2Se3/RbGeI3/Sb2S3/Pt |
0.8970 |
34.085248 |
80.88 |
24.73 |
Al/ITO/ZnSe/RbGeI3/Sb2S3/Pt |
0.8948 |
34.087093 |
80.95 |
24.69 |
Al/ITO/IGZO/RbGeI3/Sb2S3/Pt |
0.8715 |
34.089026 |
77.38 |
22.99 |
Al/ITO/SnS2/RbGeI3/Sb2S3/Pt |
0.8717 |
34.088865 |
77.34 |
22.98 |
Al/ITO/WS2/RbGeI3/Sb2S3/Pt |
0.7150 |
34.093955 |
71.79 |
17.50 |
Al/ITO/CdS/RbGeI3/Sb2S3/Pt |
0.6773 |
34.098321 |
70.81 |
16.35 |
Al/ITO/In3Se4/RbGeI3/Sb2S3/Pt |
0.5612 |
33.243283 |
42.62 |
7.95 |
Al/ITO/In2S3/RbGeI3/Sb2S3/Pt |
0.5424 |
3.074529 |
42.58 |
0.71 |
After numerous trials, the Al/ITO/TiO2/RbGeI3/Sb2S3/Pt structure proved to be an effective primary structure. Hence, further investigation, i.e., optimization of several properties such as temperature; work function of the left metal contact; series and shunt resistance; surface recombination velocity of electrons and holes; thickness of the perovskite-based absorber layer, ETM, and HTM; carrier concentration of the perovskite absorber layer, ETM, and HTM; defect density of the perovskite-based absorber layer, ETM, and HTM; and defect density at the interface of HTL and the absorber layer, and also the absorber layer and ETL, were conducted with the structure of the PSC device as Al/ITO/TiO2/RbGeI3/Sb2S3/Pt.
2.3 Finalized structure of the device
Fig. 1(a) shows the finalized and optimized structure (Al/ITO/TiO2/RbGeI3/Sb2S3/Pt) of the proposed perovskite-based PV device. It was determined that p-type RbGeI3 was the most attractive Pb-free, highly stable, efficient, and low-cost perovskite-based absorber material. On the top of the absorber, n-type TiO2 ETL was used as a heterojunction associated with the RbGeI3 absorber. It formed an n–p heterojunction interface that enhanced the collection of photogenerated electrons (PGEs) with blocking holes. On top of the TiO2 layer, highly transparent, conductive, and low-cost n+-type ITO was applied as a window layer. At the bottom of the absorber (RbGeI3), a p+-type Sb2S3 HTL was used to effectively collect photogenerated holes (PGHs) with blocking electrons that enhanced the overall efficiency of the proposed PSC. Aluminum (Al) was used as the front metal contact material, and platinum (Pt) was used as the back metal contact material. Fig. 1(b) shows the energy band diagram of the suggested device, which shows that there is a low VBO and adequately large CBO at the Sb2S3/RbGeI3 interface, and the band configuration between Sb2S3 and RbGeI3 is excellent for ensuring that hole transmission from the absorber as well as electron opposition are facilitated. Also, because of the adequately large VBO and very low CBO at the RbGeI3/TiO2 interface, the band configuration between RbGeI3 and TiO2 is excellent for ensuring smooth electron transmission from the absorber and facile hole opposition.
|
| Fig. 1 (a) Schematic diagram and (b) energy band diagram of the proposed PSC device. | |
3. Results and discussion
3.1 Consequences of ETL thickness and carrier concentrations on the performance of the device and their optimization
There is considerable impact of the ETL's (TiO2) carrier concentration as well as thickness on PSC performance, and they can be used to improve it. Therefore, when developing highly efficient PSCs, it is important to carefully select the ETL's characteristics. An optimal ETL for PSCs can assist in decreasing recombination currents and increasing transmittance.70 In this investigation, the ETL's carrier concentration and thickness were varied consecutively within the range of 1012–1019 cm−3 and 10–100 nm, while other specifications remained the same, as in Tables 1–3.
The variation in the device's performance according to the ETL's thickness is demonstrated in Fig. 2(a), which shows that all the PV performance parameters, i.e., Voc, Jsc, FF, and PCE, remain unchanged for ‘with HTL’ and ‘without HTL.’ All the PV parameters have a higher value ‘with HTL’ than ‘without HTL.’ Because there is no significant alteration in the PSC performance as the ETL's thickness is varied, 50 nm was selected as the optimal thickness for the ETL (TiO2) layer to minimize fabrication costs.
|
| Fig. 2 Consequences of the ETL (TiO2) layer's (a) thickness and (b) carrier concentration on the device's performance. | |
The variation in the device's performance according to the ETL's carrier concentration is demonstrated in Fig. 2(b). Although there were lower values without HTL than with HTL for all the PV parameters, they remained nearly unchanged, except for FF as the carrier concentration, which was varied. The FF increased very slightly beyond 1016 cm−3. Because this occurred as doping concentrations increased, they also raised the internal electric field of the ETL, thus increasing the transport velocity of the charge carriers and thereby conductivity.71 The PV parameters have a higher value with HTL than without HTL, and they remained unchanged up to 1018 cm−3. After that, Voc, Jsc, and PCE slightly decreased, but FF significantly decreased. The reason behind this is that coulomb traps resulted due to high doping concentrations, which in turn decreased the electron mobility.72 Therefore, it was determined that the optimal carrier concentration for the ETL (TiO2) layer was 1017 cm−3.
3.2 Consequences of the absorber's thickness and carrier concentrations on the device's performance and their optimization
The perovskite layer's thickness and carrier concentration are critical elements for improving the PSC's performance, and they influence diffusion length, photocarrier generation, carrier lifetime, and extraction. As the absorber layer thickness increases, the generation and recombination rates of the carriers also increase.73 In this investigation, the absorber layer's thickness and carrier concentration varied with the range (100–1000) nm and (1012–1021) cm−3 respectively, while other specifications remained the same as in Tables 1–3. The alteration in the device's performance according to absorber thickness is demonstrated in Fig. 3(a), where the open circuit voltage (Voc) significantly increased up to 500 nm for the ‘without HTL’ configuration, and after that, it slightly increased. Although the Voc for the ‘with HTL’ configuration is higher than that ‘without HTL,’ it slightly diminished with the initial increasing thickness. The reason behind this is that the recombination increases with the increase in absorber thickness, and hence the saturation current also increases.72 The current density (Jsc) significantly increased up to 600 nm with the increase in thickness of the absorber layer. The significant increase in photon absorption is the primary cause of the rise of Jsc.72 Beyond the 600 nm thickness of the absorber, the current density also increased, but not significantly in comparison. The reasons are that the photon absorption also increases, but not significantly, and the greater absorber layer influences charge carrier diffusion lengths, enhancing photogenerated carrier recombination via SRH recombination.74 The fill factor (FF) for ‘without HTL’ significantly increased to 200 nm and that is in the opposite phase for “with HTL”, and after that, the change in FF for both “with HTL” and “without HTL” is slight but in the opposite phase. In the case of “with HTL” structures, the reason for abrupt decreases in FF is that the series resistance (Rs) is growing as the thickness of the cell rises.75,76 However, the FF is boosted in the ‘with HTL’ configuration rather than ‘without HTL.’ Because the PCE is determined by the parameters Voc, Jsc, and FF as presented in eqn (9). Although Voc and FF are slightly decreased for ‘with HTL,’ as a result of a significant increase in Jsc up to 600 nm thicknesses, the PCE significantly intensified up to 600 nm. The PCE (without HTL) increased due to the same reason but is lower than ‘with HTL.’ Therefore, it was determined that 600 nm was the optimal value of the thickness for the absorber (RbGeI3), and this value was used to reduce fabrication costs.
|
| Fig. 3 Consequences of the absorber (RbGeI3) layer's (a) thickness and (b) carrier density on the device's performance. | |
The device's performance varied according to the absorber's carrier density, as demonstrated in Fig. 3(b). According to Fig. 3(b), the Voc remained unchanged with an absorber carrier concentration of up to 1016 cm−3 for the ‘with HTL’ and ‘without HTL’ structures, and after that, it significantly increased. The key reason for this is because as the concentration of the carrier increases, the saturation current of the entire cell also increases, which reinforces the increase in Voc.21 Although the Voc is lower for ‘without HTL’ than ‘with HTL,’ the Jsc remained unchanged with the variation of the carrier concentration for the ‘with HTL” structure. Thus, there is no substantial alteration of photon absorption within the variation of the carrier concentrations, and that is the principal reason for a lack of change in Jsc. The carrier concentration accelerates the recombination process by incorporating traps or recombination centers into the layer. At the front, the collection of the photogenerated carriers is then limited by this recombination process.1,77 The increase in recombination rate may be the same as the increasing rate of Jsc due to the increase in carrier concentration, and hence, no significant change in Jsc occurred. However, for the ‘without HTL’ structure, the Jsc is the same as that for the ‘with HTL’ structure. Up to 1016 cm−3 and beyond that, it rapidly decreased up to 1018 cm−3, and after that, it gradually decreased. This occurred because the recombination rate may be higher due to the rapid decrease in Jsc. The FF for the ‘with HTL’ configuration remained unchanged up to 1015 cm−3, but after that, it significantly increased up to 1017 cm−3, and beyond that, it slightly increased. The FF for ‘without HTL’ is less than that for the ‘with HTL’ configuration and remained unchanged up to 1016 cm−3; after that, it slightly decreased up to 1017 cm−3, and beyond that, it significantly increased. The fact behind this is that the increase in Voc is due to an increase in carrier concentration. Because the PCE is determined by the parameters Voc, Jsc, and FF, as presented in eqn (9), the Jsc remained unchanged for the ‘with HTL’ configuration, but as a result of a major increase in Voc as well as FF, the PCE then significantly increased up to 1019 cm−3. However, the PCE for ‘without HTL’ significantly decreased due to a considerable diminution of Jsc, although there was a slight increase in Voc and FF. The optimal value of the carrier concentration for the absorber (RbGeI3) layer was therefore taken as 1019 cm−3.
