Next Article in Journal
Features of the Process Obtaining of Mg-Zn-Y Master Alloy by the Metallothermic Recovery Method of Yttrium Fluoride Melt
Previous Article in Journal
Enhanced Li+ Ionic Conduction and Relaxation Properties of Li5+2xLa3Ta2-xGaxO12 Garnets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simulation of a New CZTS Solar Cell Model with ZnO/CdS Core-Shell Nanowires for High Efficiency

School of Optoelectronic Engineering, Changchun University of Science and Technology, 7089 Weixing Rd, Changchun 130022, China
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(6), 772; https://doi.org/10.3390/cryst12060772
Submission received: 4 April 2022 / Revised: 12 May 2022 / Accepted: 20 May 2022 / Published: 26 May 2022

Abstract

:
The numerical modeling of Cu2ZnSnS4 solar cells with ZnO/CdS core-shell nanowires of optimal dimensions with and without graphene is described in detail in this study. The COMSOL Simulation was used to determine the optimal values of core diameter and shell thickness by comparing their optical performance and to evaluate the optical and electrical properties of the different models. The deposition of a nanolayer of graphene on the layer of MoS2 made it possible to obtain a maximum absorption of 97.8% against 96.5% without the deposition of graphene.The difference between generation rates and between recombination rates of electron–hole pairs of models with and without graphene is explored.The electrical parameters obtained, such as the filling factor (FF), the short-circuit current density (Jsc), the open-circuit voltage (Voc), and the efficiency (EFF) are, respectively, 81.7%, 6.2 mA/cm2, 0.63 V, and 16.6% in the presence of graphene against 79.2%, 6.1 mA/cm2, 0.6 V, and 15.07% in the absence of graphene. The suggested results will be useful for future research work in the field of CZTS-based solar cells with ZnO/CdS core-shell nanowires with broadband light absorption rates.

1. Introduction

Recent studies on thin-film solar cells have shown that Cu–Zn–Sn–S, i.e., Cu2ZnSnS4 (CZTS), is a very good absorber [1,2,3]. In addition to its high absorption coefficient of around 104 cm−1 and its refractive index of 2.72, CZTS has an interesting bandgap of around 1.45 to 1.65 eV [4,5,6,7,8]. On the other hand, the application of nanotechnology to the solar cell improves its performance (high absorption, low reflection, etc.) and reduces the cost of nanowire materials [9,10,11].
In 2013, Research on the fabrication of MEH-PPV [2-methoxy-5-(2′-ethylhexyloxy)-p-phenylenevinylene]/Al:ZnO nanorod arrays based on ordered bulk heterojunction hybrid solar cells) has resulted in high performance (Jsc = 5.32 mA/cm2, Voc = 195 mV, FF = 27.71%,and EFF = 0.287%), an average optical transmission of approximately 78%, and an improved absorption coefficient compared to the catalyst layer seeded [12]. On the other hand, in 2014, an Experimental study of aligned ZnO nanorods on ZnO thin films doped with 2% Sn yielded high transmittance in the visible region, a large area with longer aligned ZnO nanorods, and improved performance (short-circuit current density Jsc = 2.840 mA/cm2, open-circuit voltage Voc = 0.552, fill factor FF = 0.381, and the power conversion efficiency of 0.599%) compared to the undoped film [13]. Also, in 2015, Jalal Rouhi et al. showed that vertically aligned ZnO–ZnS core-shell nanocone arrays can be useful for optoelectronic devices due to a good surface-to-volume ratio, good electrical performance (Jsc = 12.2 mA/cm2, Voc = 0.68 V, and FF = 0.49), better light trapping, and longer life than the other structures [14].
Recent photocatalytic studies indicate that ZnO/CdS core-shell nanowire arrays exhibit high cell separation efficiency compared to ZnO nanorods [15]. Experimental studies have also proven that the photocatalytic and photoelectrochemical activities of the ZnO/CdS core-shell structure are significantly higher than those of the bare ZnO structure [15,16]. This means that the choice of the diameter of the ZnO core and the thickness of the CdS shell plays an important role in improving the performance of the solar cell. Although cadmium (Cd) is a toxic product, CdSremains a very good junction zone reducer [17], which improves its interface with the absorber. Note also that in 2015, by Ansoft HFSS14 simulation on a periodic array of double-shell nanowire structures (ZnO/CdS/CZTS) embedded in a thin multilayer film, Qian Liu et al. obtained an absorption rate of about 90% in a range of wavelengths between 400 nm and 1000 nm, i.e., a frequency range of 300 to 750 THz [18,19].As molybdenum (Mo) is widely used as a back contact in CZTS-based cells considering its good electrical properties [20], due to its chemical reaction with CZTS particles, a MoS2 layer is formed, and voids are observed. Although the formed layer of MoS2 has a positive effect, the growth of the MoS2 layer near the Mo/CZTS interface during the sulfurization process can negatively impact the back contact cell parameters (series resistance and fill factor) depending on thickness or quality of MoS2 [21,22]; it can also be noted that the voids appearing in the MoS2/CZTS interlayer generally lead to a degradation of the performance of the solar cell [23].The use of graphene in the Mo/CZTS interface could help us to remedy this problem thanks to these properties [24]. Graphene, although it is not a good absorber, is increasingly used in applications as a two-dimensional material to improve surface contacts due to its unique large area properties, electron mobility high, and its superior catalytic capacity [25,26,27,28,29,30]. The graphene is a zero-bandgap semimetal [31,32] and has a thermal conductivity (3000–5000 Wm−1 K−1) [33]. It has also been shown that under certain conditions, the contact of graphene with the single layer of MoS2 effectively improves photon absorption and photocatalytic efficiency [34,35,36,37]. Chonge Wang et al. showed by simulation that the introduction of a graphene nanolayer between the ZnO core and the CdS shell could lead to an improvement in the optical absorption of the cell [38]. This shows that it is interesting to study the optical and electrical performance of the model in which graphene is sandwiched between the core and the shell and also deposited on the MoS2 layer. Transmission Electron Microscope (TEM) investigations of the two structures, Mo/CZTS and Mo/Graphene/CZTS, were carried out, and the presence of a layer of graphene between Mo and CZTS led to a reduction in thickness from 25% to 33% of the layer of MoS2 formed; the suppression of the growth of secondary phases near the MoS2/CZTS interface has also been observed [24]. Note that very recent research on the growth of MoS2-graphene heterostructures at low temperatures has made it possible to know that by depositing a seed layer of a thin film of Mo on graphene followed by an H2S sulphurization process, it was possible to obtain an atomic ratio of Mo on S, lower than the stoichiometric value 2 of the MoS2 standard and, therefore, an improved catalytic performance [39]. In addition, Xiaolong Liu et al. were able to observe by chemical vapor deposition that MoS2 develops preferentially with a network aligned with the epitaxial graphene [40]. It is therefore found that graphene, used as a Mo/CZTS interlayer, can serve as an agent for controlling the growth of MoS2. Note that the performance of these cells will be improved with optimally sized nanowires. Indeed, previous scientific studies have established guidelines for the choice of the diameter of the nanowire and the pitch of the network of nanowires [41,42,43,44,45,46,47,48,49]. It has been shown that for an optimized array with better optical performance and ultimate efficiency, increasing the length of the nanowires leads to a large increase in the pitch of the nanowires, but that the absorption of solar energy in the array InP Nanowire, for example, exhibits local maxima at nanowire diameters of approximately D1 = 170 and D2 = 410 nm, regardless of the length of the nanowires, knowing that the ultimate efficiency of D1 is greater than that of D2 [50]. Further, Goutam Kumar Dalapati et al. showed that with a thickness of CZTS of 1000 nm, obtained by the co-sputtering or sequential sputtering method during 15 to 30 min of the process of growing the precursor films, the thickness formed of MoS2 varies on average by 170 nm [51]. Also, Farjana Akter Jhuma et al. have shown that the appropriate thickness of CdS for CZTS is between 50 and 150 nm [52]. However, research has shown that the optical absorption of suspended monolayer graphene at a normal incidence is very low, around 2.3% in the visible and near-infrared regions [53].
Thus, to improve the light absorption of graphene in the visible and near-infrared regions, there have been several approaches. The one that caught our attention is that of the bound state in the continuum (BIC) [54]. This mechanism appears to be efficient and easier to implement. The method is based on confinement and therefore trapping of light. It was also shown that when the graphene monolayer is sandwiched between two dielectric layers, just like absorption, an increase in the electric field intensity on the upper and lower faces of the graphene layer is observed. Add to this that with the tuning of the geometric parameters, this device can serve as a modulator and chemical/biochemical sensor [55].
The particularity of this model is that a graphene layer is not only sandwiched between the CZTS and MoS2 layers but also placed between the ZnO core and its CdS shell. Additionally, these ZnO/CdScore-shell nanowire arrays are embedded in the CZTS absorber and MoS2 layer; the assembly is located between a Mo layer below and the CdS/ZnO/ITO structure above.
This shows the importance of using aZnO/CdS core-shell nanowire in CZTS-based cells with the incorporation, or not, of graphene. Thus, several studies have been carried out over the last ten years on these cells and have obtained encouraging results. Indeed, the experiments performed by R. Suet al. on ZnO/CdS core-shell nanowires gave Jsc = 3.35 mA/cm2 and FF = 32%, and on ZnO/CdS/CuSbS2, core-shell nanowires gave results such as Jsc = 6.48 mA and EFF = 52% [56]. Additionally, YoungjoTak et al., in another experiment on ZnO-CdSZnO/CdS core/shell nanowire heterostructure arrays, obtained Jsc = 7.23 mA/cm2 and Voc = 1.55 V and an efficiency of conversion of 3.53% [57]. Other studies also performed on the solar cell model based on ZnO/CdS core-shell nanowire arrays resulted in a maximum absorption of 92% at a wavelength of 500 nm [58]. Chonge Wang et al., by simulation on ZnO/CdS core-shell nanowire arrays, by including graphene in the CZTS-based solar cell structure, obtained Jsc = 6.39 mA/cm2, Voc = 0.63 V, EFF = 16.8%, and a maximum absorption and reflection of 97% and 40%, respectively [38].
Although such a model is difficult to achieve due to the growth deposition of nanowires through the graphene layer, models based on the same principle have been studied and implemented, and have shown the improved optical and electrical performance of the solar cell [55,59]. Experimentally, ZnO nanowires are grown by electrochemical process and a CdS shell is synthesized on ZnO nanowires using the hydrothermal method [15,52,55,59]. This same hydrothermal growth approach can be used to push the nanowires through the graphene layer [15].
In order to study optical and electrical performance, the ZnO core and CdS shell nanowires, after having obtained by simulation the optimum core radius and shell thickness, are embedded in a CZTS absorber and inserted into the model.
These solar cell designs appear to have a promising future, especially if used with optimally sized core-shell nanowires and an improved light trapping system. These new models of the solar cell, little known to researchers, are the subject of our study.

2. Materials and Methods

To obtain the optical and electrical performance of the different models, we used the COMSOL Multiphysics simulation software for the different models of nanowire solar cells. COMSOL uses the finite element method to solve for the electromagnetic fields within the modeling domains. It is also able to calculate a couple of several physical domains. This study focuses on two main components (semiconductor and electromagnetic wave and frequency domain) coupled with studies in the wavelength domain, semiconductor equilibrium, and stationary [60,61,62].
To choose the optimal diameter of the ZnO core and the optimal thickness of the CdS shell of the nanowire, we relied on the model of the CZTS solar cell containing ZnO/CdS core-shell nanowires immersed in a CZTS absorber. Figure 1a shows CZTS solar cell structures with core model ZnO/CdS core-shell nanowire arrays. The model is based on a 3 µm × 3 µm square base. The principle is to compare the optical performance of models with different values of ZnO core diameter (D) and CdS shell thickness (Sh).
It is known that for minimum reflection, antireflection coatings with refractive indices n1 and n2, of thicknesses d1, d2, of wavelength λ , satisfy the following condition: n1d1 = n2d2 = λ /4 [63,64]. The dimensions of the models were obtained by taking these criteria into account and choosing λ = 888 nm [21].
Each nanowire has a length of 1000 nm and a ZnO core covered with a CdS shell. The space between the nanowires is filled with MoS2/CZTS films of respective thicknesses of 170 nm and 830 nm. The top, along the z-axis, is formed of layers of CdS/ZnO/In2O3-SnO2 (Indium tin oxide or ITO) structures with respective thicknesses of 100 nm, 140 nm, and 170 nm. Below the nanowires, along the z-axis, is the layer of Mo with a thickness of 200 nm. An air thickness of 500 nm has been added above the model only to take into account the air parameters as shown in Figure 1b. The device as presented is composed from bottom to top of a rear contact layer in Mo, an absorbent layer in CZTS, a buffer layer in CdS, a window layer in ZnO, and a layer of transparent oxide in ITO. The substrate and the grid contact are not represented in the model. This structure was chosen to take into account the properties of the materials listed above.
By respecting the conditions and geometric constraints cited below in the model, we chose nanowires 160 nm in diameter, including the diameter (D) of the core and the thickness (Sh) of the shell, such that D + 2 × Sh = 160 nm. The core diameter and the shell thickness of the three models to be compared are, therefore: 70 nm and 45 nm, 100 nm and 30 nm, and 120 nm and 20 nm. For a more efficient nanowire array, the fill factor (nanowire diameter/nanowire pitch) should be between 0.4 and 0.5 [65]. This factor, like many other studies, allowed us to choose a nanowire pitch of 360 nm [15,43,66].
The study of this simulation proceeds in two phases. The first is a comparative study based on the calculation of the electric field, absorption, reflection, and transmission in a single nanowire for each of the three models. The results of this study will make it possible to choose an efficient model with optimal dimensions. Figure 1b shows the CZTS solar cell structures with ZnO/CdS core-shell nanowire arrays of a single nanowire model. This model is based on a 360 nm × 360 nm square base.
We considered the Solar Irradiance Spectra Air Mass (AM1.5) [67] with a power density of 100 mW/cm2 for solar lighting, whose solar zenith angle is 48.2°.
The absorption, A ( λ ) , as a function of the wavelength can be calculated by Equation (1) [68,69], but in our simulation, the absorption, A ( λ ) , was calculated using Equation (2):
A λ = 1 2 2 π ε 0 c λ E λ 2 I m ( n 2 λ ) d V
where E( λ ) is a vector of the electric field as a function of the wavelength, ε0 = 8.85 × 10−12 Fm−1 is the vacuum permittivity, c = 299,792,458 m/s is the speed of light in vacuum, λ is the photon wavelength, and n is the complex refractive index. It is recalled that the absorption, A( λ ), the transmitted incident light, T( λ ), and the reflection, R( λ ), in the cell must satisfy the relation below [65]:
R λ + A λ + T λ = 1
In the magnetic propagation wave, the diffusion parameters are determined by the parameter S in terms of the electric field. The electromagnetic and electric fields acting on the ports are, therefore, calculated after the excitation of the port1. The electric fields E1 and E2, respectively at ports 1 and 2, are transformed into a complex scalar number corresponding to the voltage. We assumed that the fields are normalized with respect to the integral of the energy flow through each cross-section of the ports. The electric field EC calculated at port1 also contains the reflected field. The coefficients S11 and S21 are calculated automatically according to Equations (3) and (4) [62]. These coefficients respectively determine the reflection, R( λ ), and the transmission, T( λ ). The absorption is calculated by the Equation (2). The results obtained by Equation (1) can be used for verification.
S 11 = P o r t 1 E c E 1 * E 1 * d A 1 P o r t 1 ( E 1 * E 1 * ) d A 1
S 21 = P o r t 2 E c * E 2 * d A 2 P o r t 2 ( E 2 * E 2 * ) d A 2
where A1 and A2 are the areas of port 1 and port 2, respectively.
The refractive indices of materials are available from Ref [70]. The refractive index of CZTS was taken from the experience of Nabeel A. Bakr et al. thanks to the WebPlotDigitizer simulator [6,71].
For more precision in the performance of the models, we calculated the ultimate absorption efficiency, η, of each model using the following relation [72,73]:
η = λ l λ g I ( λ ) A λ λ λ g d λ λ l λ u I λ d λ
where I λ is the solar spectral irradiance, and λ l = 300 nm and λ u = 4000 nm are, respectively, the negligible solar irradiance and the upper limit of the available data for the solar spectrum, and λ g = 900 nm is the photon wavelength corresponding to the average bandgap of the materials of the model. The short circuit current and open-circuit voltage give a large place to evaluate the electrical performance of the solar cell.
By using Shockley’s diode equation as a function to voltage, the current density, J, is defined by [62]
J ( V ) = J 0 [ e x p ( e K T V ) 1 ] J s c
where J 0 is saturation current density, V is the applied bias voltage, KT/e is the thermal voltage equal to 25.69 mV at 25 °C, and the short current density condition under illumination is J s c .
Table 1 contains the values of parameters used in our simulation COMSOL. The initial temperature of the cell is assumed to be very close to the ambient temperature (T = 300 K).
To emphasize the credibility of our choice, in the study of CZTS solar cells with a ZnO core diameter of 100 nm, the CdS shell thickness of 30 nm received special attention [15,21].
Taking into account the properties cited above of graphene, we propose in the second phase of the simulation to study a second model, this time containing graphene to arrive at an even more efficient model. We will call it, “the model with graphene interlayer”, designating the first model as “the model without graphene interlayer”. The simulation results of these two models will be compared. We chose the thickness of 1 nm of graphene in the simulation by referring to the experiment in which Maria Jabeen et al. showed that by choosing 1 nm thick graphene deposited on the SiO2 layer [90], the optical properties of the solar cell improved [93,94].
Figure 2 shows the basic 3D model of CZTS solar cell structures with ZnO/CdS core-shell nanowire arrays and a graphene interlayer.
Almost perfect absorption of single-layer graphene can be achieved due to the high light confinement and near-field enhancement of CZTS slabs. This absorption is calculated using Equation (1) and is very dependent on the electromagnetic frequency of light and medium. In practice, the CZTS thin film is produced by the co-sputtering technique of Cu, SnS, and ZnS for the deposition of precursors, and followed by sulfurization in an H2S atmosphere since sputter deposition is used to deposit thin-films, elemental material films, complex alloy films, etc. [95]. However, the deposited film may have a different composition from the source. Thus, after having covered the soda-lime glass with a layer of molybdenum, CZTS is deposited there. On the other hand, the chemical bath deposition being at a low temperature is simpler and more reliable for the deposition of thin layers; its use is suitable for nanomaterials, but not for semiconductors, due to the small size of the deposited crystals [96]. Thus, the p–n junction was formed by the chemical bath deposition of CdS. ZnO nanowires-based CZTS devices are fabricated by a reverse process that is used for standard thin-film solar cells, where ZnO nanowire windows are first prepared on ITO glass by the electrochemical method followed by CdS deposition, then CZTS [21].It is known that chemical vapor deposition (CVD) is useful in the process of depositing extremely thin layers of material, with high performance and high purity; therefore, the graphene monolayer is deposited directly on the Mo layer using the CVD [97,98,99].This deposition technique is chosen in this context, taking into account the fact that despite evaporation at a very high temperature of the precursors, the duration of the deposition remains very short, due to a very low thickness of the graphene.
In the visible and near-infrared regions, graphene is modeled as an ultra-thin dielectric film of a thickness of t g =   1 nm , and its new permittivity is written as ε g = 1 + j σ g ω ε 0 t g , and according to Kubo’s relation, the graphene surface conductivity, σ g , is calculated as the sum of intra band σ int ra and inter band σ int er conductivity of graphene, and given by [55]:
σ int ra ( ω ) = j e 2 k B T π 2 ( ω 2 j Γ ) [ μ c k B T + 2 ln ( e μ c K B T + 1 ) ]
σ int er ( ω ) = j e 2 4 π 2 ln [ 2 μ c ( ω j 2 Γ ) 2 μ c + ( ω j 2 Γ ) ]
where ω is the angular frequency of the photon, ε 0 is the vacuum permittivity, = 6.582 × 10−16 eVs is the reduced Plank’s constant, e = 1.6 × 10−19C is the electron charge, k B is the Boltzmann constant, µc = 0.15 eV is the chemical potential, τ = 0.5 ps is the momentum relaxation time, T = 300 K is the operating temperature, and Γ = 1 / 2 τ is the phenomenological scattering rate. KBT = 0.0256 eV, Γ = 0.11 eV, and ω = 2πf, where f is frequency. To achieve a very good approximation, the frequency must be in the range of 1 to 300 THz [99].
Thus, in the model with the graphene interlayer, the core-shell nanowires of ZnO/CdS are above the Mo/MoS2/Graphene structure. To minimize the complexity of the work, the continuation of the simulation was carried out in 2D since the difference between the results is not very large compared to 3D. In addition to the electric field and optical properties, the study was interested in determining the total rates of generation and recombination, the short-circuit current density (Jsc), the open-circuit voltage (Voc), and efficiencies. Figure 3 shows in 2D the two structural models with the graphene interlayer (Figure 3a), without graphene interlayer (Figure 3b) of the CZTS solar cell with arrays of ZnO/CdS core-shell nanowires. Both models are based on a 3µm × 3µm square base.
Also, in this model, the top, along the Z-axis, is formed by the layers of CdS/ZnO/In2O3-SnO2 (Indium tin oxide, ITO) structures with respective thicknesses of 100 nm, 140 nm, and 170 nm. Under the nanowires, along the Z-axis, is a layer of Mo with a thickness of 200 nm. Between the nanowires is the MoS2/Graphene/CZTS structure with respective thicknesses of 170 nm, 1 nm, and 829 nm. An air thickness of 500 nm has been added above each model only to take into account the air parameters.
To evaluate the behavior of charge carriers and, therefore, the density of the short-circuit current (Jsc) in the model as a function of the height, we determined through simulation the generation g(x) and recombination r(x) rates, as well as the total generation rates, G, and the total recombination rates, R, of the hole–electron pairs according to the relationships below [100,101,102]:
g ( x ) = λ i λ f [ 1 R 2 ( λ ) ] I ( λ ) α ( λ ) e α x d λ
r ( x ) = n ( x ) p ( x ) n i 2 τ p [ n ( x ) + n i ] + τ n [ p ( x ) + n i ]
G = 0 d g ( x ) d x
R = 0 d r ( x ) d x
where 𝑛 = 𝑛0 + Δ𝑛, 𝑝 = 𝑝0 + Δ𝑝, 𝑛0, and 𝑝0 are the densities of carriers at equilibrium, Δ𝑛 and Δ𝑝 are the densities of excess carriers, 𝜏n and 𝜏p are the lifetimes of the carriers at pendant bond state, and 𝑛i is the intrinsic density of carriers; d is the thickness of the model, and λ i = 200 nm and λ f = 2000 nm (wavelength study interval).
The current density (J) as a function of voltage (V) is obtained from Equation (6). The ultimate absorption efficiency is calculated using Equation (5).
Since the model is composed of several materials, we were interested in the effective permittivity, but as the system is a bit complex, in this simulation, we adopted a simplified method by calculating the average permittivity. This average permittivity can be calculated according to the following relationship [103]:
ε m o y = i = 1 7 ε i × f i
f i = V i 1 7 V j
with
1 7 f i = 1
where ε 1 , ε 2 , ε 3 , ε 4 , ε 5 , ε 6 , and ε 7 are relative permittivity of air, ITO, ZnO, CdS, CZTS, MoS2, and, Mo, respectively; f 1 , f 2 , f 3 , f 4 , f 5 , f 6 , and f 7 are volume fractions of air, ITO, ZnO, CdS, CZTS, MoS2, and Mo, respectively. According to Table 1, the permittivity of graphene is much lower than that of MoS2 and CZTS.
The values of the average permittivities with and without the graphene interlayer found, according to Equation (13), are ε m o y ( w i t h _ g r a _ int ) = 9.08 and ε m o y ( w i t h o u t _ g r a _ int ) = 9.1 , respectively.