3.3 Consequences of the HTL's thickness and carrier concentrations on device performance and their optimization
The HTL's (Sb2S3) thickness and the concentration of the carrier have the ability to greatly influence the PSC's performance, and can improve it. The HTL increased the SC effectiveness due to minimization of the surface recombination velocity (SRV) by producing a strong field near the rear electrode.1 To examine the consequences of concurrent changes in thickness and concentration of the carrier of the HTL (Sb2S3) on the PSC device performance, a contour map was constructed with simultaneous variation of thickness and density of the carrier of the HTL (Sb2S3) within the range of 10–200 nm and 1014–1021 cm−3, respectively, while other specifications remained the same as in Tables 1–3. Fig. 4 illustrates the influence of simultaneous variation of thickness and concentration of the carrier of the HTL on the parameters that determine the performance of the PV devices.
|
| Fig. 4 Consequences to PV performance parameters (a) Voc, (b) Jsc, (c) FF, and (d) PCE due to concurrent changes in thickness and density of the HTL carrier (Sb2S3). | |
According to Fig. 4(a), for the range of the variation of carrier density from 1014 cm−3 to 8 × 1020 cm−3, the Voc increased very negligibly within the increase of thickness for the range of 10 nm to 60 nm, and there was no alteration in the Voc after 60 nm. Beyond 8 × 1020 cm−3, the change in thickness did not considerably impact the Voc of the proposed PSC device. Also, for a particular thickness, the Voc was nearly unchanged for the range of 1014 cm−3 to 8 × 1020 cm−3, and the Voc slightly increased beyond 8 × 1020 cm−3. The reason for this negligible increase in Voc is that the Sb2S3 HTL produced an advanced built-in potential at the Sb2S3/RbGeI3 interface,56,78 as well as an increase in Jsc.79
Fig. 4(b) shows a very negligible increase in Jsc within the increase in thickness from 10 nm to 200 nm for the range of carrier concentrations of 1014 cm−3 to 8 × 1020 cm−3, and the Jsc slightly increased beyond 8 × 1020 cm−3 within the increase in thickness. Also, for the range of thickness of 10–100 nm, Jsc was almost unchanged for the range of 1014 cm−3 to 8 × 1020 cm−3, and for further than 8 × 1020 cm−3, the Jsc slightly increased for the range of thickness of 100–200 nm. Photon absorption at longer wavelength spectra results in a slight increase in electron–hole pair generation through the tail-states-assisted (TSA) two-step photon upconversion process, which is the main reason for the slight increase in Jsc.16,19,56
According to Fig. 4(c), for the range of 1014 cm−3 to 8 × 1020 cm−3 for carrier concentration, the FF increased very negligibly within the increase in thickness for the range of 10 nm to 30 nm, and afterwards, the FF remained nearly unaffected. Beyond 8 × 1020 cm−3, the variation in thickness had no impact on the FF of the proposed PSC device, and this occurred because Voc has no impact on the thickness variation. Additionally, for a particular thickness, the FF was nearly unchanged from 1014 cm−3 to 8 × 1020 cm−3. Away from 8 × 1020 cm−3, the FF slightly increased because higher carrier concentrations reduced the series resistance of the cell, hence grading the FF of the cell.80
From Fig. 4(d), it is clear that the PCE increased very negligibly for the range of carrier concentrations from 1014 cm−3 to 8 × 1020 cm−3 within the increase in thickness for the range of 10 nm to 60 nm, and afterward, the PCE remained nearly unaffected. Beyond 8 × 1020 cm−3, the variation in thickness had no impact on the conversion efficiency of the photon's power in the proposed PSC device. Additionally, for a particular thickness, the PCE was nearly unchanged for the range of 1014 cm−3 to 8 × 1020 cm−3, and outside of 8 × 1020 cm−3, the PCE slightly increased. This occurred because of all the combined effects of Voc, Jsc, and FF.
Sb2S3 with a small thickness cannot offer a contact channel of low resistance, whereas increased layer thickness prevents carriers from transferring to an electrode. Hence, extending the length of diffusion and allowing recombination to occur via a trade-off may be considered to determine the optimal value of the HTL's thickness, which was determined to be 50 nm. Also, the reasons behind the increase in overall performance are that increased carrier concentration moves the Fermi level towards the valence band (VB), which accelerates the hole build-up at the anode via creation of an ohmic contact with the rare metal contact,81 and also due to increased carrier concentration. The HTL's internal electric field also increases, which implies a higher transport velocity for the charge carriers, and consequently, the conductivity.72 The HTL's carrier concentration should be greater than the absorber layer's carrier concentration, and therefore, the optimal HTL carrier concentration was chosen as 1021 cm−3.
3.4 Consequences of the bulk defect density of different layers on device performance and their optimization
3.4.1 In the case of the ETL (TiO2).
In this investigation, the density of bulk defect (Nt) of the single-acceptor type for the n-type ETL (TiO2) was used. The defect density varied within the range of 1012–1018 cm−3, while other specifications remained the same, as in Tables 1–3. Fig. 5 illustrates the consequences of bulk defect density for the ETL, absorber, and HTL.
|
| Fig. 5 Consequences of the bulk defect density of the (a) ETL (TiO2), (b) absorber (RbGeI3), and (c) HTL (Sb2S3) on the device performance. | |
According to Fig. 5(a), for the ‘with HTL’ and ‘without HTL’ states, all the performance-determining parameters for the PV device, i.e., Voc, Jsc, FF, and PCE, were almost unchanged until 1017 cm−3. After that, they significantly decreased, with the exclusion of Voc. The density of bulk defect exhibits no influence on Voc. However, all the PV performance parameters for ‘with HTL’ have a higher value as compared to ‘without HTL.’ The recombination rate of SRH increased due to the contribution of the TiO2 layer's defect density, which lowered the SC's overall performance.82 The optimal value of the bulk defect density was therefore taken as 1017 cm−3 for the best performance.
3.4.2 In the case of the absorber (RbGeI3) layer.
The construction of the device or the grade of the material's light-absorbing capacity has little effect on the PCE of the PSCs because photoelectrons are produced when the sun's light strikes the absorber layer. Poor morphology results from insufficient perovskite layer distribution on the ETL. Because of the reduced film excellence, the defect density is higher, which results in greater recombination.76 In this investigation, the bulk defect density of the single-donor type for the p-type absorber (RbGeI3) layer was used. The defect density varied within the range of 1012–1018 cm−3, and all other specifications remained the same, as in Tables 1–3. Fig. 5(b) illustrates the consequences of the bulk defect density of the absorber layer on device performance.
According to Fig. 5(b), for the ‘without HTL’ configuration, all the performance-determining parameters of the PV device, i.e., Voc, Jsc, FF, and PCE, continued almost unchanged until 1016 cm−3, and after that, they significantly decreased. For the ‘with HTL’ configuration, Jsc remained almost unchanged up to 1015 cm−3, but after that, it significantly decreased up to 1016 cm−3, and beyond that, it dramatically degraded. Initially, the Voc slightly decreased to 1013 cm−3, and beyond that, it sharply degraded. The FF and PCE significantly decreased at startup to 1016 cm−3, and beyond that, they dramatically degraded. However, all the PV performance parameters for ‘with HTL’ had a higher value than that for ‘without HTL.’ The RbGeI3 absorber layer's defect density enhances the recombination rate of SRH, which reduces the number of PGCs and thereby reduces the values of Voc, Jsc, FF, and PCE.83 The optimal value of the bulk defect density for the RbGeI3 layer was therefore taken as 1014 cm−3 for best performance.
3.4.3 In the case of the HTL (Sb2S3).
In this investigation, the bulk defect density of the single-donor type for the p+-type HTL (Sb2S3) was used. The defect density varied within the range of 1012–1021 cm−3, and all other specifications remained the same, as in Tables 1–3. Fig. 5(c) illustrates the consequences of the bulk defect density of the HTL.
According to Fig. 5(c), all the performance-determining parameters of the PV device, i.e., Voc, Jsc, FF, and PCE, were strictly unchanged until 1020 cm−3 within the variation of the bulk defect density, and beyond that, they sharply decreased. This implies that there was no influence of the HTL's defect density up to 1020 cm−3 on the performance of the proposed PSC. The optimal value of the bulk defect density of the HTL was therefore taken as 1014 cm−3 for best performance.
3.5 Consequences of defect densities at different interfaces on the performance of the device and their optimization
3.5.1 In the case of the absorber/ETL interface.
In the heterojunction PV device, an interfacial defect develops during the fabrication process as a result of various structural imperfections. This implies that testing the consequences of interfacial imperfections on SC properties is urgent. To replicate more realistic conditions in this investigation, the interfacial defect (nt) of neutral type for the absorber/ETL interface was added. The defect density varied within the range of 1010–1018 cm−2, and all other specifications remained the same, as in Tables 1–3. Fig. 6 exemplifies the consequences of the defect density of the interface.
|
| Fig. 6 Consequences of the defect at the (a) absorber (RbGeI3)/ETL (TiO2) and (b) HTL (Sb2S3)/absorber (RbGeI3) interface on the device performance. | |
According to Fig. 6(a), for the ‘without HTL’ configuration, all the performance-determining parameters of the PV device, i.e., Voc, Jsc, FF, and PCE, were almost unchanged until 1016 cm−2; after that, they significantly decreased. For the ‘with HTL’ condition, Jsc and FF remained almost unchanged up to 1016 cm−2, and after that, they significantly decreased. Initially, the Voc slightly decreased to 1013 cm−2, and beyond that, it sharply degraded. The PCE remained almost unchanged up to 1013 cm−2, but after that, it significantly decreased up to 1016 cm−2, and beyond that, it dramatically degraded. The diminution in device performance could be due to the increase in defect density at the interface of the absorber/ETL. Also, electrons have a greater risk of being trapped, dispersed, or recombined while traveling from the absorber to the ETL, which reduces the charge carriers produced by the photon and hinders the accumulation of carriers.84 However, all the performance-determining parameters of the PV device for the ‘with HTL’ configuration have a higher value as compared to those ‘without HTL.’ The Voc is significantly more sensitive to nt than Jsc, and the nt range was preferred to 1012 cm−2 for Voc. The interface recombination Voc limit may be calculated using the formula in eqn (10):85
| | (10) |
where
St denotes interface recombination velocity,
A symbolizes the ideality factor of the hetero-junction,
k symbolizes the Boltzmann constant, and
ϕc denotes an effective barrier height.
3.5.2 In the case of the HTL/absorber interface.
To replicate more realistic conditions in this investigation, an interfacial defect density of neutral type for the HTL/absorber interface was added. The defect density varied within the range of 1010–1018 cm−2, while other specifications remained the same, as in Tables 1–3. Fig. 6(b) illustrates the consequences of interface defect density.
According to Fig. 6(b), for the ‘with HTL’ configuration, all the performance-determining parameters of the PV device, i.e., Voc, Jsc, FF, and PCE, continued almost unchanged until 1015 cm−2, excluding Jsc, which remained unchanged up to 1016 cm−2, and beyond that, it significantly decreased. The reason behind this is similar to that explained earlier in the above section. Therefore, the optimal density of the interface defect for the HTL/absorber interface was chosen at 1012 cm−2.
3.6 Consequences of different resistances on device performance and their optimization
The electrical resistances in the PSCs are introduced by the metal contacts used as front and back contacts in the perovskite absorbers, HTLs, and ETLs. The series (Rs) and shunt (Rsh) resistances play a key role in influencing the PV characteristics of SCs. Therefore, it is obligatory to understand the impacts of Rs and Rsh to improve the cell function. The Shockley equations determine the J–V properties of SCs when considering ideal one-sun illumination conditions, as in eqn (11) and (12).86,87 | | (11) |
where q, A, k, T, Jsc, Jph, Jo, V, Rs, and Rsh denote the elementary charge, ideality factor of the heterojunction, Boltzmann constant, ambient temperature, short circuit current density, photo-irradiated constant current density, reverse bias saturation current density, output voltage, and series and shunt resistances, respectively.
Because of an open-circuit condition, i.e., Jsc ≈ 0 mA cm−2, Voc is denoted as follows:
| | (12) |
3.6.1 In the case of series resistance.
In this investigation, series resistance variation was conducted in the range of (0–10) Ω cm2 to study the influence on the PV performance parameters, while other specifications remained the same, as in Tables 1–3. Fig. 7(a) illustrates the consequences of series resistance, and shows that for the ‘with HTL’ and ‘without HTL’ configurations, the series resistance has no effect on Jsc and Voc, but has a greater impact on FF and PCE. It is evident from the figure that FF significantly decreases from the beginning of the variation of Rs, and thereby PCE as well, which occurred because the increased series resistance accelerated the power loss. However, all the PV performance parameters for ‘with HTL,’ excluding FF, have a higher value as compared to ‘without HTL.’ Andriessen and colleagues found that the increased series resistance of electrodes caused a greater loss of conversion efficiency in PSC devices once the active regions were augmented.88 Although there are limitations to reducing series resistance due to the fabrication process and material's inherent properties, the optimal series resistance for the proposed PSC was chosen as 1 Ω cm2.
|
| Fig. 7 Consequences of (a) series resistance (Rs) and (b) shunt resistance (Rsh) on device performance. | |
3.6.2 In the case of shunt resistance.
Shunt resistance was introduced by several charge recombination routes in the PSC.72 In this investigation, the shunt resistance was varied within the range of 101–109 Ω cm2 to study its influence on the PV performance parameters, while other specifications remained the same, as in Tables 1–3. Fig. 7(b) illustrates the consequences of shunt resistance, and shows that for the ‘with HTL’ and ‘without HTL’ configuration, the Jsc and Voc sharply increase up to 102 Ω cm2, and beyond that, the shunt resistance has no effect on Jsc or Voc. It is also evident from the figure that the FF and PCE significantly increase from the beginning of the variation up to 103 Ω cm2, and after that, they remain strictly unchanged. The reason behind the grading in cell performance is that the increase in shunt resistance degrades current loss. For the proposed PSC, 105 Ω cm2 was chosen as the optimal shunt resistance.
3.7 Consequences of surface recombination velocity (SRV) on device performance and its optimization
The consequences of SRV at the back metal contact on the performance-determining parameters of the planned PSC device were thoroughly investigated, as seen in Fig. 8.
|
| Fig. 8 Consequence of the surface recombination velocity of (a) electrons and (b) holes on device performance. | |
3.7.1 In the case of the electron's SRV.
The SRV of electrons varied within the range of 101–1010 cm s−1, while other specifications remained the same, as in Tables 1–3. The influence on the performance-determining parameters by reason of the variation of electron SRV is illustrated in Fig. 8(a). According to Fig. 8(a), for the ‘without HTL’ configuration, all the performance-determining parameters for the PV device, i.e., Voc, Jsc, FF, and PCE, significantly degraded until 104 cm s−1. After that, there was no impact of electron SRV on the performance parameters. Before reaching metal contact, electron and hole recombination occurred with greater SRV, which implied a significant decrease in performance parameters. This drawback may be overcome by adding an HTL, which is visible in Fig. 8(a). This figure shows the evidence that there is no effect of electron SRV on the performance parameters for the ‘with HTL’ structure. Although the electrons have a higher SRV, they are prevented by the HTL from attaining back metal contact, and thus no electron–hole recombination occurs, resulting in no limitation on performance parameters.
3.7.2 In the case of the hole's SRV.
The SRV of holes varied within the range of 101–1010 cm s−1, while other specifications remained the same, as in Tables 1–3. Fig. 8(b) illustrates the influence on the performance-determining parameters by reason of the variation of hole SRV. According to Fig. 8(b), for the ‘without HTL’ configuration, all the performance-determining parameters of the PV device, i.e., Voc, Jsc, FF, and PCE, remain unchanged with the variation in the hole's SRV. However, for the ‘with HTL’ configuration, although there is no influence of hole SRV variation on Jsc and Voc, the FF significantly increased up to 103 cm s−1, and beyond that, it remained unchanged, which implies enhancement of the PCE. This occurred because holes are accelerated by the HTL and do not attain back metal contact, which thereby enhances the FF and PCE of the proposed PSC device.
3.8
C–V attribution and Mott–Schottky plot analysis of the device
3.8.1 Inspection of the C–V attribution and Mott–Schottky plot.
Capacitance–voltage (C–V) inspections were conducted at different frequencies, i.e., within the range of 0.5 kHz to 1 MHz, while other specifications remained the same, as in Tables 1–3, by applying the DC bias potential difference, as depicted in Fig. 9(a). The net capacitance generally appears from the depletion as well as diffusion capacitances linked with the p–n junctions for all the PV devices.
|
| Fig. 9 (a) C–V attribution curve and (b) (1/C2) − V (Mott–Schottky plot) curve of the device. | |
The depletion capacitance, i.e., space charge capacitance, appeared by reason of the reverse bias condition in the depletion region and in a similar manner to that of the parallel plate capacitor, whose capacitance is given by eqn (13):89
| | (13) |
where,
ε,
A, and
W denote the permittivity of the material, the area of the SC, and depletion width at the junction interface, respectively.