3. Meshing Geometric Models

For simulation calculations, it is mandatory to define the mesh of each of the geometric models. Figure 4 shows the schematic mesh structures of the single nanowire of the CZTS solar cell with ZnO/CdS core-shell nanowire arrays.
The transverse magnetic field (TM) source is determined by setting the port limit as a “periodic port”. Incident light (port wave excitation) enters through the top port while the bottom port is located under the models. Two periodic conditions are necessary for the simulation. The periodic type is that of the “Floquet period”, and its vector, k, comes from the periodic port. The simulation region is defined as follows: “swept mesh” for the air region and “freely divided tetrahedral mesh” for the other regions. For this physical interface, the maximum size of the cells of the grid are preferably less than a certain fraction of the wavelength. The incident light wavelength range of this simulation model is 200 nm to 2000 nm, so the maximum cell size is 60 nm and the minimum cell size is 4 nm. All simulation models use the same grid cell size to maintain computational rigor. In “Parameters1” of the global definitions, we have chosen to add a layer of 500 nm above the model to take into account the air parameters in the simulation.
Figure 5 shows the schematic mesh structures of the models with the graphene interlayer (Figure 5a) and without the graphene interlayer (Figure 5b) of the CZTS solar cell with ZnO/CdS core-shell nanowire arrays.
The bibliographical references of the equations used in the simulation are given in the Supplementary Materials (Table S1).

4. Results and Discussion

The simulation study on the model with a single nanowire gave us the following results:
Figure 6, Figure 7 and Figure 8 show the evolution of the electric field, the absorption, and the reflection as a function of the wavelength in the different models with a single nanowire.
The synthesis of these three results is given in Table 2.
Through this summary, Table 2, we see that the optimal values of the electric field, absorption, and reflection, respectively, of 1000 V/m, 96%, and 43% are obtained by the model of the CZTS solar cell with ZnO/CdS core-shell nanowire whose the diameter of the ZnO core of is 100 nm and the thickness of the CdS shell is 30 nm.
Figure 9 shows the electric field along with the height of the model with an optimal core diameter and shell thickness values in CZTS solar cells with ZnO/CdS core-shell nanowire arrays; Figure 9a is shown with the graphene interlayer, and Figure 9b without the graphene interlayer.
We then observe that the electric field is not uniform with the height of the model. It increases on average towards the top of the model by considering Equation (16) [104].
E = q 4 π ε d 2
where q is the charge of a carrier whose absolute value is 1.602 × 10−19 C, therefore, of an electron, ε is the permittivity of the medium, and d is the distance from the top of the model.
Note that the average permittivity for the graphene interlayer model is lower than that for the model without a graphene interlayer. According to Equation (16), a decrease in the permittivity (ε) leads to an increase in the electric field. Therefore, the electric field in the model with the graphene interlayer increases on average compared to that of the model without the graphene interlayer. Note that the maximum electric field is 2093.352 V/m at a distance of 332.352 nm from the top of the cell for the model with graphene interlayer, and 2020 V/m at a distance of 259.128 nm from the top of the cell for the model without an interlayer of graphene. The average electric field is 687 V/m for the model with a graphene interlayer and 658 V/m for the model without a graphene interlayer. The addition of a 1 nm thick layer of graphene results in a slight increase in the depth, d, from the MoS2/CZTS interlayer; therefore, there is a slight decrease in the electric field from the MoS2/CZTS interlayer to the bottom of the model. Taking into account the low thickness and the low permittivity of graphene, this reduction is not too visible.
In the model with the graphene interlayer, according to Equation (1), absorption increases on average with increasing electric field. Figure 10 shows absorption as a function of wavelength (from 200 nm to 2000 nm) for models with a graphene interlayer (green curve) and without graphene interlayer (blue curve). The angle of incidence is equal to 0°. We note that in half of the ultraviolet range (between 250 and 300 nm), the two models have almost the same absorption. Note that, thanks to the unique large area properties of graphene, a slight improvement in absorption is observed in the graphene interlayer model. The maximum absorption is 97.8% for the model with the graphene interlayer against 96.5% for the model without graphene interlayer. These values were reached at a wavelength of 500 nm. By comparison with a single nanowire, an improvement in maximum absorption of 0.005% = (96.5 − 96)/96 is observed for the model without a graphene interlayer with a nanowire pitch of 360 nm. This explains the fact that with an increasing number of nanowires, the light ray trapping factor increases, thus leading to a decrease in reflection and consequently an improvement in absorption.
The results found in the literature are close to those of our model with ZnO/CdS core-shell nanowires [18,19], but the deposition of the graphene on the MoS2 layer enabled us to obtain an absorption rate of 97.8%. This core-shell nanowire model proposed in this work can therefore play a promising role in the fabrication of broadband light absorption rate solar cells [19]. This model is very cost-effective in the visible spectrum regions [105].
Figure 11 shows the reflection as a function of the wavelength (from 200 nm to 2000 nm) for the models with a graphene interlayer (green curve) and without graphene interlayer (blue curve). The angle of incidence is equal to 0°. Note that the model without a graphene interlayer has, on average, the greatest reflection, largely due to the presence of air in the MoS2/CZTS. According to Figure 11, the maximum reflection is obtained in the model without a graphene interlayer, which is 42% against 40% for the model with a graphene interlayer. By comparing the models with all the nanowires and with a single nanowire of the solar cell without a graphene interlayer, a decrease of 0.047% = (43 − 42)/43 is observed, due to an increasing number of trapped light rays due to the neighboring nanowires, and thus causing a decrease in the reflection rate.
Figure 12 and Figure 13, respectively, show the evolution of the total generation rates and the total recombination rates of the models with and without a graphene interlayer as a function of the height of the model. Taking into account the acquired values of the generation and recombination rates, the scale of the axis of these rates (𝑦 axis) is configured as a decimal logarithmic axis. It should be noted that the total generation rate of the model with a graphene interlayer is, on average, higher than that of the model without a graphene interlayer. The reverse is observed for the recombination curves. Like all the generations of charge carriers, the maximums of the generation rate totals are at the surface of the incident light side; this process is explained by Equations (9) and (11). The difference between the curves of the totals of the generation rate and that between the curves of the totals of the recombination rate varies on average by 2% and becomes almost zero at the rear contact (Mo). This difference is mainly due to the extinction coefficient, the low bandgap of the graphene, the lifetime of the carriers, and the densities of carriers in the materials. We also notice that there is a very low rate of generation and recombination of electron–hole pairs near the Mo layer.
Figure 14 shows the current density as a function of the voltage under light illumination in the two models (with and without graphene interlayer). The model without a graphene interlayer has a short-circuit current density (Jsc) of 6.1 mA/cm2, while that with a graphene interlayer Jsc is equal to 6.2 mA/cm2. This slight advantage of the model without a graphene interlayer can be explained through Equation (17) [106].
J s c = q G ( L n + L p )
where q is the amount of charge of the carriers, G total of the generations of the charge carriers, Ln and Lp are the electron and hole scattering lengths respectively. It is known that the generation and recombination rates are respectively high and low in the model with graphene interlayer thanks to the presence of graphene in the MoS2/CZTS interface. On the other hand, a decrease in the rate of recombination leads to an increase in Ln and Lp [107].
We also note that the open-circuit voltage (Voc) of the model with graphene interlayer, which is 0.63 V, is higher than that of the model without graphene interlayer, whose value is 0.6V. The explanation for this difference can be deduced from the relation (18) [107].
V o c = K T e ln ( J s c J 0 + 1 ) ,
where kT/e is the thermal voltage equal to 25.69 mV at 25 °C, Jsc is the short-circuit current density under lighting conditions, and J0 is the saturation current density, depending on the recombination in the solar cell [106]. In addition, the saturation current, J0, in the model with the graphene interlayer is the smallest, due to the difference in recombination rates.
Table 3 shows the numerical values of open-circuit voltage (Voc), short current density (Jsc), fill factor (FF), and efficiency (EFF) for models with and without a graphene interlayer. The fill factor (FF) was determined from the data of the relationship between current density (J) and voltage (V) using the relationship below [108]:
FF = J m p × V m p J s c × V o c ,
where V m p and J m p are, respectively, the voltage and the current density corresponding to the point of the curve where their product is maximum.
Note that the maximum values of Voc, Jsc, EFF, and FF are obtained by the model with a graphene interlayer.
It can be seen that the results of our study are, for the most part, similar to those given in the literature. Thanks to the introduction of graphene and an improved cell design, our model shows the best results on average, as evidenced by the information in Table 4. This table gives a summary of the results including those of previous research. Indeed, taking into account the improvement of the contact surface by graphene on the surface of the CZTS and between the core and the shell, and thanks to the significant trapping of light in the nanowires, our model and that of C Wang et al. turned out to be slightly advantageous. We see, on the other hand, that the generation rate is maximum and almost identical to the surface for the models with and without a graphene interlayer, and almost zero at a 1510 nm depth for the model with a graphene interlayer and at 1560 nm depth for the model without a graphene interlayer.

5. Conclusions

The formation of the MoS2 layer in the Mo/CZTS interface during the sulfurization process, at a certain thickness, can have negative effects on the performance parameters of the cell. This study reports the interest in using the graphene layer to control and improve the contact surface and the growth of the thickness of the MoS2 interfacial layer formed in the interface between the rear molybdenum (Mo) contact and the CZTS layer. This work was carried out using the COMSOL Multiphysics simulation software based mainly on the study and analysis of the optical and electrical characteristics of CZTS solar cells, based on ZnO/CdS core-shell nanowires with back contact of Mo. After having obtained by simulation the optimal values of the diameter of the ZnO core and of the thickness of the CdS shell, which are respectively 100 nm and 30 nm, two structural models with core-shell nanowires in ZnO/CdS immersed in MoS2/CZTS thin-film were studied, namely the structure with an interlayer of graphene (Mo/MoS2/Graphene/CZTS/CdS/ZnO/ITO) and the structure without an interlayer of graphene (Mo/MoS2/CZTS/CdS/ZnO/ITO). For the model with a graphene interlayer, the graphene light absorption without the prior device is very low; its improvement in the visible and near-infrared regions was obtained thanks to the principle of confinement of light of the bound state in the continuum (BIC). Thus, the best performances were obtained by the model with a graphene interlayer. For this model, the fill factor (FF), the short-circuit current density (Jsc), and the open-circuit voltage (Voc) are, respectively, 81.7%, 630 mV, and 6.2 mA/cm2, with a maximum absorption rate of 97.8%; while the model without a graphene interlayer recorded values of FF, Jsc, Voc, and maximum absorption rates of 79.2%, 600 mV, 6.1 mA/cm2, and 96.5%, respectively. These performances are due to certain properties of graphene such as the high mobility of the carriers, the higher catalytic capacity, the good optical conductivity, and the capacity to improve the surface contacts. In addition, the rates of recombination and generation of electron–hole pairs are respectively low and high in the model with a graphene interlayer, compared to the model without a graphene interlayer. The lowest maximum reflection, 40%, was obtained in the model with a graphene interlayer. Thanks to the proposed graphene interlayer model, an efficiency of 16.6% was obtained with the model with a graphene interlayer against 15.07% for the model without a graphene interlayer. These simulation results must first be experimentally verified since, in practice, the graphene layer is deposited directly on the Mo layer. These results, once experimentally verified, will serve as an aid for the manufacturers in improving the performance of CZTS solar cells based on ZnO/CdS core-shell nanowire arrays. This core-shell nanowire model presented in this study can be very useful in the fabrication of broadband light absorption rate solar cells. It is also very cost-effective in visible spectrum regions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12060772/s1, Table S1: List of equations used in the simulation (References [109,110,111,112,113,114,115,116,117,118,119,120] are cited in the Supplementary Materials).