Conversely, the capacitance formed by diffusion appeared per unit area due to the forward bias condition, and this capacitance can be given by eqn (14):90
| | (14) |
where
q denotes the absolute value of electronic charge,
L denotes the layer's thickness for the accumulation of charge,
n0 denotes the concentration of the minority carrier at equilibrium,
η denotes the diode ideality factor,
k denotes the Boltzmann constant, and
T denotes the absolute temperature.
According to Fig. 9(a), the capacitance of the proposed PSC was nearly 78 nF cm−2 at zero bias. Two regions are specified: the first region within the range of −1.0 V to 0.7 V implies that the capacitance linearly changes with a particular slope for all frequencies, whereas the second region beyond 0.7 V implies that the capacitance exponentially increases for all frequencies, excluding 1 MHz, which has the same slope as before. The exponential nature of the capacitance is similar to that of the previous study.90 It is also evident from the figure that the capacitance decreases at a particular bias potential with an increase in frequency, and then rises at a particular frequency with an increase in bias potential. Because of the insensitivity of absorber traps at higher frequencies, traps functionally do not respond and lower the functional charge to ensure decreased capacitance (C = Q/V) of the device. A noticeable trap in the figure is that the capacitance of the device is almost insensitive to the bias voltage at the frequency in the visible range.
Mott–Schottky (MS) plots are a renowned and reliable tool for determining the built-in potential (Vbi) of the variance between an electrode's work functions91 and a device's carrier concentration level. Fig. 9(b) describes the MS plot at different frequencies for the proposed PSC device. The flat band potential at the interface is an essential characteristic of SC design. The charge carrier transmission at the interface is facilitated by the larger flat-band potential. The SC's flat-band potential results from the joining point of the 1/C2 curve and the voltage axis in the MS plot. According to Fig. 9(b), the slope of each curve for different frequencies is negative, and all are linear in nature. This implies that holes are the majority carriers, and the space-charge region occupies the majority of the p-type RbGeI3 layer.1 Additionally, the linear section of the MS curves across all observed frequencies exhibits an essentially constant slope. Conversely, for these observed frequencies, the intersection of each linear region at 1/C2 = 0 is different. The localized deep states in the absorber material might be the cause of this deviation, according to certain theories. The deep state contribution is negligible at higher frequencies, and the majority of the capacitance contribution is derived from the inflection of majority carriers near to the depletion region's boundary.92
3.8.2 Consequences of the thickness and concentration of the carrier of the absorber layer on the C–V attribution.
To investigate the consequences of thickness and concentration of the carrier of the absorber layer on the C–V attribution, the thickness as well as carrier density of the absorber layer were varied within the range of 300–900 nm and 1014–1019 cm−3, respectively, while other specifications remained the same, as in Tables 1–3, at a constant frequency of 1 MHz. According to Fig. 10(a), the capacitance of the device linearly increases with nearly constant slope for all thicknesses within the variation of the bias potential. Also, the figure depicts that the change in capacitance with the variation in thickness is insignificant, and these results were also obtained in a previous study.93
|
| Fig. 10 The effect of the absorber (RbGeI3) layer's (a) thickness and (b) concentration of the carrier on the C–V curve of the device. | |
Also, according to Fig. 10(b), the capacitance of the device increases with the rise in bias potential for all carrier concentrations and acts as an MS junction. The figure also shows that at a fixed bias potential, the capacitance dramatically increases with the carrier concentration. The increase in doping density increases the charge buildup at the interface, which implies an increase in capacitance, as reported in a prior study.1
3.9 Consequences of the back metal contact work function on device performance and its optimization
The back metal contact's work function exerts a crucial influence on the PSC's stability and performance. To gather holes from the external circuit on the HTM layer, an electrode of metal is situated. Making contact with an ohmic type is a prerequisite for the suitable assemblage of the charge of the majority carrier, i.e., a hole through the contact of the back metal. In this investigation, to study the consequence of the work function of the back metal on the performance-determining parameters of PV devices, ten back metal work functions were used (5.00 eV (C, Co, Ge, Ir), 5.04 eV (Ni), 5.12 eV (Pt), 5.22 eV (Ni, Pd, Pt, W), 5.31 eV (Au), 5.35 eV (Ni), 5.37 eV (Au), 5.42 eV (Ir), 5.47 eV (Au), and 5.60 eV (Pd)),48,49,94 while other specifications remained the same, as in Tables 1–3. Fig. 11(a) illustrates the influence of alteration of the back metal work function on the performance-determining parameters of the PV devices. For the ‘without HTL’ structure, all the performance-determining parameters of the PV device, i.e., Voc, Jsc, FF, and PCE, remained almost unchanged until 5.22 eV. After that, they substantially increased to 5.42 eV, and beyond that, they again remained unchanged. The greater the metal's work function for the back contact, the lower the barrier height, and therefore, from the absorber, the transmission of holes becomes smoother. Consequently, cell efficiency has been substantially increased.52 The figure also shows that there is no consequence of the back metal work function on the device's performance for the ‘with HTL’ structure. Hence, a suitable back metal, Pt (5.22 eV), was used in this investigation.
|
| Fig. 11 Consequences of the (a) back metal work function (Φ) and (b) temperature (T) on device performance. | |
3.10 Consequences of temperature on device performance
Temperature crucially impacts the PSC's stability and performance. In this investigation, real environmental conditions were used to study the impact of temperature on the performance-determining parameters of the PV devices and the stability of the proposed PSC device. The temperature was varied within the range of 280–500 K, while other specifications remained the same, as in Tables 1–3. The influence on the performance value due to the alteration of the temperature is demonstrated in Fig. 11(b). According to the figure, for ‘with HTL’ and ‘without HTL,’ except for Jsc, all the performance parameters of the PV device, i.e., Voc, FF, and PCE, significantly decreased from the initiation of the temperature variation. More electron–hole pairs were formed with the increase in temperature, which implies increasing recombination rates between energy bands. The reverse saturation current was increased by the carrier's internal recombination rate, which lowered the Voc,95 and it can be realized by eqn (7). In addition, raising the temperature increased interface defects, which may be responsible for lowering Voc. When the temperature was increased, the FF and PCE values decreased, which was probably due to a decrease in the shunt resistance.96 However, all the PV performance parameters ‘with HTL’ have a higher value as compared to ‘without HTL.’ Fig. 11(b) recommends that to chose at room temperature (300 K) of the proposed PSC device with HTL.
3.11 Optimized features of the proposed PSC device
3.11.1 Justification of the appropriateness of Sb2S3 and TiO2 as HTLs and ETLs, respectively.
A diagram of the energy state for the perovskite absorber layer, different HTLs and ETLs, as well as back and front contacts, are shown in Fig. 12. The band alignment in the heterojunction solar cell layers, specifically, the conduction band offset (CBO) at the interface of the ETL and absorber layers as well as the valence band offset (VBO) at the interface of the absorber and HTLs, are the main determinants of the collection efficiency for the photogenerated carrier.97 Using eqn (15) and (16),97 the CBO and VBO at the ETL/absorber and absorber/HTL interfaces for different HTLs and ETLs were calculated, and are tabulated in Table 7. | CBO = χabsorber − χETL | (15) |
| VBO = (χETL − χabsorber) + (Eg(HTL) − χg(Absorber)) | (16) |
where χabsorber, χETL, and χHTL denote the electron affinity of the three layers, i.e., absorber, ETL, and HTL materials, respectively. Also, χg(Absorber) and Eg(HTL) denote the bandgap of absorber and HTL materials, respectively.
|
| Fig. 12 Energy band for perovskite absorber, and different HTMs and ETMs, as well as back and front contacts. | |
Table 7 VBO and CBO of different HTMs and ETMs
Different layers |
VBO (eV) |
CBO (eV) |
BaSi2 (HTL) |
+0.61 |
+0.60 |
AgInTe2 (HTL) |
+0.45 |
+0.30 |
Cu2Te (HTL) |
−0.08 |
−0.20 |
CdSe (HTL) |
−0.42 |
−0.03 |
SnS (HTL) |
+0.01 |
+0.30 |
CuO (HTL) |
−0.37 |
−0.17 |
Sb2Se3 (HTL) |
−0.36 |
−0.14 |
MoS2 (HTL) |
−0.29 |
+0.10 |
CuSCN (HTL) |
+0.11 |
+2.20 |
Sb
2
S
3
(HTL)
|
−0.11
|
+0.20 |
TiO
2
(ETL)
|
−1.99 |
−0.10
|
Cd0.5Zn0.5S (ETL) |
−1.59 |
−0.10 |
In2Se3 (ETL) |
−0.99 |
+0.10 |
ZnSe (ETL) |
−1.58 |
−0.19 |
IGZO (ETL) |
−1.27 |
−0.34 |
SnS2 (ETL) |
−2.00 |
−0.26 |
WS2 (ETL) |
−0.96 |
−0.40 |
CdS (ETL) |
−1.61 |
−0.50 |
In3Se4 (ETL) |
−1.14 |
−0.65 |
In2S3 (ETL) |
−2.19 |
−0.10 |
A larger negative or positive value of VBO promotes hole accumulation at the HTL/absorber interface, which results in a higher rate of recombination at the interface, and hence, the SC performance will be degraded. Therefore, a lower negative or positive VBO should have been expected. Again, the transit of holes generated by photons from the layer of absorbers to the HTLs is hindered by positive VBO, whereas when negative VBO is considered, no such type of hindrance occurred.98
From the table, it is evident that there are lower negative VBOs of −0.08 eV and −0.11 eV for Cu2Te and Sb2S3, respectively, and thus, Cu2Te is more acceptable as an HTL material because if VBO is negative, there is no barrier preventing the photogenerated holes from moving toward the back electrode. Also, interface recombination increases as a result of an increase in negative VBO.98 However, according to Table 6, the PCE of Cu2Te is lower than that of Sb2S3. Therefore, Sb2S3 may be realized as an appropriate HTL material.