Author Contributions

Conceptualization, C.W. and F.Y.; Methodology, C.W. and F.Y.; Validation, C.W. and F.Y.; Formal analysis, L.N. and B.D.; Investigation, B.D. and L.N.; Data curation, B.D.; Writing—review and editing, B.D.; Visualization, F.Y.; Supervision, F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahbub, R.; Islam, M.S.; Anwar, F.; Satter, S.S.; Ullah, S.M. Simulation of CZTS thin-film solar cell for different buffer layers for high-efficiency performance. South Asian J. Eng. Tech. 2016, 2, 1–10. [Google Scholar]
  2. Benami, A. Effect of CZTS Parameters on Photovoltaic Solar Cell from Numerical Simulation. J. Energy Power Eng. Sci. 2019, 13, 32–36. [Google Scholar]
  3. Gupta, G.K.; Dixit, A. Simulation Studies of CZT(S,Se) Single and Tandem Junction Solar Cells Towards Possibilities for Higher Efficiencies up to 22%. arXiv 2018, arXiv:1801.08498. Available online: https://arxiv.org/ftp/arxiv/papers/1801/1801.08498.pdf (accessed on 28 March 2022).
  4. Jimbo, K.; Kimura, R.; Kamimura, T.; Yamada, S.; Maw, W.S.; Araki, H.; Oishi, K.; Katagiri, H. Cu2ZnSnS4-type Thin-film Solar Cells Using Abundant Materials. Thin Solid Film 2007, 515, 5997–5999. [Google Scholar] [CrossRef]
  5. Wang, J.; Xin, X.; Lin, Z. Cu2ZnSnS4 nanocrystals and graphene quantum dots for photovoltaics. Nanoscale 2011, 3, 3040–3048. [Google Scholar] [CrossRef]
  6. Bakr, N.A.; Khodair, Z.T.; Mahdi, H.I. Influence of Thiourea Concentration on Some Physical Properties of Chemically Sprayed Cu2ZnSnS4 Thin Films. Int. J. Mater. Sci. 2016, 5, 261–270. [Google Scholar]
  7. Swati, S.I.; Matin, R.; Bashar, S.; Mahmood, Z.H. Experimental study of the optical properties of Cu2ZnSnS4 thin-film absorber layer for solar cell application. J. Phys. Conf. Ser. 2018, 1086, 012010. [Google Scholar] [CrossRef]
  8. Katagiri, H.; Jimbo, K.; Yamada, S.; Kamimura, T.; Maw, W.S.; Fukano, T.; Ito, T.; Motohiro, T. Enhanced Conversion Efficiencies of Cu2ZnSnS4-Based Thin-film Solar Cells by Using Preferential Etching Technique. Appl. Phys. Express 2008, 1, 041201. [Google Scholar] [CrossRef]
  9. Sethi, V.K.; Pandey, M.; Shukla, P. Use of Nanotechnology in Solar PV Cell. Int. J. Chem. Eng. Appl. 2011, 2, 2. Available online: http://www.ijcea.org/papers/79-A555.pdf (accessed on 28 March 2022). [CrossRef] [Green Version]
  10. Michallon, J.; Bucci, D.; Morand, A.; Zanuccoli, M.; Consonni, V.; Cachopo, A.K. Light trapping in ZnO nanowires arrays covered with an absorbing shell for solar cells. Opt. Express 2014, 22, A1174–A1189. [Google Scholar] [CrossRef] [Green Version]
  11. Anwar, F. Simulation and Performance Study of Nanowires CdS/CdTe Solar Cell. Int. J. Renew. Energy Res. 2016, 7, 885–893. [Google Scholar] [CrossRef]
  12. Malek, M.F.; Sahdan, M.Z.; Mamat, M.H.; Musa, M.Z.; Khusaimi, Z.; Husairi, S.S.; Sin, N.D.; Rusop, M. A novel fabrication of MEH-PPV/Al:ZnO nanorod arrays based ordered bulk heterojunction hybrid solar cells. Appl. Surf. Sci. 2013, 275, 75–83. [Google Scholar] [CrossRef]
  13. Saurdi, I.; Mamat, M.H.; Malek, M.F.; Rusop, M. Preparation of Aligned ZnO Nanorod Arrays on Sn-Doped ZnO Thin Films by Sonicated Sol-Gel Immersion Fabricated for Dye-Sensitized Solar Cell. Adv. Mater. Sci. Eng. 2014, 2014, 636725. [Google Scholar] [CrossRef] [Green Version]
  14. Rouhi, J.; Mamat, M.H.; Raymond, O.C.H.; Mahmud, S.; Mahmood, M.R. Based on Morphology-Controllable Synthesis of ZnO–ZnS Heterostructure Nanocone Photoanodes. PLoS ONE 2015, 10, e0123433. [Google Scholar] [CrossRef] [Green Version]
  15. Ding, M.; Yao, N.; Wang, C.; Huang, J.; Shao, M.; Zhang, S.; Li, P.; Deng, X.; Xu, X. ZnO@CdS Core-Shell Heterostructures: Fabrication, Enhanced Photocatalytic, and Photoelectrochemical Performance. Nanoscale Res. Lett. 2016, 11, 205. [Google Scholar] [CrossRef] [Green Version]
  16. Kumar, K.; Pattanaik, S.; Dash, S.K. Modeling of the Nanowires CdS-CdTe Device Design for Enhanced Quantum Efficiency in Window Absorber Type Solar Cells. Int. J. Sci. Res. Rev. 2019, 7, 2279–2543. [Google Scholar]
  17. Rokade, A.; Rondiya, S.; Date, A.; Sharma, V.; Prasad, M.; Pathan, H.; Jadkar, S. Electrochemical synthesis of core-shell ZnO/CdS nanostructure for photocatalytic water splitting application. In Proceedings of the 1st International Conference on Electronics Packaging, Power Energy, ICEP2016, RMIT University, Melbourne, Australia, 14–16 December 2016. [Google Scholar]
  18. Yur, K.; Zolotkov, A.; Unit Convertor. Convert Terahertz [THz] to Wavelength in Meters. December. 2021. Available online: https://www.translatorscafe.com/unit-converter/en-US/frequency-wavelength/4-27/terahertz-wavelength%20in%20metres/ (accessed on 28 March 2022).
  19. Liu, Q.; Sandgren, E.; Barnhart, M.; Zhu, R.; Huang, G. Photonic Nanostructures Design and Optimization for Solar Cell Application. Photonics 2015, 2, 893–905. [Google Scholar] [CrossRef]
  20. Björkman, C.P.; Barreau, N.; Choubrac, M.B.L.; Grenet, L.; Heo, J.; Kubart, T.; Mittiga, A.; Sanchez, Y.; Scragg, J.; Sinha, S.; et al. Back and front contacts in kesterite solar cells: State-of-the-art and open questions. J. Phys. Energy 2019, 1, 044005. [Google Scholar] [CrossRef]
  21. Sun, W.; Brozak, M.; Armstrong, J.C.; Cui, J. Solar Cell Structures Based on ZnO/CdS core-shell nanowires Arrays Embedded in Cu2ZnSnS4 Light Absorber. In Proceedings of the IEEE Photovoltaic Specialists Conference (PVSC), 39th PVSC, Tampa, FL, USA, 16–21 June 2013; pp. 2042–2046. [Google Scholar]
  22. Çetinkaya, S. Study of electrical effect of transition-metal dichalcogenide-MoS2 layer on the performance characteristic of Cu2ZnSnS4 based solar cells using wxAMPS. Optik 2019, 181, 627–638. [Google Scholar] [CrossRef]
  23. Liu, X. Molybdenum Back Contact Treatment for Cu2ZnSnS4 Solar Cells. UNSW August 2015 Open ACCESS. Available online: http://unsworks.unsw.edu.au/fapi/datastream/unsworks:41049/SOURCE02?view=true (accessed on 28 March 2022).
  24. Vishwakarma, M.; Thota, N.; Karakulina, O.; Habermann, J.; Mehta, B.R. Role of graphene interlayer on the formation of the MoS2-CZTS interface during growth. AIP 2018, 1953, 100064. [Google Scholar] [CrossRef]
  25. Novoselov, K.S.; Falko, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef]
  26. Hu, Y.H.; Wang, H.; Hu, B. Thinnest Two-Dimensional Nanomaterial-Graphene for Solar Energy. ChemSusChem 2010, 3, 782–796. [Google Scholar] [CrossRef] [PubMed]
  27. Du, X.; Skachko, I.; Barker, A.; Andrei, E.Y. Approaching ballistic transport in suspended graphene. Nat. Nano 2008, 3, 491–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Han, B.; Hu, Y.H. MoS2 as a co-catalyst for photocatalytic hydrogen production from water. Energy Sci. Eng. 2016, 128, 285–304. [Google Scholar] [CrossRef] [Green Version]
  29. Jaramillo, T.F.; Jorgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Zhang, K.; Kim, J.K.; Ma, M.; Yim, S.Y.; Lee, C.L.; Shin, H.; Park, J.H. Delocalized Electron Accumulation at Nanorod Tips: Origin of Efficient H2 Generation. Adv. Funct. Mater. 2016, 26, 4527–4534. [Google Scholar] [CrossRef]
  31. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.V.; Dubonos, S.V.; Firsov, A.A. Two-Dimensional Gas of Massless Dirac Fermions in Graphene. Nature 2005, 438, 197–200. [Google Scholar] [CrossRef] [PubMed]
  32. Geim, A.K. Graphene: Status and Prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [Green Version]
  33. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and graphene oxide: Synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef]
  34. Britnell, L.; Ribeiro, R.M.; Eckmann, A.; Jalil, R.; Belle, B.D.; Mishchenko, A.; Kim, Y.J.; Gorbachev, R.V.; Georgiou, T.; Morozov, S.V.; et al. Strong Light-Matter Interactions in Heterostructures of Atomically Thin Films. Science 2013, 340, 1311–1314. [Google Scholar] [CrossRef] [Green Version]
  35. Jiang, J.W. Graphene versus MoS2: A short review. Front. Phys. 2015, 10, 287–302. [Google Scholar] [CrossRef] [Green Version]
  36. Seo, D.B.; Trung, T.N.; Bae, S.S.; Kim, E.T. Improved Photoelectrochemical Performance of MoS2 through Morphology-Controlled Chemical Vapor Deposition Growth on Graphene. Nanomaterials 2021, 11, 1585. [Google Scholar] [CrossRef]
  37. Dong, H.; Li, J.; Chen, M.; Wang, H.; Jiang, X.; Xiao, Y.; Tian, B.; Zhang, X. High-throughput Production of ZnO-MoS2 Graphene Heterostructures for Highly Efficient PhotocatalyticHydrogen Evolution. Materials 2019, 12, 2233. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, C.; Drame, B.; Niare, L.; Yuegang, F. Optimization of the Shell Thickness of the ZnO/CdS Core-Shell Nanowire Arrays in a CZTS Absorber. Int. J. Opt. 2022, 2022, 5301790. [Google Scholar] [CrossRef]