Also, a larger negative CBO value promotes electron accumulation at the absorber/ETL interface, resulting in a higher rate of recombination at the interface, and hence, the SC performance will be degraded. Therefore, a lower negative CBO should have been expected. Again, the transit of electrons generated by the photons from the absorber towards the ETLs is hindered by a positive CBO, whereas there is no such type of hindrance for a negative CBO.98 From Table 7, it is also evident that TiO2, Cd0.5Zn0.5S, and In2S3 have a lower negative CBO of −0.10 eV. However, according to Table 6, the PCE of TiO2 is higher than that of other ETLs. Therefore, we chose TiO2 as an appropriate ETL material.
3.11.2 QE vs. wavelength and J vs. V characteristics.
Fig. 13(a) and (b) demonstrate the QE-wavelength and J–V characteristics correspondingly for the ‘with HTL’ and ‘without HTL’ structures of the optimized PSC. From Fig. 13(a), a noticeable decrease in QE occurred with the longer wavelength, whereas the insertion of a HTL remarkably enhanced the QE for shorter and longer wavelength photons. The photon absorption capacity increased from 82% to nearly 99.5% for the photon wavelength of 500 nm due to the insertion of the HTL. Also, the longer-wavelength photon absorption was enhanced by the insertion of the HTL. The insertion of the Sb2S3 HTL with an appropriate band gap and satisfactory band alignment next to the RbGeI3 absorber layer dramatically advanced the absorption of light at longer wavelengths (>900 nm), which is the key reason why the cell performance increased from 19.42% to 30.35%.
|
| Fig. 13 (a) QE-wavelength curve and (b) J–V curve of the optimized RbGeI3-based proposed PSC. | |
The J–V curve is a well-organized tool for estimating the carrier's loss due to recombination in the SC. Fig. 13(b) shows that there was a dramatic increase in the Voc and Jsc after simply inserting an appropriate HTL. The insertion of the HTL resulted in a lower value of current density by deducing the recombination of the minority carrier, and thereby enhancing Voc, Jsc, and hence PCE. The maximum PCE of the optimized (‘without HTL’) PSC was 19.42%, which obeys the limitations of Shockley–Queisser (highest PCE 30% at 1.1 eV,99 and highest PCE 33.16% at 1.34 eV100). Conversely, in this investigation, a boosted PCE of 30.35% was achieved by inserting an additional HTL.
On the basis of previous studies, the reason behind the improvement in cell performance for a double junction can be established by the TSA two-step photon upconversion process.16,17,56 The sub-band gap photons are conspicuously engrossed in an arrangement by Urbach tail-states of materials that create further electron–hole pairs in the TSA process. As a result, an HTL with a satisfactory band gap and carrier density and a higher coefficient of absorption might bring about an effective TSA upconversion process in the area of longer wavelengths,16,17,56,79 which may be the key reason for the performance enhancement from 19.42% to 30.35%.
3.11.3 Optimized value of PV performance parameters.
For the ‘without HTL’ structure, i.e., Al/ITO/TiO2/RbGeI3/Pt, the optimized values of Voc, Jsc, FF, and PCE were obtained as 0.923 V, 24.97 mA cm−2, 84.26%, and 19.42%, respectively.
For the ‘with HTL’ structure, i.e., Al/ITO/TiO2/RbGeI3/Sb2S3/Pt, the optimized values of Voc, Jsc, FF, and PCE were obtained as 1.067 V, 33.15 mA cm−2, 85.82%, and 30.35%, respectively, through numerical simulation with the SCAPS-1D simulator.
3.11.4 Comparison of the present investigation with prior work.
The output performance of this investigation was compared with prior work. The output performance of different RbGeI3-based PSC devices is listed in Table 8, which shows that the output performance of the proposed PSC device with its new structure has been comprehensively increased.
Table 8 Performance of the RbGeI3-based PSC devices for different configurations
Sl. no. and ref. |
Configuration |
Types |
V
oc (V) |
J
sc (mA cm−2) |
FF (%) |
PCE (%) |
Present investigation, Theo. = theoretical.
|
0126 |
FTO/RbGeI3/TiO2/Ag |
Theo. |
0.291 |
25.386 |
48.637 |
3.601 |
02101 |
FTO/TiO2/RbGeI3/NiO/Ag |
Theo. |
0.5311 |
28.89 |
63.68 |
10.11 |
0325 |
ITO/CuCrO2/RbGeI3/TiO2/Au |
Theo. |
0.89 |
33.70 |
79.20 |
23.80 |
04102 |
FTO/PCBM/RbGeI3/PTAA/C |
Theo. |
0.8325 |
33.03193 |
79.8447 |
21.9567 |
05a |
Al/ITO/TiO2/RbGeI3/Pt |
Theo. |
0.923 |
24.97 |
84.26 |
19.42 |
06a |
Al/ITO/TiO2/RbGeI3/Sb2S3/Pt |
Theo. |
1.067 |
33.15 |
85.82 |
30.35 |
4. Conclusion
In this investigation, various HTMs and ETMs were incorporated into the Pb-free perovskite-based (RbGeI3) absorber to discover the primary structure of the proposed SC device with higher performance. To evaluate the PV performance of the proposed PSC device for different structures, the SCAPS-1D simulator was utilized. After many trials, Al/ITO/TiO2/RbGeI3/Sb2S3/Pt proved to be an effective primary structure. After obtaining the finalized structure, different parameters such as temperature; work function of the left metal contact; series and shunt resistance; surface recombination velocity of holes and electrons; perovskite layer, ETM, and HTM thickness; perovskite layer, ETM, and HTM carrier concentration; defect density of the perovskite layer, ETM, and HTM; and HTL/absorber and absorber/ETL interface defect density were varied so that the proposed device would be more practical. For the ‘without HTL’ and ‘with HTL’ configurations, all the influencing factors were optimized. The optimized thickness of the ETL, absorber, and HTL for the ‘without HTL’ and ‘with HTL’ configurations was found to be 50 nm, 600 nm, and 50 nm, respectively. Also, the optimized carrier concentrations for the ETL, absorber, and HTL for ‘without HTL’ and ‘with HTL’ were found to be 1017 cm−3, 1019 cm−3, and 1021 cm−3, respectively. Moreover, the temperature impact also ensured that the suggested device was unaffected by the typical operating temperature range of 300 K to 321 K. The optimized structure ‘without HTL’ showed optimal values for Voc, Jsc, FF, and PCE of 0.923 V, 24.97 mA cm−2, 84.26%, and 19.42%, respectively, whereas the optimized structure ‘with HTL’ showed optimal values for Voc, Jsc, FF, and PCE of 1.067 V, 33.15 mA cm−2, 85.82%, and 30.35%, respectively. Comparative performances with other structures were also studied, and the investigated new structure seems to be important in the long run for experimental designs.
Author contributions
Md. Mojahidur Rahman – conceived, designed, and performed the simulation process; analyzed and interpreted the data; prepared figures; and wrote the manuscript. Md. Hasan Ali – conceived, designed, analyzed, and interpreted the data; prepared figures; wrote, reviewed, and edited the manuscript. Md. Dulal Haque – conceived, designed, and analyzed the data; and reviewed the manuscript. Abu Zafor Md. Touhidul Islam – conceived, designed, and interpreted the data; and reviewed the manuscript.
List of abbreviations
Abbreviations Elaborations
PV | Photovoltaic |
TW | Terawatts |
CTL | Charge transport layer |
HTL | Hole transport layer |
ETL | Electron transport layer |
FF | Fill factor |
PCE | Power conversion efficiency |
QE | Quantum efficiency |
PGC | Photo generated carrier |
PGH | Photo generated hole |
PGE | Photo generated electron |
PSC | Perovskite solar cell |
CB | Conduction band |
VB | Valence band |
CTM | Charge transport material |
HTM | Hole transport material |
ETM | Electron transport material |
TSA | Tail-states-assisted |
List of symbols
Symbols Meaning
L
| Thickness |
E
g
| Band gap |
χ
| Electron affinity |
ε
r
| Dielectric permittivity (relative) |
N
C
| CB effective density of states |
N
V
| VB effective density of states |
v
T,n
| Electron thermal velocity |
v
T,p
| Hole thermal velocity |
μ
n
| Electron mobility |
μ
p
| Hole mobility |
N
D
| Shallow uniform donor density |
N
A
| Shallow uniform acceptor density |
σ
n
| Capture cross section (electrons) |
σ
p
| Capture cross section (holes) |
N
t
| Total defect density (bulk) |
n
t
| Total defect density (interface) |
R
s
| Series resistance |
R
sh
| Shunt resistance |
Φ
| Work function |
ε
0
| Dielectric permittivity (vacuum) |
ρ
p
| Hole distribution |
ρ
n
| Electron distribution |
J
p
| Hole current density |
J
n
| Electron current density |
J
sc
| Short circuit current density |
J
o
| Reverse bias saturation current density |
q
| Charge of an electron |
n
| Diode ideality factor |
k
| Boltzmann constant |
T
| Ambient temperature |
V
oc
| Open circuit voltage |
η
| Power conversion efficiency |
Data availability
The data that has been used is confidential. The data will be available upon reasonable request to the corresponding author(s).