  39. Kim, H.U.; Kim, M.; Jin, Y.; Hyeon, Y.; Kim, K.S.; Anb, B.S.; Yangb, C.W.; Kanade, V.; Moonc, J.Y.; Yeom, G.Y.; et al. Low-temperature wafer-scale growth of MoS2-graphene heterostructures. Appl. Surf. Sci. 2019, 470, 129–134. [Google Scholar] [CrossRef]
  40. Liu, X.; Balla, I.; Bergeron, H.; Campbell, G.P.; Bedzyk, M.J.; Hersam, M.C. Rotationally Commensurate Growth of MoS2 on Epitaxial Graphene. ACS Nano 2016, 10, 1067–1075. [Google Scholar] [CrossRef] [Green Version]
  41. Anttu, N.; Xu, H.Q. Coupling of light into nanowire arrays and subsequent absorption. J. Nanosci. Nanotechnol. 2010, 10, 7183–7187. [Google Scholar] [CrossRef]
  42. Wen, L.; Zhao, Z.; Li, X.; Shen, Y.; Guo, H.; Wang, Y. Theoretical analysis and modeling of light trapping in high-efficiency GaAs nanowire array solar cells. Appl. Phys. Lett. 2011, 99, 143116. [Google Scholar] [CrossRef]
  43. Kupec, J.; Stoop, R.L.; Witzigmann, B. Light absorption and emission in nanowire array solar cells. Opt. Express 2010, 18, 27589–27605. [Google Scholar] [CrossRef]
  44. Gu, Z.; Prete, P.; Lovergine, N.; Nabet, B. On optical properties of GaAs and GaAs/AlGaAs core-shell periodic nanowire arrays. J. Appl. Phys. 2011, 109, 064314–064316. [Google Scholar] [CrossRef]
  45. Huang, N.; Lin, C.; Povinelli, M.L. Broadband absorption of semiconductor nanowire arrays for photovoltaic applications. J. Opt. 2012, 14, 024004. [Google Scholar] [CrossRef] [Green Version]
  46. Hu, L.; Chen, G. Analysis of optical absorption in silicon nanowire arrays for photovoltaic applications. Nano Lett. 2007, 7, 3249–3252. [Google Scholar] [CrossRef] [PubMed]
  47. Lin, C.; Povinelli, M.L. Optical absorption enhancement in silicon nanowire arrays with a large lattice constant for photovoltaic applications. Opt. Express 2009, 17, 19371–19381. [Google Scholar] [CrossRef] [PubMed]
  48. Li, J.; Yu, H.; Wong, S.M.; Li, X.; Zhang, G.; Lo, P.G.Q.; Kwong, D.L. Design guidelines of periodic Si nanowire arrays for solar cell application. Appl. Phys. Lett. 2009, 95, 243113. [Google Scholar] [CrossRef]
  49. Li, J.; Yu, H.; Li, Y. Solar energy harnessing in hexagonally arranged Si nanowire arrays and effects of array symmetry on optical characteristics. Nanotechnology 2012, 23, 194010. [Google Scholar] [CrossRef] [PubMed]
  50. Anttu, N.; Xu, H.Q. Efficient light management in vertical nanowire arrays for photovoltaics. Opt. Express 2013, 21, A558–A575. [Google Scholar] [CrossRef] [PubMed]
  51. Dalapati, G.K.; Zhuk, S.; Panah, S.M.; Kushwaha, A.; Seng, H.L.; Chellappan, V.; Suresh, V.; Su, Z.; Batabyal, S.K.; Tan, C.C.; et al. Impact of molybdenum out-diffusion and interface quality on the performance of sputter grown CZTS based solar cells. Sci. Rep. 2017, 2, 1350. [Google Scholar] [CrossRef] [PubMed]
  52. Jhuma, F.A.; Shaily, M.Z.; Rashid, M.J. Towards High-Efficiency CZTS Solar Cell through Buffer Layer Optimization Mater. Renew. Sustain. Energy 2019, 8, 6. Available online: https://link.springer.com/article/10.1007/s40243-019-0144-1 (accessed on 28 March 2022). [CrossRef] [Green Version]
  53. Nair, R.R.; Blake, P.; Grigorenko, A.N.; Novoselov, K.S.; Booth, T.J.; Stauber, T.; Peres, N.M.R.; Geim, A.K. Fine structure constant defines visual transparency of graphene. Science 2008, 320, 1308. [Google Scholar] [CrossRef] [Green Version]
  54. Zhang, M.; Zhang, X. Ultrasensitive optical absorption in graphene-based on bound states in the continuum. Sci. Rep. 2015, 5, 8266. [Google Scholar] [CrossRef]
  55. Wang, Q.; Ouyang, Z.; Lin, M.; Zheng, Y. High-Quality Graphene-Based Tunable Absorber Based on Double-Side Coupled-Cavity Effect. Nanomaterials 2021, 11, 2824. [Google Scholar] [CrossRef]
  56. Shin, D.M.; Tsege, E.L.; Kang, S.H.; Seung, W.; Kim, S.W.; Kim, H.K.; Hong, S.W.; Hwang, Y.H. Freestanding ZnO nanorod/graphene/ZnO nanorod epitaxial double-heterostructure for improved piezoelectric nanogenerators. Nano Energy 2015, 12, 268–277. [Google Scholar] [CrossRef]
  57. Introduction to Comsol Multiphysics 5.3, 1998–2017. pp. 6–153. Available online: https://doc.comsol.com/5.3a/doc/com.comsol.help.comsol/IntroductionToCOMSOLMultiphysics.pdf (accessed on 20 February 2020).
  58. Semiconductor Module User’s Guide—Comsol Multiphysics 5.4, 1998–2018. pp. 11–264. Available online: https://doc.comsol.com/5.4/doc/com.comsol.help.semicond/SemiconductorModuleUsersGuide.pdf (accessed on 20 February 2020).
  59. RF Module User’s Guide-Comsol Multiphysics 5.3, 1998–2017. Available online: https://doc.comsol.com/5.4/doc/com.comsol.help.rf/RFModuleUsersGuide.pdf (accessed on 12 December 2020).
  60. Kavakl, I.G.; Kantarli, K. Single and Double-Layer Antireflection Coatings on Silicon. Turk. J. Phys. 2002, 26, 349–354. [Google Scholar]
  61. Kosoboutskyy, P.; Karkulovska, M.; Morgulis, A. The Principle of Multilayer Plane-Parallel Structure Antireflection. Opt. Appl. 2010, XL, 4. Available online: https://www.dbc.wroc.pl/dlibra/publication/98216/edition/58485/content?ref=struct (accessed on 28 March 2022).
  62. Guo, H.; Wen, L.; Li, X.; Zhao, Z.; Wang, Y. Analysis of Optical Absorption in GaAs Nanowires Arrays. Nano Res. Lett. 2011, 6, 617. Available online: http://www.nanoscalereslett.com/content/6/1/617 (accessed on 28 March 2022). [CrossRef] [Green Version]
  63. Wilson, D.P.; Lapierre, R.R. Simulation of optical absorption in conical nanowires. Opt. Express 2021, 29, 9544–9552. [Google Scholar] [CrossRef]
  64. Oriel Product Training—On Solar Simulation. Available online: https://www.semanticscholar.org/paper/ORIEL-PRODUCT-TRAINING-Solar-Simulation/9dcad5adc1e65f9f3099d4e5e74a2e6c9d76069a (accessed on 25 November 2021).
  65. Aghaeipour, M.; Pettersson, H. Enhanced broadband absorption in nanowire arrays with integrated Bragg reflectors. Nanophotonics 2018, 7, 819–825. [Google Scholar] [CrossRef] [Green Version]
  66. Byrnes, S.J. Multilayer Optical Calculations. arXiv 2021, arXiv:1603.02720. Available online: https://arxiv.org/pdf/1603.02720.pdf (accessed on 3 January 2022).
  67. Refractive Index Database of Materials. Available online: https://refractiveindex.info/ (accessed on 20 November 2020).
  68. Rohatgi, A. WebPlotDigitizer User Manual, Version 4.3, July 2020. Available online: https://automeris.io/WebPlotDigitizer/userManual.pdf (accessed on 4 January 2022).
  69. Sturmberg, B.C.P.; Dossou, K.B.; Botten, L.C.; Asatryan, A.A.; Poulton, C.G.; Sterke, C.M.; McPhedran, R.C. Modal analysis of enhanced absorption in silicon nanowire arrays. Opt. Express 2011, 19, A1067–A1081. [Google Scholar] [CrossRef] [Green Version]
  70. Tivanov, M.; Moskalev, A.; Kaputskaya, I.; Żukowski, P. Calculation of the Ultimate Efficiency of p-n-Junction Solar Cells Taking into Account the Semiconductor Absorption Coefficient. Prz. Elektrotech. 2016, 92, 85–87. Available online: https://sigma-not.pl/publikacja-99983-2016-8.html (accessed on 28 March 2022). [CrossRef]
  71. Wang, W.; Winkler, M.T.; Gunawan, O.; Gokmen, T.; Todorov, T.K.; Zhu, Y.; Mitzi, D.B. Device characteristics of CZTSSe thin-film solar cells with 12.6% efficiency. Adv. Energy Mater. 2014, 4, 1301465. [Google Scholar] [CrossRef]
  72. Guo, S.D. Phonon Transport in Janus monolayer MoSSe: A first-principles study. Phys. Chem. Chem. Phys. 2018, 20, 7236–7242. [Google Scholar] [CrossRef] [Green Version]
  73. Peng, B.; Zhang, H.; Shao, H.; Xu, Y.; Zhang, X.; Zhu, H. Thermal conductivity of monolayer MoS2, MoSe2, and WS2: Interplay of mass effect, interatomic bonding, and anharmonicity. RSC Adv. 2016, 6, 5767. [Google Scholar] [CrossRef] [Green Version]
  74. Cozza, D.; Ruiz, C.M.; Duché, D.; Giraldo, S.; Saucedo, E.; Simon, J.J.; Escoubas, L. Optical modeling and optimizations of Cu2ZnSnSe4 solar cells using the modified transfer matrix method. Opt. Express 2016, 24, A1201–A1209. [Google Scholar] [CrossRef]
  75. Song, N.; Green, M.A.; Huang, J.; Hu, Y.; Hao, X. Study of sputtered Cu2ZnSnS4 thin films on Si. Appl. Surf. Sci. 2018, 459, 700–706. [Google Scholar] [CrossRef]
  76. Reya, G.; Larramonab, G.; Bourdaisb, S.; Chońeb, C.; Delatoucheb, B.; Jacobb, A.; Dennlerb, G.; Siebentritta, S. On the origin of band-tails in kesterite. Sol. Energy Mater. Sol. Cells 2018, 179, 142–151. [Google Scholar] [CrossRef]
  77. Cozza, D.; Ruiz, C.; Duché, M.; Simon, J.J.; Escoubas, L. Modeling the back contact of Cu2ZnSnS4 solar cells. IEEE J. Photovolt. 2016, 6, 1292–1297. [Google Scholar] [CrossRef]
  78. Zhang, Z.; Cui, T.; Lv, R.; Zhu, H.; Wang, K.; Wu, D.; Kang, F. Improved efficiency of graphene/Si heterojunction solar cells by optimizing hydrocarbon feed rate. J. Nanomater. 2014, 2014, 359305. [Google Scholar] [CrossRef]
  79. Miao, X.; Tongay, S.; Petterson, M.K.; Berke, K.; Rinzler, A.G.; Appleton, B.R.; Hebard, A.F. High-efficiency graphene solar cells by chemical doping. Nano Lett. 2012, 12, 2745–2750. [Google Scholar] [CrossRef] [Green Version]