Conflicts of interest
There are no conflicts of interest among the authors.
Acknowledgements
This research has not been funded by any research organization or university. The authors express their heartfelt appreciation to Dr Marc Burgelman and his colleagues at the EIS Department at the University of Gent in Belgium for providing us with the opportunity to conduct research using the SCAPS-1D software.
References
- M. H. Ali, M. A. Al Mamun, M. D. Haque, M. F. Rahman, M. K. Hossain and A. Z. Abu, Performance Enhancement of an MoS2-Based Heterojunction Solar Cell with an In2Te3 Back Surface Field: A Numerical Simulation Approach, ACS Omega, 2023, 8(7), 7017–7029 CrossRef CAS PubMed .
- H. Chen, Z. Wang, S. Xu, Y. Zhao, Q. Cheng and B. Zhang, Energy demand, emission reduction and health co-benefits evaluated in transitional China in a 2 °C warming world, J. Cleaner Prod., 2020, 264, 121773 CrossRef .
- Y. Yu, N. Zhang and J. D. Kim, Impact of urbanization on energy demand: an empirical study of the Yangtze River Economic Belt in China, Energy Policy, 2020, 139, 111354 CrossRef .
- C. A. Wolden, J. Kurtin, J. B. Baxter, I. Repins, S. E. Shaheen, J. T. Torvik, A. A. Rockett, V. M. Fthenakis and E. S. Aydil, Photovoltaic manufacturing: present status, future prospects, and research needs, J. Vac. Sci. Technol., A, 2011, 29(3), 030801 CrossRef .
-
D. Meissner, Solar Power Implications of our Climate Crisis, Source: https://www.quantsol.org/pub/pub11_08.pdf.
- S. Ahmmed, A. Aktar, M. H. Rahman, J. Hossain and A. B. Abu, Design and simulation of a high-performance CH3NH3Pb(I1−xClx)3-based perovskite solar cell using a CeOx electron transport layer and NiO hole transport layer, Semicond. Sci. Technol., 2021, 36(3), 035002 CrossRef CAS .
- M. Spalla, L. Perrin, E. Planes, M. Matheron, S. Berson and L. Flandin, Effect of the Hole Transporting/Active Layer Interface on the Perovskite Solar Cell Stability’, ACS Appl. Energy Mater., 2020, 3(4), 3282–3292 CrossRef CAS .
- J. Y. Kim, J.-W. Lee, H. S. Jung, H. Shin and N. G. Park, High-Efficiency Perovskite Solar Cells, Chem. Rev., 2020, 120(15), 7867–7918 CrossRef CAS PubMed .
- A. Kuddus, M. F. Rahman, J. Hossain and A. B. M. Ismail, Enhancement of the performance of CdS/CdTe heterojunction solar cell using TiO2/ZnO bi-layer ARC and V2O5 BSF layers: a simulation approach, EPJ Appl. Phys., 2020, 92(2), 20901 CrossRef .
- M. A. Rahman, Design and simulation of a high-performance Cd-free Cu2SnSe3 solar cells with SnS electron-blocking hole transport layer and TiO2 electron transport layer by SCAPS-1D, SN, Appl. Sci., 2021, 3(2), 253 CAS .
- M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami and H. J. Snaith, Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites, Science, 1979, 2012338(6107), 643–647 CrossRef PubMed .
- B. Conings, J. Drijkoningen, N. Gauquelin, A. Babayigit, J. D'Haen, L. D'Olieslaeger, A. Ethirajan, J. Verbeeck, J. Manca, E. Mosconi and F. D. Angelis, Intrinsic Thermal Instability of Methylammonium Lead Trihalide Perovskite, Adv. Energy Mater., 2015, 5(15), 1500477 CrossRef .
- I. Hanif, S. M. Faraz Raza, P. Gago-de-Santos and Q. Abbas, Fossil fuels, foreign direct investment, and economic growth have triggered CO2 emissions in emerging Asian economies: some empirical evidence, Energy, 2019, 171, 493–501 CrossRef .
- A. Kuddus, S. K. Mostaque and J. Hossain, Simulating the performance of a high-efficiency SnS-based dual-heterojunction thin film solar cell, Opt. Mater. Express, 2012, 11(11), 3812 CrossRef .
- S. K. Mostaque, B. K. Mondal and J. Hossain, Numerical simulation on the impurity photovoltaic (IPV) effect in c-Si wafer-based dual-heterojunction solar cell, Mater. Today Commun., 2022, 33, 104442 CrossRef CAS .
- A. Kuddus, A. B. M. Ismail and J. Hossain, Design of a highly efficient CdTe-based dual-heterojunction solar cell with 44% predicted efficiency, Sol. Energy, 2021, 221, 488–501 CrossRef CAS .
- B. K. Mondal, S. K. Mostaque and J. Hossain, Theoretical insights into a high-efficiency Sb2 Se3-based dual-heterojunction solar cell, Heliyon, 2022, 8(3), E09120 CrossRef CAS PubMed .
- M. Mostefaoui, H. Mazari, S. Khelifi, A. Bouraiou and R. Dabou, Simulation of High Efficiency CIGS Solar Cells with SCAPS-1D Software, Energy Procedia, 2015, 74, 736–744 CrossRef CAS .
- S. K. Mostaque, B. K. Mondal and J. Hossain, Simulation approach to reach the SQ limit in CIGS-based dual-heterojunction solar cell, Optik, 2022, 249, 168278 CrossRef CAS .
- M. S. Rana, M. M. Islam and M. Julkarnain, Enhancement in efficiency of CZTS solar cell by using CZTSe BSF layer, Sol. Energy, 2021, 226, 272–287 CrossRef CAS .
- Y. H. Khattak, F. Baig, S. Ullah, B. Marí, S. Beg and H. Ullah, Numerical modeling baseline for high efficiency (Cu2FeSnS4) CFTS based thin film kesterite solar cell, Optik, 2018, 164, 547–555 CrossRef CAS .
- F. Kouadio Konan, H. Joël Tchognia Nkuissi and B. Hartiti, Numerical Simulations of Highly Efficient Cu2FeSnS4 (CFTS) Based Solar Cells, Int. J. Renewable Energy Res., 2019, 9(4), 1865–1872 Search PubMed .
- M. D. A. H. Pappu, A. Kuddus, B. K. Mondal, A. T. Abir and J. Hossain, Design of n-CdS/p-CuInTe2/p+-MoS2 thin film solar cell with a power conversion efficiency of 34.32%, Opt. Continuum, 2023, 2(4), 942 CrossRef CAS .
- D. K. Sarkar, M. Mottakin, A. K. Mahmud Hasan, V. Selvanathan, K. Sobayel, M. N. I. Khan, A. F. M. Masum Rabbani, M. Shahinuzzaman, M. Aminuzzaman, F. H. Anuar, T. Suemasu, K. Sopian and Md Akhtaruzzaman, A comprehensive study on RbGeI3 based inorganic perovskite solar cell using green synthesized CuCrO2 as hole conductor, J. Photochem. Photobiol., A, 2023, 439, 114623 CrossRef CAS .
- G. Pindolia, S. M. Shinde and P. K. Jha, Optimization of an inorganic lead free RbGeI3 based perovskite solar cell by SCAPS-1D simulation, Sol. Energy, 2022, 236, 802–821 CrossRef CAS .
- M. T. Ekwu, E. Danladi, N. N. Tasie, I. S. Haruna, O. E. Okoro, P. M. Gyuk, O. M. Jimoh and R. C. Obasi, A QUALITATIVE THEORETICAL STUDY OF INORGANIC HTM-FREE RbGeI3 BASED PEROVSKITE SOLAR CELLS USING SCAPS-1D AS A PATHWAY TOWARDS 3.601% EFFICIENCY, East, Eur. J. Phys., 2023, 2023(1), 118–124 Search PubMed .
- A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, Organometal halide perovskites as visible-light sensitizers for photovoltaic cells, J. Am. Chem. Soc., 2009, 131(17), 6050–6051 CrossRef CAS PubMed .
- W. Ke and M. G. Kanatzidis, Prospects for low-toxicity lead-free perovskite solar cells, Nat. Commun., 2019, 10(1), 965 CrossRef PubMed .
- N. K. Noel, N. K. Noel, S. D. Stranks, A. Abate, C. Wehrenfennig, S. Guarnera, A. A. Haghighirad, A. Sadhanala, G. E. Eperon, S. K. Pathak, M. B. Johnston and A. Petrozza, Lead-free organic–inorganic tin halide perovskites for photovoltaic applications, Energy Environ. Sci., 2014, 7(9), 3061–3068 RSC .
- G. Thiele, H. W. Rotter and K. D. Schmidt, Die Kristallstrukturen und Phasentransformationen des tetramorphen RbGel3, Zeitschrift für anorganische und all gemeine Chemie, 1989, 571(1), 60–68 CrossRef CAS .
- B. M. Wong and J. G. Cordaro, Coumarin dyes for dye-sensitized solar cells: a long-range-corrected density functional study, J. Chem. Phys., 2008, 129(21), 214703 CrossRef PubMed .
- K. A. Parrey, T. Farooq, S. A. Khandy, U. Farooq and A. Gupta, First principle studies on structure, magneto-electronic and elastic properties of photovoltaic semiconductor halide (RbGeI3) and ferromagnetic half metal oxide (RbDyO3), Comput. Condens. Matter, 2019, 19, E00381 CrossRef .
- U.-G. Jong, C.-J. Yu, Y.-H. Kye, Y.-G. Choe, W. Hao and S. Li, First-Principles Study on Structural, Electronic, and Optical Properties of Inorganic Ge-Based Halide Perovskites, Inorg. Chem., 2019, 58(7), 4134–4140 CrossRef CAS PubMed .
- K. Deepthi Jayan and V. Sebastian, Ab initio DFT determination of structural, mechanical, optoelectronic, thermoelectric and thermodynamic properties of RbGeI3 inorganic perovskite for different exchange-correlation functionals, Mater. Today Commun., 2021, 28, 102650 CrossRef CAS .
- Z. Qin, Y. Zhang and J. Guo, SCAPS simulation and DFT study of ultra-thin lead-free perovskite solar cells based on RbGeI3, Opt. Commun., 2024, 554, 130187 CrossRef CAS .
-
A. Kowsar, M. Billah, S. Dey, S. C. Debnath, S. Yeakin and S. F. Uddin Farhad, Comparative Study on Solar Cell Simulators, 2019 2nd International Conference on Innovation in Engineering and Technology (ICIET), IEEE, 2019, pp. 1–6 Search PubMed .
- A. Kowsar, S. N. Sakib, M. B. Billah, S. Dey, K. N. Babi, A. N. Bahar and S. F. Farhad, A novel simulator of multijunction solar cells MSCS-1D, Int. J. Renewable Energy Res., 2020, 10(3), 1369–1375 Search PubMed .
- M. A. Rahman, Numerical modeling of ultra-thin CuSbS2 heterojunction solar cell with TiO2 electron transport and CuAlO2:Mg BSF layers, Opt. Mater. Express, 2022, 12(8), 2954 CrossRef CAS .
- A. Isha,
et al., High efficiency Cu2MnSnS4 thin film solar cells with SnS BSF and CdS ETL layers: a numerical simulation, Heliyon, 2023, 9(5), E15716 CrossRef CAS PubMed .
-
U. C. Obi, Investigation Of Lead-Free Bismuth Perovskite By Numerical, African University Of Science And Technology, Abuja, PhD. diss., 2019 Search PubMed .
-
B. M. Soucase, I. Guaita Pradas and K. R. Adhikari, Numerical Simulations on Perovskite Photovoltaic Devices, Perovskite Materials: Synthesis, Characterisation, Properties, and Applications, 2016, ch. 15, pp. 445–488 Search PubMed.
- M. Minbashi, A. Ghobadi, M. H. Ehsani, H. Rezagholipour Dizaji and N. Memarian, Simulation of high efficiency SnS-based solar cells with SCAPS, Sol. Energy, 2018, 176, 520–525 CrossRef CAS .
- M. Burgelman, J. Verschraegen, S. Degrave and P. Nollet, Modeling thin-film PV devices, Prog. Photovoltaics Res. Appl., 2004, 12(23), 143–153 CrossRef CAS .
-
M. Burgelman, J. Verschraegen, B. Minnaert and J. Marlein, Numerical simulation of thin film solar cells: practical exercises with SCAPS, Proceedings of NUMOS (Int. Workshop on Numerical Modeling of Thin Film Solar Cells, Gent (B)), Gent., U Gent & Academia Press, 2007 Search PubMed .
- Y. Kawano, J. Chantana and T. Minemoto, Impact of growth temperature on the properties of SnS film prepared by thermal evaporation and its photovoltaic performance, Curr. Appl. Phys., 2015, 15(8), 897–901 CrossRef .
- H. Movla, Optimization of the CIGS based thin film solar cells: numerical simulation and analysis, Optik, 2014, 125(1), 67–70 CrossRef CAS .
-
A. Niemegeers, S. Gillis and M. Burgelman, A user program for realistic simulation of polycrystalline heterojunction solar cells: SCAPS-1D, Proceedings of the 2nd World Conference on Photovoltaic Energy Conversion, JRC, European Commission, juli, 1998, 672–675.
-
M. Yoshitake, Work Function and Band Alignment of Electrode Materials, National Institute for Materials Science, Japan, 2021 Search PubMed .
-
J. Hsizl and F. K. Schulte, Work Function of Metals, Springer Tracts in Modern Physics, 2006, vol. 85, pp. 1–150 Search PubMed.
- K. Ali and Z. Ali, Analytical study of electrical performance of SiGe-based n+–p–p+ solar cells with BaSi2 BSF structure, Sol. Energy, 2021, 225, 91–96 CrossRef CAS .
- C. Doroody,
et al., Impact of Back Surface Field (BSF) Layers in Cadmium Telluride (CdTe) Solar Cells from Numerical Calculation, Int. J. Recent Technol. Eng., 2019, 8(4), 6218–6222 Search PubMed .
- Md. N. H. Riyad, A. Sunny, M. M. Khatun, S. Rahman and S. R. Al Ahmed, Performance evaluation of WS2 as buffer and Sb2S3 as hole transport layer in CZTS solar cell by numerical simulation, Eng. Rep., 2023, 5(5), E12600 CrossRef CAS .
- M. K. Hossain, M. H. K. Rubel, G. F. I. Toki, I. Alam, M. F. Rahman and H. Bencherif, Effect of Various Electron and Hole Transport Layers on the Performance of CsPbI3-Based Perovskite Solar Cells: A Numerical Investigation in DFT, SCAPS-1D, and wxAMPS Frameworks, ACS Omega, 2022, 7(47), 43210–43230 CrossRef CAS PubMed .
- S. Yasin, Z. A. Waar, T. Al Zoubi and M. Moustafa, Optoelectronic simulation of a high efficiency C2N based solar cell via buffer layer optimization, Opt. Mater., 2021, 119, 111364 CrossRef CAS .
- S. Tripathi, Sadanand, P. Lohia and D. K. Dwivedi, Contribution to sustainable and environmental friendly non-toxic CZTS solar cell with an innovative hybrid buffer layer, Sol. Energy, 2020, 204, 748–760 CrossRef CAS .
- B. K. Mondal, S. K. Mostaque, M. A. Rashid, A. Kuddus, H. Shirai and J. Hossain, Effect of CdS and In3Se4 BSF layers on the photovoltaic performance of PEDOT:PSS/n-Si solar cells: simulation based on experimental data, Superlattices Microstruct., 2021, 152, 106853 CrossRef CAS .
- F. Belarbi, W. Rahal, D. Rached, S. Benghabrit and M. Adnane, A comparative study of different buffer layers for CZTS solar cell using Scaps-1D simulation program, Optik, 2020, 216, 164743 CrossRef CAS .
- K. Kanchan, A. Sahu and B. Kumar, Numerical Simulation of Copper Indium Gallium Diselenide Solar Cell with Ultra-Thin BaSi2 Back Surface Field Layer Using the Non-Toxic In2Se3 Buffer Layer, Silicon, 2022, 14(18), 12675–12682 CrossRef CAS .
- H. A. Mohamed, Dependence of efficiency of thin-film CdS/CdTe solar cell on optical and recombination losses, J. Appl. Phys., 2013, 113(9), 093105 CrossRef .
- N. R. Paudel, K. A. Wieland and A. D. Compaan, Ultrathin CdS/CdTe solar cells by sputtering, Sol. Energy Mater. Sol. Cells, 2012, 105, 109–112 CrossRef CAS .
- P. Singh and N. M. Ravindra, Temperature dependence of solar cell performance—an analysis, Sol. Energy Mater. Sol. Cells, 2012, 101, 36–45 CrossRef CAS .
- Z. Hawash, L. K. Ono, S. R. Raga, M. V. Lee and Y. Qi, Air-Exposure Induced Dopant Redistribution and Energy Level Shifts in Spin-Coated Spiro-MeOTAD Films, Chem. Mater., 2015, 27(2), 562–569 CrossRef CAS .
- A. F. Akbulatov,
et al., Effect of Electron-Transport Material on Light-Induced Degradation of Inverted Planar Junction Perovskite Solar Cells, Adv. Energy Mater., 2017, 7(19), 1700476 CrossRef .
- A. K. Jena, Y. Numata, M. Ikegami and T. Miyasaka, Role of spiro-OMeTAD in performance deterioration of perovskite solar cells at high temperature and reuse of the perovskite films to avoid Pb-waste, J. Mater. Chem. A, 2018, 6(5), 2219–2230 RSC .