  80. Zhong, S.; Hua, X.; Shen, W.; Hua, X. Simulation of high-efficiency crystalline silicon solar cells with homo–heterojunctions. IEEE Trans. Electron. Devices 2013, 60, 2104–2110. [Google Scholar] [CrossRef]
  81. Gangopadhyay, U.; Roy, S.; Garain, S.; Jana, S.; Das, S. Comparative simulation study between n-type and p-type silicon solar cells and the variation of efficiency of n-type solar cell by the application of passivation improvement layer with different thickness using AFORS HET and PC1D. IOSR J. Eng. 2012, 2, 41–48. [Google Scholar] [CrossRef]
  82. Patel, K.; Tyagi, P.K. Multilayer graphene as a transparent conducting electrode in silicon heterojunction solar cells. AIP Adv. 2015, 5, 077165. [Google Scholar] [CrossRef] [Green Version]
  83. Zandi, S.; Saxena, P.; Razaghi, M.; Gorjic, N.E. Simulation of CZTSSe thin-film solar cells in COMSOL: 3D coupled optical, electrical, and thermal model. IEEE J. Photovolt. 2020, 10, 1503–1507. [Google Scholar] [CrossRef]
  84. Huang, P.; Riccardi, E.; Messelot, S.; Graef, H.; Valmorra, F.; Tignon, J.; Taniguchi, T.; Watanabe, K.; Dhillon, S.; Plaçais, B.; et al. Ultra-long carrier lifetime in neutral graphene-hBN van der Waals heterostructures under mid-infrared illumination. Nat. Commun. 2020, 11, 863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Cismaru, A.; Dragoman, M.; Dinescu, A.; Dragoman, D.; Stavrinidis, G.; Konstantinidis, G. Microwave and millimeter-wave electrical permittivity of graphene monolayer. arXiv 2013, arXiv:1309.0990. [Google Scholar]
  86. Pop, E.; Varshney, V.; Roy, A.K. Thermal properties of graphene: Fundamentals and applications. Mater. Res. Soc. Symp. Proc. 2012, 37, 1273–1281. [Google Scholar] [CrossRef] [Green Version]
  87. Jahan, F.; Islam, M.H.; Smith, B.E. Bandgap and refractive index determination of Mo-black coatings using several techniques. Sol. Energy Mater. Sol. Cells 1995, 37, 283–293. [Google Scholar] [CrossRef]
  88. Latrous, A.R.; Mahamdi, R.; Touafek, N.; Pasquinelli, M. Performance Enhancement in CZTS Solar Cells by SCAPS-1D. Int. J. Thin. Fil. Sci. Techol. 2021, 10, 75–81. [Google Scholar]
  89. Jhuma, F.A.; Rashid, M.J. Simulation study to find suitable dopants of CdS buffer layer for CZTS solar cell. J. Theor. Appl. Phys. 2020, 14, 75–84. [Google Scholar] [CrossRef] [Green Version]
  90. Jabeen, M.; Haxha, S. High rear reflectance and light trapping in textured graphene-based silicon thin-film solar cells with back dielectric-metal reflectors. OSA Contin. 2019, 2, 1807–1821. [Google Scholar] [CrossRef]
  91. Nakamura, R.; Tanaka, K.; HisaoUchikiHisao, H.; Nakamura, R.; Jimbo, K.; Washio, T.; Katagiri, H. Cu2ZnSnS4 thin film deposited by sputtering with Cu 2 ZnSnS 4 compound target. Jpn. J. Appl. Phys. 2014, 53, 02BC10. [Google Scholar] [CrossRef]
  92. Hanson, G.W. Dyadic Green’s functions and guided surface waves for a surface conductivity model of graphene. J. Appl. Phys. 2008, 103, 064302. [Google Scholar] [CrossRef] [Green Version]
  93. Zhong, H.; Liu, Z.; Liu, X.; Fu, G.; Liu, G.; Chen, J.; Tang, C. Ultra-high quality graphene perfect absorbers for high performance switching manipulation. Opt. Express 2020, 28, 37294–37306. [Google Scholar] [CrossRef]
  94. Wang, D.W.; Zhao, W.S.; Gu, X.Q.; Chen, W.; Fellow, W.Y.Y. Wideband Modeling of Graphene-Based Structures at Different Temperatures Using Hybrid FDTD Method. IEEE Trans. Nanotechnol. 2015, 14, 250–258. [Google Scholar] [CrossRef]
  95. Xu, S.; Zhang, L.; Wang, B.; Ruoff, R.S. Chemical vapor deposition of graphene on thin-metal films. Cell Rep. Phys. Sci. 2021, 24, 100372. [Google Scholar] [CrossRef]
  96. Hodes, G. Semiconductor and ceramic nanoparticle films deposited by chemical bath deposition. Phys. Chem. Chem. Phys. 2007, 9, 2181–2196. [Google Scholar]
  97. Hamel, A.; Chibani, A. Characterization of texture surface for solar cells. J. Appl. Sci. 2010, 10, 231–234. [Google Scholar] [CrossRef]
  98. Shockley, V.; Read, W.T. Statistics of the recombination of holes and electrons. Phys. Rev. 1952, 87, 835–842. [Google Scholar] [CrossRef]
  99. Hall, R.N. Electron-hole recombination in germanium. Phys. Rev. 1952, 87, 387. [Google Scholar] [CrossRef]
  100. Hingerl, J. Data Analysis for Nanomaterials: Effective Medium Approximation, Its Limits and ImplementationsK. Ellipsometry at the Nanoscale March 2013. pp. 145–177. Available online: http://old.ceitec.eu/2-data-analysis/f33667 (accessed on 2 February 2020).
  101. Permittivity, Maxwell’s Equations. 2012. Available online: https://maxwells-equations.com/materials/permittivity.php (accessed on 1 December 2020).
  102. ElsayedGhitas, A. Studying the effect of spectral variations intensity of the incident solar radiation on the Si solar cells performance. NRIAG J. Astron. Geophys. 2012, 1, 165–171. [Google Scholar]
  103. Nair, K.K.; Jose, J.; Ravindran, A. Analysis of temperature-dependent parameters on solar cell efficiency using MATLAB. Int. J. Eng. Dev. Res. 2016, 4, 2321–9939. [Google Scholar]
  104. Honsberg, C.; Bowden, S. Diffusion Length. PV Education. Available online: https://www.pveducation.org/pvcdrom/pn-junctions/diffusion-length (accessed on 10 December 2020).
  105. Al-Jumaili, M.H.; Abdalkafor, A.S.; Taha, M.Q. Analysis of the hard and soft shading impact on photovoltaic module performance using solar module tester. Int. J. Eng. Dev. Res. 2019, 10, 1014–1021. [Google Scholar] [CrossRef]
  106. Su, R.R.; Yu, Y.X.; Xiao, Y.H.; Yang, X.F.; Zhang, W.D. Earth-abundant ZnO/CdS/CuSbS2 core-shell nanowire arrays as highly efficient photoanode for hydrogen evolution. Int. J. Hydrog. Energy 2018, 43, 6040–6048. [Google Scholar] [CrossRef]
  107. Tak, Y.; Hong, S.J.; Leeb, J.S.; Yong, K. Fabrication of ZnO/CdS core/shell nanowire arrays for efficient solar energy conversion. J. Mater. Chem. 2009, 19, 5945–5951. [Google Scholar] [CrossRef] [Green Version]
  108. Salazar, R.; Delamoreanub, A.; Clémentc, C.L.; Ivanova, V. ZnO/CdTe and ZnO/CdS core-shell nanowire arrays for extremely thin absorber solar cells. Energy Procedia 2011, 10, 122–127. [Google Scholar] [CrossRef] [Green Version]
  109. Van Dam, D.; Van Hoof, N.J.; Cui, Y.; Van Veldhoven, P.J.; Bakkers, E.P.; Gómez Rivas, J.; Haverkort, J.E. High-efficiency nanowire solar cells with omnidirectionally enhanced absorption due to self-aligned indium-tin-oxide Mie scatterers. ACS Nano 2016, 10, 11414–11419. [Google Scholar] [CrossRef]
  110. Khvol’son, O.D. Course of Physics; Gosizdat RSFSR: Berlin, Germany, 1923; Volume 2. [Google Scholar]
  111. Shockley, W.; Queisser, H.J. Detailed balance limit of efficiency of p-n junction solar cells. J. Appl. Phys. 1961, 32, 510–519. [Google Scholar] [CrossRef]
  112. Lee, K.S.; Chung, Y.D.; Park, N.M.; Cho, D.H.; Kim, K.H.; Kim, J. Analysis of the Current-voltage Curves of a Cu(In,Ga)Se2 Thin-film Solar Cell Measured at Different Irradiation Conditions. J. Opt. Soc. Korea 2010, 14, 321–325. [Google Scholar] [CrossRef] [Green Version]
  113. Varshni, Y.P. Temperature dependence of the energy gap in semiconductors. Physica 1967, 34, 149–154. [Google Scholar] [CrossRef]
  114. Han, J.Z.; Chen, R.S. Tunable broadband terahertz absorber based on a single-layer graphene metasurface. Opt. Express 2020, 28, 30289. [Google Scholar] [CrossRef] [PubMed]
  115. Landau, L.D.; Lifshitz, E.M. Electrodynamics of Continuous Media. Section 9, 2nd ed.Pergamon Press: Oxford, UK, 1984. [Google Scholar]
  116. Looyenga, H. Dielectric Constants of Heterogeneous Mixtures. Physica 1965, 31, 401. [Google Scholar] [CrossRef]
  117. Jackson, J.D. Classical Electrodynamics; Wiley: New York, NY, USA, 1962. [Google Scholar]
  118. Fowler, M. Maxwell’s Equations and Electromagnetic Waves. Physics Department, UVa. Available online: http://galileo.phys.virginia.edu/classes/109N/more_stuff/Maxwell_Eq.html (accessed on 22 April 2022).
  119. Sinton, R.A.; Cuevas, A. Contactless determination of current-voltage characteristics and minority-carrier lifetimes in semiconductors from quasi-steady-state photoconductance data. Appl. Phys. Lett. 1996, 69, 2510–2512. [Google Scholar] [CrossRef]
  120. Sabbar, E.H.; Saleh, M.H.; Salih, S.M.; Salih, S.H. Deposition of SeTe/Si Thin Film via Thermal Evaporation. Am. J. Condens. Matter Phys. 2013, 3, 119–122. [Google Scholar]
Figure 1. CZTS solar cell structures with ZnO/CdS core-shell nanowire arrays of (a) the basic model and of (b) the single nanowire model.
Figure 1. CZTS solar cell structures with ZnO/CdS core-shell nanowire arrays of (a) the basic model and of (b) the single nanowire model.
Crystals 12 00772 g001
Figure 2. Basic model of CZTS solar cell structures with ZnO/CdS core-shell nanowire arrays and a graphene interlayer.