- K. Norrman, M. V. Madsen, S. A. Gevorgyan and F. C. Krebs, Degradation Patterns in Water and Oxygen of an Inverted Polymer Solar Cell, J. Am. Chem. Soc., 2010, 132(47), 16883–16892 CrossRef CAS PubMed .
- J. Liang,
et al., All-Inorganic Perovskite Solar Cells, J. Am. Chem. Soc., 2016, 138(49), 15829–15832 CrossRef CAS PubMed .
- N. A. N. Ouedraogo, Stability of all-inorganic perovskite solar cells, Nano Energy, 2020, 67, 104249 CrossRef CAS .
- A. K. Singh, S. Srivastava, A. Mahapatra, J. K. Baral and B. Pradhan, Performance optimization of lead free-MASnI3 based solar cell with 27% efficiency by numerical simulation, Opt. Mater., 2021, 117, 111193 CrossRef CAS .
- H. Wei,
et al., Challenges and strategies of all-inorganic lead-free halide perovskite solar cells, Ceram. Int., 2022, 48(5), 5876–5891 CrossRef CAS .
- A. Tara, V. Bharti, S. Sharma and R. Gupta, Device simulation of FASnI3 based perovskite solar cell with Zn(O0.3,S0.7) as electron transport layer using SCAPS-1D, Opt. Mater., 2021, 119, 111362 CrossRef CAS .
- A. Ghosh, S. S. Dipta, S. S. S. Nikor, N. Saqib and A. Saha, Performance analysis of an efficient and stable perovskite solar cell and a comparative study of incorporating metal oxide transport layers, J. Opt. Soc. Am. B, 2020, 37(7), 1966 CrossRef CAS .
- D. Saikia, J. Bera, A. Betal and S. Sahu, Performance evaluation of an all inorganic CsGeI3 based perovskite solar cell by numerical simulation, Opt. Mater., 2022, 123, 111839 CrossRef CAS .
- J. Barbé,
et al., Amorphous Tin Oxide as a Low-Temperature-Processed Electron-Transport Layer for Organic and Hybrid Perovskite Solar Cells, ACS Appl. Mater. Interfaces, 2017, 9(13), 11828–11836 CrossRef PubMed .
- K. K. Subedi,
et al., Enabling bifacial thin film devices by developing a back surface field using CuxAlOy, Nano Energy, 2021, 83, 105827 CrossRef CAS .
- L. Lin, L. Jiang, P. Li, B. Fan and Y. Qiu, A modeled perovskite solar cell structure with a Cu2O hole-transporting layer enabling over 20% efficiency by low-cost low-temperature processing, J. Phys. Chem. Solids, 2019, 124, 205–211 CrossRef CAS .
- M. K. Hossain, S. Bhattarai, A. A. Arnab, M. K. Mohammed, R. Pandey, M. H. Ali, M. F. Rahman, M. R. Islam, D. P. Samajdar, J. Madan and H. Bencherif, Harnessing the potential of CsPbBr3-based perovskite solar cells using efficient charge transport materials and global optimization, RSC Adv., 2023, 13(30), 21044–21062 RSC .
- A. Chatterjee, S. Biswas and A. Sinha, Tunneling Current of an AlGaAs/GaAs Multiple-Quantum-well Solar Cell Considering a Trapezoidal Potential Barrier, Int. J. Renewable Energy Res., 2018, 8(2), 672–681 Search PubMed .
- J. Hossain, M. M. A. Moon, B. K. Mondal and M. A. Halim, Design guidelines for a highly efficient high-purity germanium (HPGe)-based double-heterojunction solar cell, Opt Laser Technol., 2021, 143, 107306 CrossRef CAS .
- J. Hossain, B. K. Mondal and S. K. Mostaque, Computational investigation on the photovoltaic performance of an efficient GeSe-based dual-heterojunction thin film solar cell, Semicond. Sci. Technol., 2022, 37(1), 015008 CrossRef CAS .
- S. Ahmmed, A. Aktar, J. Hossain and A. B. M. Ismail, Enhancing the open circuit voltage of the SnS based heterojunction solar cell using NiO HTL, Sol. Energy, 2020, 207, 693–702 CrossRef CAS .
- M. M. Khatun, A. Sunny and S. R. Al, Ahmed, Numerical investigation on performance improvement of WS2 thin-film solar cell with copper iodide as hole transport layer, Sol. Energy, 2021, 224, 956–965 CrossRef CAS .
- S. Kohnehpoushi, P. Nazari, B. A. Nejand and M. Eskandari, MoS2: a two-dimensional hole-transporting material for high-efficiency, low-cost perovskite solar cells, Nanotechnology, 2018, 29(20), 205201 CrossRef PubMed .
- S. Taheri, A. Ahmadkhan kordbacheh, M. Minbashi and A. Hajjiah, Effect of defects on high efficient perovskite solar cells, Opt. Mater., 2021, 111, 110601 CrossRef CAS .
- M. S. S. Basyoni,
et al., On the Investigation of Interface Defects of Solar Cells: Lead-Based vs Lead-Free Perovskite, IEEE Access, 2021, 9, 130221–130232 Search PubMed .
- N. Jensen, R. M. Hausner, R. B. Bergmann, J. H. Werner and U. Rau, Optimization and characterization of amorphous/crystalline silicon heterojunction solar cells, Prog. Photovoltaics Res. Appl., 2002, 10(1), 1–13 CrossRef CAS .
- M. K. Hossain,
et al., Numerical Analysis in DFT and SCAPS-1D on the Influence of Different Charge Transport Layers of CsPbBr3 Perovskite Solar Cells, Energy Fuels, 2023, 37(8), 6078–6098 CrossRef CAS .
- Y. Li,
et al., Ultra-high open-circuit voltage of perovskite solar cells induced by nucleation thermodynamics on rough substrates, Sci. Rep., 2017, 7(1), 46141 CrossRef CAS PubMed .
- Y. Galagan, E. W. C. Coenen, W. J. H. Verhees and R. Andriessen, Towards the scaling up of perovskite solar cells and modules, J. Mater. Chem. A, 2016, 4(15), 5700–5705 RSC .
-
S. M. Sze, Y. Li and K. K. Ng, Physics of semiconductor devices, John Wiley & Sons, 2021 Search PubMed .
- I. Mora-Seró, G. Garcia-Belmonte, P. P. Boix, M. A. Vázquez and J. Bisquert, Impedance spectroscopy characterisation of highly efficient silicon
solar cells under different light illumination intensities, Energy Environ. Sci., 2009, 2(6), 678–686 RSC .
- G. G. Malliaras, J. R. Salem, P. J. Brock and C. Scott, Electrical characteristics and efficiency of single-layer organic light-emitting diodes, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 58(20), R13411 CrossRef CAS .
- G. K. Gupta, A. Garg and A. Dixit, Electrical and impedance spectroscopy analysis of sol-gel derived spin coated Cu2ZnSnS4 solar cell, J. Appl. Phys., 2018, 123(1), 013101 CrossRef .
- R. K. Zahoo, Effect of carrier concentration and thickness of absorber layer on performance CBTS solar cell, Turkish J. Comput. Math. Educ., 2021, 12(10), 5056–5064 CrossRef .
- H. B. Michaelson, The work function of the elements and its periodicity, J. Appl. Phys., 1977, 48(11), 4729–4733 CrossRef CAS .
- S. R. Al Ahmed, A. Sunny and S. Rahman, Performance enhancement of Sb2Se3 solar cell using a back surface field layer: a numerical simulation approach, Sol. Energy Mater. Sol. Cells, 2021, 221, 110919 CrossRef .
- N. Singh, A. Agarwal and M. Agarwal, Numerical simulation of highly efficient lead-free perovskite layers for the application of all-perovskite multi-junction solar cell, Superlattices Microstruct., 2021, 149, 106750 CrossRef CAS .
- S. Ahmmed,
et al., Performance analysis of lead-free CsBi3I10-based perovskite solar cell through the numerical calculation, Sol. Energy, 2021, 226, 54–63 CrossRef CAS .
- S. Ahmed, F. Jannat, Md. A. K. Khan and M. A. Alim, Numerical development of eco-friendly Cs2TiBr6 based perovskite solar cell with all-inorganic charge transport materials via SCAPS-1D, Optik, 2021, 225, 165765 CrossRef CAS .
- W. Shockley and H. J. Queisser, Detailed Balance Limit of Efficiency of p–n Junction Solar Cells, J. Appl. Phys., 1961, 32(3), 510–519 CrossRef CAS .
- S. Rühle, Tabulated values of the Shockley–Queisser limit for single junction solar cells, Sol. Energy, 2016, 130, 139–147 CrossRef .
- D. K. Sarkar,
et al., A comprehensive study on RbGeI3 based inorganic perovskite solar cell using green synthesized CuCrO2 as hole conductor, J. Photochem. Photobiol., A, 2023, 439, 114623 CrossRef CAS .
- G. Pindolia, S. M. Shinde and P. K. Jha, Void of lead and non-carcinogenic germanium based RbGeI3 PSC using organic charge transport layers: towards a clean and green future, J. Mater. Sci.: Mater. Electron., 2023, 34(9), 804 CrossRef CAS .
|
This journal is © The Royal Society of Chemistry 2024 |