Figure 2. Basic model of CZTS solar cell structures with ZnO/CdS core-shell nanowire arrays and a graphene interlayer.
Crystals 12 00772 g002
Figure 3. Structural models (a) with graphene interlayer, (b) without graphene interlayer of CZTS solar cell with ZnO/CdS core-shell nanowires arrays.
Figure 3. Structural models (a) with graphene interlayer, (b) without graphene interlayer of CZTS solar cell with ZnO/CdS core-shell nanowires arrays.
Crystals 12 00772 g003aCrystals 12 00772 g003b
Figure 4. Schematic diagram of the mesh structures of the single nanowire of the CZTS solar cell with ZnO/CdS core-shell nanowire arrays.
Figure 4. Schematic diagram of the mesh structures of the single nanowire of the CZTS solar cell with ZnO/CdS core-shell nanowire arrays.
Crystals 12 00772 g004
Figure 5. Schematic diagram of the mesh of structural models with the graphene interlayer (a) and without the graphene interlayer (b) of the CZTS solar cell with ZnO/CdS core-shell nanowire arrays.
Figure 5. Schematic diagram of the mesh of structural models with the graphene interlayer (a) and without the graphene interlayer (b) of the CZTS solar cell with ZnO/CdS core-shell nanowire arrays.
Crystals 12 00772 g005
Figure 6. Electric field in CZTS solar cell with single core-shell ZnO/CdS nanowires of ZnO core diameter/CdS shell thickness: (a) 70 nm/45 nm, (b) 100 nm/30 nm, and (c) 120 nm/20 nm.
Figure 6. Electric field in CZTS solar cell with single core-shell ZnO/CdS nanowires of ZnO core diameter/CdS shell thickness: (a) 70 nm/45 nm, (b) 100 nm/30 nm, and (c) 120 nm/20 nm.
Crystals 12 00772 g006
Figure 7. Absorption as a function of wavelength in CZTS solar cell with single core-shell ZnO/CdS nanowires of ZnO core diameter/CdS shell thickness: 70 nm/45 nm, 100 nm/30 nm, and 120 nm/20 nm.
Figure 7. Absorption as a function of wavelength in CZTS solar cell with single core-shell ZnO/CdS nanowires of ZnO core diameter/CdS shell thickness: 70 nm/45 nm, 100 nm/30 nm, and 120 nm/20 nm.
Crystals 12 00772 g007
Figure 8. Reflection as a function of wavelength in CZTS solar cell with single core-shell ZnO/CdS nanowires of ZnO core diameter/CdS shell thickness: 70 nm/45 nm, 100 nm/30 nm, and 120 nm/20 nm.
Figure 8. Reflection as a function of wavelength in CZTS solar cell with single core-shell ZnO/CdS nanowires of ZnO core diameter/CdS shell thickness: 70 nm/45 nm, 100 nm/30 nm, and 120 nm/20 nm.
Crystals 12 00772 g008
Figure 9. Electric field along with height in CZTS solar cells with ZnO/CdS core-shell nanowire arrays: (a) with graphene interlayer and (b) without graphene interlayer.
Figure 9. Electric field along with height in CZTS solar cells with ZnO/CdS core-shell nanowire arrays: (a) with graphene interlayer and (b) without graphene interlayer.
Crystals 12 00772 g009
Figure 10. Absorption as a function of wavelength for models (with and without graphene interlayer).
Figure 10. Absorption as a function of wavelength for models (with and without graphene interlayer).
Crystals 12 00772 g010
Figure 11. Reflection as a function of wavelength, with graphene interlayer and without graphene interlayer.
Figure 11. Reflection as a function of wavelength, with graphene interlayer and without graphene interlayer.
Crystals 12 00772 g011
Figure 12. Total generation rate of electron–hole pairs in the cell (with and without graphene interlayer) as a function of its height. The rear surface of the cell is at 0 nm.
Figure 12. Total generation rate of electron–hole pairs in the cell (with and without graphene interlayer) as a function of its height. The rear surface of the cell is at 0 nm.
Crystals 12 00772 g012
Figure 13. Total recombination rate of electron–hole pairs in the cell (with and without graphene interlayer) as a function of its height. The rear surface of the cell is at 0 nm.
Figure 13. Total recombination rate of electron–hole pairs in the cell (with and without graphene interlayer) as a function of its height. The rear surface of the cell is at 0 nm.
Crystals 12 00772 g013
Figure 14. Current density as a function of voltage for the models with and without graphene interlayer.
Figure 14. Current density as a function of voltage for the models with and without graphene interlayer.
Crystals 12 00772 g014
Table 1. The values of the electrical and thermal parameters used in COMSOL as partially reported in Refs [1,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92].
Table 1. The values of the electrical and thermal parameters used in COMSOL as partially reported in Refs [1,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92].
ParametersMoGrapheneMoS2CZTSCdSZnOITO
W (nm)20011701000100140150
εr-3.31013.6999
Nc (cm−3)-1 × 10180.7 × 10182.2 × 10182.2 × 10182.2 × 10182.2 × 1018
Nv (cm−3)-1 × 10190.3 × 10191.8 × 10191.8 × 10191.8 × 10191.8 × 1019
μn (cm2/V·s)-1 × 10910010030100100
χ (eV)-3.54.144.24.24.43.6
Eg (eV)1.960–0.251.11.452.43.33.6
CB (cm−3)2.2 × 10183 × 109–3 × 10222.2 × 10182.2 × 10182.2 × 10192.2 × 10194 × 1019
VB (cm−3)1.8 × 10193 × 109–3 × 10221.8 × 10191.8 × 10191.8 × 10181.8 × 10191 × 1019
NA (cm−3)--1 × 10151 × 10161 × 10131 × 10131 × 1013
ND (cm−3)-1 × 10195 × 10135 × 10135 × 10175 × 10181 × 1019
CB (cm−3)2.2 × 10181 × 10222.2 × 10182.2 × 10182.2 × 10192.2 × 10192.2 × 1019
VB (cm−3)1.8 × 10191 × 10221.8 × 10191.8 × 10191.8 × 10181.8 × 10191.8 × 1019
Tvn (cms−1)1 × 1071 × 1071 × 1071 × 1071 × 1071 × 1071 × 107
Tvp (cms−1)1 × 1071 × 1071 × 1071 × 1071 × 1071 × 1071 × 107
D (g·cm−1) 4.692.3286.94.564.825.617.12
K (W/m·K)1381401032.956.223.410
Cp (Jg–1 K–1)0.2770.70.120.410.210.4040.36
τn/τp (ns)-0.03/0.032.4/2.45.4/5.40.005/0.0050.003/0.0030.003/0.003
h (W/m2K)363 × 10630 × 102605 × 10614 × 104124 × 104468 × 10633 × 106
h (W/m2K)363 × 10630 × 102605 × 10614 × 104124 × 104468 × 10633 × 106
Cn (cm6 s−1)0001 × 10−291 × 10−291 × 10−291 × 10−29
Cp (cm6 s−1)0001 × 10−291 × 10−291 × 10−291 × 10−29
C (cm3 s−1)0005 × 10−95 × 10−95 × 10−95 × 10−9
W: Thickness. εr: Relative permittivity. Nc: Effective conduction band density. Nv: Effective valence band density. μn: Electron mobility. χ: Electron Affinity. Eg: Bandgap. CB: effective density of states. VB: effective density of states. NA: Acceptor concentration. ND: Donor concentration. CB: Effective conduction band density. VB: Effective valence band density. Tvn: Thermal velocity of electrons. Tvp: Thermal velocity of holes. D: Density. K: Thermal Conductivity. Cp: Specific heat. τn/τp: Lifetime electron/hole. h: Heat transfer coefficient. Cn: Auger recombination coefficient for electron. Cp: Auger recombination coefficient for hole. C: Direct band-to-band recombination coefficient.
Table 2. Comparative results of electric fields, absorptions, and reflections of different models with a single nanowire.
Table 2. Comparative results of electric fields, absorptions, and reflections of different models with a single nanowire.
Single Nanowire Model-(Core Diameter/Shell Thickness)Maximum ElectricField in the Nanowire
(V/m)
Maximum Absorption in the NanowireMaximum Reflection in the Nanowire
70 nm/45 nm100089% at 390 nm64% at 1820 nm
100 nm/30 nm130096% at 500 nm 43% at 1800 nm
120 nm/20 nm120094% at 420 nm62% at 1810 nm
Table 3. The numerical value of Jsc, Voc, FF, and EFF in the models with and without a graphene interlayer.
Table 3. The numerical value of Jsc, Voc, FF, and EFF in the models with and without a graphene interlayer.
Model of Solar CellVoc (V)Jsc (mA/cm2)FF (%)EFF (%)
With graphene interlayer0.636.281.716.6
Without graphene interlayer0.66.179.215.07
Table 4. Comparative results with previous works.
Table 4. Comparative results with previous works.
Physical Model ofVoc (V)Jsc (mA/cm2)FF (%)EFF (%)Max Absoption (nm)Wavelength (nm)
W. Sun et al. [21]0.170.729.410.035-500
R. Su et al. [56]-6.48-52-480
Y. Tak et al. [57]1.557.2364.513.53-500
C. Wang et al. [38]0.636.3941.216.897%500
M. Malek et al. [12]0.1955.3227.710.287
I. Saurdi et al. [13]0.5522.8400.3810,599
J. Rouhi et al. [14]0.6812.20.494.07
Our study0.636.281.716.697.8500
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, C.; Drame, B.; Niare, L.; Yuegang, F. Simulation of a New CZTS Solar Cell Model with ZnO/CdS Core-Shell Nanowires for High Efficiency. Crystals 2022, 12, 772. https://doi.org/10.3390/cryst12060772

AMA Style

Wang C, Drame B, Niare L, Yuegang F. Simulation of a New CZTS Solar Cell Model with ZnO/CdS Core-Shell Nanowires for High Efficiency. Crystals. 2022; 12(6):772. https://doi.org/10.3390/cryst12060772

Chicago/Turabian Style

Wang, Chonge, Boubacar Drame, Lucien Niare, and Fu Yuegang. 2022. "Simulation of a New CZTS Solar Cell Model with ZnO/CdS Core-Shell Nanowires for High Efficiency" Crystals 12, no. 6: 772. https://doi.org/10.3390/cryst12060772

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop