Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Numerical modeling of ultra-thin CuSbS2 heterojunction solar cell with TiO2 electron transport and CuAlO2:Mg BSF layers

Open Access Open Access

Abstract

The ternary chalcostibite copper antimony sulfide (CuSbS2) system, with its very high optical absorption coefficient, low-cost, vacuum-free fabrication techniques, and earth-abundant elements, is a rising candidate as solar absorber material for ultrathin film solar cells. However, due to the Schottky barrier formed at the back-contact and high carrier recombination at the CuSbS2/CdS interface, the efficiency of conventional CuSbS2/CdS heterojunction solar cell is very poor. This article proposes titanium dioxide (TiO2) as an alternative to CdS layer for the CuSbS2-based thin film solar cells (TFSCs). Using TiO2, CuSbS2, and Mg-doped CuAlO2 (CuAlO2:Mg) as an electron transport layer (ETL), absorber layer, and back-surface field (BSF) layer, respectively, a novel (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au)-based npp+ heterojunction solar cell has been designed and simulated by SCAPS-1D solar cell simulator. The effects of integrating the CuAlO2:Mg BSF layer on the PV responses of the CuSbS2-based heterojunction solar cell in terms of the built-in potential and the back-contact carrier recombination have been studied. In addition, an investigation on the influences of various device parameters viz. carrier concentration and thickness of each layer, back-contact metal work function, shunt and series resistance, and working temperature have been carried out systematically. The results are analyzed in correlation with the PV parameters of the device to optimize the efficiency of the proposed solar cell. The optimized CuSbS2-based solar cell shows good performance stability at high temperature, with a maximum efficiency of 23.05% (Voc = 969 mV, Jsc= 34.61 mA/cm2, FF = 68.71%).

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In response to global energy demand expected to reach terawatt (TW) level by 2050, extensive use of renewable energies is essential without affecting the environment. Solar energy is a kind of renewable energy, which can meet the world energy demand in TW level with a large amount of solar energy to spare [1]. A potentially effective way to use solar energy is to use photovoltaic (PV) devices to convert it directly into electrical energy. From the very beginning of the invention of the solar cell, it has been improving every day. The first generation crystalline-Si (c-Si) based devices has been a leader of mainstream the solar PV market. It occupy approximately 95% of the global PV market share, while all other thin films PV technologies (i.e. CdTe, Cu(In,Ga)Se2 (CIGS) and thin film Si) occupy only 5% [2]. However, at present, c-Si technology and c-Si solar cells have the limitations of thick and expensive Si wafer, high processing temperature, high-vacuum fabrication process and other complicated and expensive processing technologies. Over the past five decades, new cost-effective and earth-abundant thin film materials and device fabrication technologies have been explored and developed by many researchers to make more efficient and environment friendly solar cells at lower cost. During this period, many TFSCs technologies came out, of which very few such as CdTe, CIGS, and perovskite solar cells have got widespread attention because their efficiency crosses commercial demarcation. Despite remarkable advances of CdTe and CIGS technology, the use of toxic and rare earth materials in CdTe solar cells, material or resource constraints, and complicated fabrication process of CIGS technology are the delimiting factors for the industrial scalability of the aforementioned solar cells [3,4] and therefore, these materials need to replace for future clean energy technologies [5]. In the last 10 years, kesterite structure such as Cu2ZnSn(S/Se)4 with its non-toxic, low-cost, earth-abundant elements were investigated vigorously as PV absorbers to overcome the material bottlenecks in CIGS and CdTe technology. So far, the record experimental efficiency of CZTS solar cells is 13.0% in laboratory level [6]. As CZTS is a defect-prone system, the large Voc deficiency (Eg/q – Voc) in CZTS/CZTSe devices is due to high defects density and large potential fluctuations at the energy band edges. However, recently Atowar et al. [7,8] theoretically anticipated over 30% PCE of CZTS and CTSe-based solar cells by reducing non-radiative recombination losses through the formation of appropriate energy band structure at buffer/absorber interface and utilizing a suitable BSF layer. As a result, it is obvious that in order to find inexpensive, non-toxic, less defective, easily controllable stoichiometry, chemically stable and earth-abundant materials for efficient TFSCs, further exploration and development of PV absorber materials is needed.

Ternary Copper-antimony-sulfide and copper-bismuth-sulfide (commonly known as CAS and CBS, respectively) systems have been gaining attention in PV communities as an alternative active layer materials for TFSCs, due to their inexpensive, nontoxic and earth-abundant elements [915]. The first theoretical study on CuSbS2 thin films was conducted by Dufton et al. in 2012 [16] and identified CuSbS2 thin films as future absorber materials for TFSCs. In 2013, Kumar and Persson [17] affirmed the primary finding from Dufton and predicted very high optical absorption of CuSbS2 thin films in the visible and infrared region, which is due to localized p-like Sb state near to the conduction band edge. Four main phases of copper-antimonysulfide system are CuSbS2 (chalcostibite), Cu3SbS4 (famatinite), Cu12Sb4S13 (tetrahedrite), and Cu3SbS3 (skinnerite) with an energy band gap that depends on the crystal structure. Among all these phases, Cu12Sb4S13 and Cu3SbS4 with their low thermal conductivity are being studied as thermoelectric materials [18,19]. The less-explored I-V-VI2 chalcogenides (CuSbS2) thin films with basic optoelectronic properties such as near-ideal direct band gap ranging from 1.38-1.58 eV, a high optical absorption in broad spectrum (α ≈ 105 cm−1), and moderate acceptor concentration (1015–1018 cm−3) and the low melting point (551 °C) and chemically stable phase [20], seems to be an upcoming solar absorber material for PV applications [2123]. Concerning the production cost, the development of inexpensive and easily scalable PV absorber layers depends largely on growth techniques. Many synthesis routes including physical and chemical processes such as chemical bath deposition (CBD) [12], spray pyrolysis [24], heating glass/Sb2S3/Cu layer in low vacuum condition [25], hybrid ink method [26], electrodeposited Sb-Cu alloys over sulfur [27], co-evaporation of Cu, Sb and S pure elements [23], low-temperature atomic layer deposition (ALD) [28] and low-cost spin coating method [29], to prepare CuSbS2 thin films are reported in the literature. Therefore, as a potential active layer material for TFSCs, interest in CuSbS2 thin films is increasing due to its relatively inexpensive synthesis process than other TFSCs technology, non-toxicity of all elements, and Sb, which is more abundant and less demanding than indium and gallium. In addition, unlike CZT(S/Se) system, the theoretical calculations of CuSbS2 thin films shows that low-energy defects are far away from the band gap center, resulting in a low density of bulk recombination centers in CuSbS2 thin films [30].

Although there are many reports on the electrical, optical and structural properties of CuSbS2 thin films that have emphasized its potential as a PV absorber, so far only a few reports have demonstrated its practical use in PV devices [31]. Using, chemical bath deposition process, the first CuSbS2-based solar cells with Voc of 345 mV and very low Jsc of 0.2 mA/cm2 were reported in 2005 [32] and since then, different methods have been utilized to construct CuSbS2-based solar cells. However, all of the CuSbS2-based solar cells demonstrated so far have been struggling with very poor PCE largely below 2% [23,27,30,3336]. In 2016, Banu et al. demonstrated CuSbS2-based solar cells using a hybrid ink process (non-vacuum) and achieved the highest PCE of 3.22% with a Voc of only 470 mV [26]. It is widely acknowledged that the PV performance of CuSbS2-based PV devices is mostly limited by their defects in absorber layer, dominant interfacial recombination of minority carriers, and unfavourable heterojunction configurations [37]. Moreover, the effective photo-generated carrier separation in CuSbS2 thin films is limited by the short diffusion length of carriers [33], thereby resulting in a low Jsc.

CuSbS2-based solar cells reported so far use CdS as the pn junction partner, because the CuSbS2-based device structure is directly inherited from its CIGS counterpart. However, a “cliff-like” conduction band offset (CBO) is formed at the CuSbS2/CdS interface [30,38], which is not within the favourable region of 0 to -0.4 eV “spike-like” CBO [39]. This results in a low Voc in CuSbS2 based devices. Therefore, more study is needed to reduce the recombination of minority carriers and to align the energy band at the CuSbS2/ETL interface to improve the efficiency of CuSbS2-based PV devices [38]. Metal oxides are widely used in TFSCs as ETL. TiO2 is a wide band gap (Eg = 3.26 eV) n-type semiconductor with high electron affinity has been used successfully in perovskite and dye-sensitized solar cells [4042], thus it is a promising candidate as ETL layer for fabricating high-performance solar cells as well.

In this work, we demonstrate significant improvements of Voc, Jsc, and thereby, PCE in CuSbS2-based solar cells by using Mg-doped CuAlO2 thin films as the BSF layer and TiO2 thin films as ETL to form heterojunction on both sides of the CuSbS2 absorber layer. In general, for the BSF layer in solar cells, direct band gap thin films materials with high carrier density and mobility, and low electron affinity are desired requirements. Thermally stable ternary delafossite copper aluminate (CuAlO2) is a p-type semiconductor. It has relatively low electron affinity compared to CuSbS2-absorber layer, tunable band gap in the range of 2.5 to 3.5 eV [15,43], low resistivity [43], and tunable acceptor concentration, making it suitable as the BSF layer for the CuSbS2-based solar cell. Carrier concentration in CuAlO2 thin films can be increased by incorporating excess oxygen and divalent species into the delafossite structure. Ruijian Liu et al. [44] reported that carrier concentration of CuAlO2 thin films can be increased to the order of 1018 cm−3 with 6.0% of Mg doping. As far as we know, for the first time, we design and simulate CuSbS2-based heterojunction solar cells with the novel device structure: Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au(100) using SCAPS-1D. The impact of physical parameters of different layers on the PV performances of the newly designed device has been investigated.

2. Methodology, material parameters, and device structure

Computer-aided design and simulation are becoming more and more compelling in the wafer-based silicon PV industry. To design and evaluate silicon PV devices, simulation tools such as PC1D, AFORS-HET, ATLAS, and Sentaurus TCAD are commonly used [4547]. Due to the complex structure and quantum-mechanical behavior of thin-film devices, the development of thin-film PV simulation tools is quite slow as compared with wafer-based silicon simulation tools. To date, most of the TFSC simulation packages are based on 1D, which are freely available to the PV community. Some of the popular thin-film PV simulation packages are AMPS-1D, wxAMPS, SCPAS-1D, etc. [4850]. According to Burgelman [50], an ideal thin-film solar cell simulator should meet some requirements such as allowing multiple layers (minimum six layers), capable of handling graded materials, capable of describing the possible discontinuities in the energy band at the interface between layers etc. Based on the requirements specified in the Refs. [50], we searched popular simulation software packages to select the most suitable package for CuSbS2-based solar cell simulation. SCAPS-1D was selected because it meets all the requirements listed in the Refs. [50] for 1D devices. It also provides the most extensive set of simulation outputs, including energy band diagrams, J-V, QE, C-V, and C-f. SCAPS-1D simulation package adopts Poisson-drift-diffusion (PDD) equation sets to compute the electrical properties of PV devices. PDD model involves three coupled equations, namely, Poisson, carrier transport, and continuity equations of free holes and electrons. Under one-dimensional and steady-state conditions, SCAPS-1D solves these three coupled differential equations with appropriate boundary conditions to compute numerically the electrical properties of PV devices.

The schematic of (Al/ITO/n-TiO2/p-CuSbS2/p+-Mg:Cu AlO2/Au (100)) n-p-p+ heterojunction device considered in this simulation is shown in Fig. 1(a). In the solar cell layout, n-type TiO2 was selected as the heterojunction partner with the active p-type CuSbS2 absorber layer forming the n-p heterojunction interface that collects the photogenerated carriers (electrons) efficiently. A window layer on the top of TiO2 layer is used consisting of highly transparent conductive and low-cost ITO. The p+-CuAlO2 layer is used as the BSF layer, which improves the photo-generated carrier separation and collection. In this simulation, the shunt and series resistance have been considered as 103 Ω.cm2 and 2.5 Ω.cm2, respectively. Simulation of the considered device was performed under global AM 1.5 spectrum at 1 sun illumination (100 mW.cm−2) from the window layer (ITO) side at 300 K temperature. The physical properties of different layers and baseline parameters, as it is important and ultimately determine the relative accuracy of simulated results, used for computing the PV response of CuSbS2-based solar cells are selected from experimental data, various reasonable estimates, or literature which are given in Table 1. The optical absorption spectra for the different layers, namely, TiO2, CuSbS2, and CuAlO2:Mg of the device under consideration are taken from the previously reported experimental values [5861]. As the bulk recombination loss is concerned, a single-acceptor/donor-like bulk defects with Gaussian energetic distribution is introduced into the bulk region of each semiconductor layer. Losses due to minority carrier recombination at each interface are considered by incorporating reasonable neutral interfacial defects, shown in Table-1. The values of electrical parameters for the front and back-contact are listed in Table 2.

 figure: Fig. 1.

Fig. 1. (a) The schematic structure of the proposed (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) n-p-p+-heterojunction solar cell, and (b) J-V characteristics under illumination for CuSbS2-based single and heterojunction solar cells with different device configurations.

Download Full Size | PDF

Tables Icon

Table 1. Baseline material parameters, used for simulating CuSbS2-based heterojunction ultrathin film solar cells [5154].a

Tables Icon

Table 2. Electrical properties of front and back contact materials used in simulation

3. Results and discussion

3.1 Enhancement of the PV performance of the CuSbS2-based solar cells

The basic model of the CuSbS2-based solar cell (Cell-1) reported so far has the traditional Mo/CuSbS2/CdS/n-ZnO/Al device structure, in which CdS is used as the ETL layer and Mo is used as the back-contact material, which has been developed, simulated and obtained PCE of 12.21%. This is the close approximation to the reported result in Ref. [62]. The J-V relationship of the CuSbS2/CdS device (Cell-1) is shown in Fig. 1(b). As mentioned earlier, the low PV performance of the Cell-1 is due to the dominant interface recombination, high series resistance, and unfavourable heterojunction configurations. This CuSbS2 based solar cell uses CdS as ETL, which is toxic and more convenient to form a “cliff-like” CBO at the CuSbS2/CdS interface, thereby acting as the barrier to the photo-induced electrons transport and hindering the PV performance of the device. The J-V characteristics of Cell-2 is illustrated in Fig. 1(b). The CdS layer of the Cell-1 is replaced by another suitable TiO2 ETL in Cell-2. The solar cell having TiO as ETL (Cell-2) exhibits slightly enhanced PV performance: Voc ≈ 0.708 V, Jsc ≈ 31.59 mA/cm2, FF ≈ 57.88%, and PCE ≈ 12.93%. This improvement in the PV performance of Cell-2 relative to Cell-1 is associated with the replacement of the CdS layer with the TiO2 layer.

It is worth mentioning that the application of low-work function metal back contact on photovoltaic devices will result in the formation of Schottky barriers that impede photo-generated carriers (holes), thereby reducing Voc and pernicious to the device’s performance. In fact, a feature called “rollover” is observed in the J-V curves of Cell-1 and Cell-2 (see Fig. 1(b)), indicating that a barrier for photogenerated carriers is developed at the back contact. Moreover, in case of Mo back-contact, the formation of a thin MoS2 layer is highly probable during CuSbS2 film growth on the Mo substrate because of the high reactivity of S with the Mo. A thin MoS2 layer between Mo back-contact and absorber layer can also adversely affect the cell performance. Therefore, this result suggests that Mo back-contact is not suitable for the CuSbS2-based solar cell, and it needs to be replaced by another suitable back-contact metal with a higher work function than Mo. Metal with higher work function may weaken the barrier of the photogenerated carriers at the back-contact.

Aiming at this point, another solar cell (Cell-3) was modeled and simulated, in which the back-contact metal (Mo) of Cell-2 was replaced by nickel (Ni) (its work function is higher than Mo), and Fig. 1(b) demonstrates the J-V characteristic curve for the Cell-3. The Cell-3 provides PCE of 18.89%, which is still below the Shockley-Queisser (SQ) limit for a single-junction solar cell with the absorber layer bandgap of ∼ 1.5 eV [63]. This is due to the short photogenerated carrier lifetime in CuSbS2 absorber as compared to CIGS and CdTe absorber. However, when the CuAlO2:Mg is added as BSF layer to the heterojunction of Cell-3, the efficiency of the device with double-heterojunction increases to ≈ 22.23%. The final cell with CuAlO2:Mg as BSF layer is identified as Cell-4 and the J-V characteristic curve of the Cell-4 is also presented in Fig. 1(b), from which the main PV parameters are derived. The PV performance parameters which are derived from the J-V curves of all cells are presented in Table 3. Therefore, data in Table 3 and the simulated J-V characteristics curves suggest a superior PV response for Cell-4 with respect to other cells studied herein.

Tables Icon

Table 3. PV performance parameters for CuSbS2-based solar cells with different cell configurations

3.2 Energy band diagram of the CuSbS2-based heterojunction solar cell

The energy band matching at the interface between two semiconductors is very important for the proper functioning of their bulk properties. The energy band diagram under illumination of (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) heterojunction solar cell (Cell-4) is shown in Fig. 2. It has been specified in the introduction section that the low Voc in conventional CuSbS2-based solar cell (Cell-1) is due to the poor energy band alignment with the CdS ETL. The too higher values of the spike and cliff CBO at the heterojunction interface (ETL/absorber interface) will reduce Voc and Jsc, respectively. Therefore, for high-efficiency solar cells, the energy mismatch at the heterojunction interface should be minimal. It is observed from Fig. 2 that a small energy mismatch (spike-like CBO) between the conduction band of CuSbS2-absorber and TiO2 ETL is developed, according to literature which is beneficial, ensuring an easy transport of photoelectrons at the TiO2/CuSbS2 interface, whereas blocking photogenerated holes at the TiO2/CuSbS2 interface, as the valance band edge of TiO2 ETL layer is much lower in energy than the CuSbS2 absorber. On the other hand, the holes from the CuSbS2 absorber layer are easily extracted by the CuAlO2:Mg BSF layer and transported to the back electrode, since the VBO between CuSbS2-absorber and CuAlO2:Mg BSF layer is very small. In contrast, the large CBO between the CuSbS2 absorber and CuAlO2:Mg BSF layers will prevent photoelectrons from moving to the CuAlO2:Mg BSF layer.

 figure: Fig. 2.

Fig. 2. Schematic energy band diagram of the proposed (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) heterojunction solar cell.

Download Full Size | PDF

3.3 Built-in-potential of the CuSbS2-based solar cell

The built-in potential at junction yields a significant electric field to split the photogenerated electron-hole pairs in the space charge region (SCR). Therefore, the carrier recombination in the SCR area decreases as the built-in potential increases. There are two heterojunctions in our proposed device structure, namely, n-p heterojunction formed by the p-CuSbS2 absorber and n-TiO2 ETL layer and p+-p heterojunction formed by the p+-CuAlO2:Mg and p-CuSbS2 absorber layer with the carrier densities: NA in p-CuSbS2, ND in n-TiO2, and Na in p+-CuAlO2:Mg layers, respectively. The junction potential, which is defined as the splitting of the quasi-Fermi levels in both n-type and p-type semiconductor, of the n-p heterojunction and p+-p heterojunction can be expressed as the following expressions [65]:

$${\Psi _{n - p}} = \frac{{\varDelta {E_C} - \varDelta {E_V}}}{{2q}} + \frac{{kT}}{q}ln\frac{{{N_D}{N_A}}}{{{n_{{i_n}}}{n_{{i_p}}}}} + \frac{{kT}}{{2q}}ln\frac{{{N_{{v_n}}}{N_{{c_p}}}}}{{{N_{{v_p}}}{N_{{c_n}}}}}$$
$${\Psi _{{p^ + } - p}} = \frac{{\varDelta {E_C} - \varDelta {E_V}}}{{2q}} + \frac{{kT}}{q}ln\frac{{{N_a}{N_A}}}{{{n_{{i_p}}}{n_{{i_{{p^ + }}}}}}} + \frac{{kT}}{{2q}}ln\frac{{{N_{{v_p}}}{N_{{c_{{p^ + }}}}}}}{{{N_{{v_{{p^ + }}}}}{N_{{c_p}}}}}$$
where, NC and NV are the effective density of states at conduction band and valance band respectively, T is the temperature, k is the Boltzmann constant, q is the electron’s charge and ${n_{{i_n}}},$ ${n_{{i_p}}},$ and ${n_{{i_{{p^ + }}}}}$ are the intrinsic (undoped) carrier density of the corresponding n-TiO2 (8.9 × 1017 cm-3 [66]), p-CuSbS2 (1016 cm-3 [56]) and p+-CuAlO2:Mg (1.78× 1018 cm-3 [44]) layers, respectively. Therefore, the cell potential (Ψ) for the (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) heterojunction solar cell can be written as Eq. (3):
$$\Psi \; = \; {\Psi _{n - p}}\; + {\Psi _{{p^ + } - p}}$$

Consequently, the calculated potentials of the n-p junction, p+-p interface, and n-p-p+ heterojunctions are ≈ 0.95 V, ≈ 0.83 V, and ≈ 1.78 V, respectively. This high potential is due to the double-heterojunction structure of the solar cell we designed. When the built-in potential is high, carrier recombination and trapping are lower under irradiation than when the built-in potential is low. The reduced photo-generated carrier trapping at the defect sites in the interface region further increases the effective bulk field under illumination, thereby reinforced the performance of the device as it is observed in Cell-4.

3.4 Recombination analysis of the CuSbS2-based solar cells

Among various recombination processes in solar cells, Shockley-Read-Hall (SRH) recombination is known to be a dominant recombination process which deteriorates the cell performance mostly. Qualitative and quantitative information about the rate of carrier recombination at different regions of the device such as at the ETL/absorber interface (Ri,f), at BSF layer/absorber interface (Ri,b), in the SCR zone (Rd), and in the bulk region (Rb) can be obtained using SRH recombination statistics and analytical approximation [67]. The recombination rates depend on various factors such as thickness, carrier density, carrier lifetime, bias voltage, working temperature of the devices, etc. However, the temperature (T) and the bias voltage (V)-dependent recombination rates can be expressed as follows [67]:

$${R^{i,f}} = R_o^{i,f}{e^{\frac{{qV}}{{{k_B}T}}}}$$
$${R^d} = R_o^d{e^{\frac{{qV}}{{2{k_B}T}}}}$$
$${R^b} = R_o^b{e^{\frac{{qV}}{{{k_B}T}}}}$$
$${R^{i,b}} = R_o^{i,b}{e^{\frac{{qV}}{{{k_B}T}}}}$$
where, Roi,f, Rod, Rob, and Roi,b are recombination coefficients at the ETL/absorber interface, in the SCR zone, in the bulk region, and at the absorber/BSF interface with zero bias voltage (V = 0), respectively. Assume that total generation across the absorber GaW (W = width of the absorber layer) is equal to the total recombination, the Voc can be expressed as follows:
$${V_{oc}} = \frac{{2kT}}{q}\textrm{ln}[{K_1}\left( {\sqrt {{G_a}{K_2} + 1} - 1} \right)$$
where,
$${K_1} = \frac{1}{2}\frac{{R_0^d}}{{R_0^{i,f} + R_0^b + R_0^{i,b}}}$$
and
$${K_2} = \frac{{4W({R_0^{i,f} + R_0^b + R_0^{i,b}} )}}{{{{({R_0^d} )}^2}}}$$

Using Eq. (9) and Eq. (10), the recombination coefficient in the depletion region is

$$R_0^d = \frac{{2W}}{{{K_1}{K_2}}}$$
is clearly seen from the above expression that all recombination coefficients can be obtained by extracting the value of K1 and K2 from the Voc analysis based on temperature (T) and excitation light intensity (Ga) variation. In particular, using Eq. (11), the recombination coefficients at the SCR zone ($R_0^d)$, can be obtained by extracting values of K1and K2 from the intensity-dependent Voc analysis under white light irradiation condition. However, the influence of introducing CuAlO2:Mg as BSF on the back-contact recombination can be revealed through wavelength-dependent Voc analysis [67].

Now focusing on the extraction of the recombination coefficients at absorber/BSF interface ($R_0^{i,b}$). For the analysis of Voc with monochromatic short-wavelength (SW) excitation, since the absorber band gap is much lower than the energy of the SW photon, electron and hole pairs are generated near the surface region. Therefore, the SRH recombination is localized in the vicinity of the front interface (TiO2/CuSbS2 interface). As a result, bulk SRH recombination has the weak impact, and rear (back) interface SRH recombination has negligible (i.e. $R_0^{i,b}$ 0) impact on total SRH recombination. Therefore, under monochromatic short-wavelength (SW) illumination, the coefficient K1and K2 can be expressed as:

$$K_1^{SW} = \frac{1}{2}\frac{{R_o^d}}{{({R_o^{i,f} + R_o^b} )}}$$
and
$$K_2^{SW} = \frac{{4W({R_o^{i,f} + R_o^b} )}}{{{{({R_o^d} )}^2}}}$$

On the other hand, with long-wavelength (LW) excitation (photon energy is nearly equal to the band gap of the absorber), one can expect a nearly uniform generation profile across the absorber layer and therefore, the coefficient K1and K2 can be expressed:

$$K_1^{LW} = \frac{1}{2}\frac{{R_o^d}}{{({R_o^{i,f} + \; R_o^b\; \; + \; R_o^{i,b}} )}}$$
and
$$K_2^{LW} = \frac{{4W({R_o^{i,f} + \; R_o^b\; + \; R_o^{i,b}} )}}{{{{({R_o^d} )}^2}}}$$

Using the above two eq.(12) and eq.(14) and after simplification, the recombination coefficient at the back-contact interface can be expressed as:

$$\; R_o^{i,b} = \frac{1}{2}\left[ {\frac{1}{{K_1^{LW}}} - \frac{1}{{K_1^{SW}}}} \right] \times R_o^d$$

Therefore, using the value of $R_0^d$ obtained previously from Eq. (11) and simplified Eq. (16), we can get the recombination coefficients at the back-contact interface ($R_o^{i,b}$). Finally, the recombination rate in the SCR zone and at the back-contact interface can be obtained using Eq. (5) and Eq. (7), and from known values of $R_0^d$ and $R_o^{i,b}$. In this study, aiming to extract qualitative and quantitative information about the recombination at the back-interface, two different monochromatic wavelengths (450 nm and 800 nm) are used to illuminate the designed solar cell. The intensity (Ga) dependent Voc curves at excitation wavelength λexc = 450 nm, λexc= 800 nm, and white light (WL) for Cell-1 and Cell-4 are shown in Fig. 3.

 figure: Fig. 3.

Fig. 3. Intensity-dependent Voc with different excitation wavelength of white light (WL), λexc= 450 nm, and λexc= 800 nm at constant temperature (300 K) for (a) Cell-1 and, (b) Cell-4.

Download Full Size | PDF

We have done the same Voc analysis for all devices but intensity-dependent Voc curves for Cell-2 and Cell-3 are not shown here. The value of the coefficients K1and K2 are obtained from the fitting of the intensity-dependent Voc curves of Fig. 3 by using Eq. (9) and Eq. (10), and other recombination coefficients extracted from K1and K2 are provided in Table 4. From the data in Table 4, it can be evidenced that the poor PV performance of the Cell-1 is due to high carrier recombination at the SCR and the back-contact region. Replacing the CdS ETL layer of Cell-1 with TiO2 ETL layer in Cell-2 slightly reduces carrier recombination rate at the SCR (Rd) and the back-contact region (Ri,b) from 2.55×1016 cm-2s-1 and 9.26×108 cm-2s-1 to 2.43×1016 cm-2s-1 and 8.69×108 cm-2s-1, respectively, results in a small enhancement of PV performance of Cell-2. However, a significant improvement of PV performance of the Cell-3 is observed, as the carrier recombination at the SCR and back-contact region is reduced (see Table 4) by diminishing barrier height for photogenerated holes at back-contact (replacing Mo with Ni). The effect of incorporating the CuAlO2:Mg BSF layer into CuSbS2-based solar cells (Cell-4) is obvious because the carrier recombination rate at the SCR and back-contact region is significantly reduced to 2.51×1014 cm-2s-1 and 2.37×104 cm-2s-1, respectively, thereby further enhancement of PV performance of the device. The presence of the electric field at the CuSbS2/CuAlO2:Mg junction would contribute to increase the hole transportation from CuSbS2 absorber to CuAlO2:Mg layer, and can minimize the flow of minority carriers (electrons) toward the the CuSbS2/CuAlO2:Mg interface and thereby reducing the carrier recombination rate at the CuSbS2/CuAlO2:Mg interface [68]. The electric field at that junction can also improve the electric potential along the absorber, which in turn improve the separation of photo-induced carriers and enhances the collection efficiency at the respective electrodes.

Tables Icon

Table 4. Extracted values of K1 and K2, and recombination coefficients for CuSbS2-based solar cells of different configurations. ${\boldsymbol R}_0^{\boldsymbol d}$and${\boldsymbol \; R}_{\boldsymbol o}^{{\boldsymbol i},{\boldsymbol b}}$are calculated at bias voltage of 0.9 V.

3.5 Performance optimization of the CuSbS2 solar cell

3.5.1 Impact of CuSbS2 layer thickness and carrier concentration

In this work, the PV performance of the final device with the structure of (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) is optimized by tuning the thickness and carrier density of the CuSbS2, ETL, and BSF layers. In this section, the PV response of the device is evaluated by varying carrier concentration and thickness of the CuSbS2 layer from 1015 to 1018 cm−3, and 200 nm to 900 nm, respectively, keeping all the other materials parameters for different layers at the baseline values presented in Table 1 and Table 2. The computational results, i.e., the change in PV parameters of the device in response to the variation of the absorber layer’s thickness and concentration, are shown in Fig. 4. It is observed from Fig. 4(a) that except FF, all PV parameter values increase steeply up to 400 nm, stabilized between 400 nm and 500 nm, and then fall off slowly with the CuSbS2 layer thickness. In particular, the maximal values of Voc, Jsc, and PCE were obtained (995.4 mV, 32.266 mA/cm2, and 22.55%, respectively) with a thickness of only 425 nm, which is mainly due to the strong absorption light in the visible region by the CuSbS2 layer. However, due to enhanced carrier recombination in the thicker absorber layer, Voc decreases slowly after 500 nm of the absorber layer thickness. According to Beer-Lambert law, the initial increase in Jscwith absorber thickness is due to more photons being absorbed within the absorber layer of the device. In addition, the initial increase in PCE is due to the fact that when the diffusion length of photogenerated carriers is greater than the absorber layer thickness, most of the carriers arrive at the electrodes before recombination, and therefore, PCE increases with the thickness of the absorber layer. However, for the thick absorber layer, bulk recombination occurs, and hence, PCE declines with the further increase in thickness. FF of the device showing a different trend of decreasing value with the absorber thickness, indicating that the series resistance increases with the increase in thickness. The optimum thickness for the absorber layer can also be estimated in the future by carried out simulation based on first principles (ab initio) study.

 figure: Fig. 4.

Fig. 4. PV responses of CuSbS2-based solar cells as a function of absorber layer (a) thickness and (b) acceptor concentration.

Download Full Size | PDF

In addition to the absorber layer thickness, the PV performance of the device is significantly affected by the acceptor concentration (NA) in the absorber layer. The p-type CuSbS2 system has a high density of acceptor defects with low formation energy, which mostly occurs during the synthesis of this materials. To understand the effect of NA on the PV responses of the device, NA in the CuSbS2-absorber layer is varied from 1015 to 1018 cm−3, with optimized absorber thicknesses (425 nm). The other materials parameters of different layers are kept unchanged. Figure 4(b) provides the changes of PV responses of the devices relative to acceptor density variation of the absorber layer. We observe that Voc initially increases linearly with NA up to 5×1017 cm−3, and then falls off sharply when NA is further increased. As NA increases, the quasi-Fermi energy level of holes moves downward and hence Voc increased. Another aspect is that increased built-in potential, which is due to an increase of NA, enhances photogenerated carrier separation in the SCR of the device, thereby increasing the Voc. Any further increase in NA resulted in sharply reduced Voc, which is due to enhanced carrier recombination at the absorber layer. The Jsc of the device also showed a reduction with an increase in NA beyond a certain concentration (2×1016 cm−3). The enhanced carrier recombination after critical concentration diminishes the minority carrier lifetime and resulting in the reduced Jsc. This is due to the fact that the Coulomb interactions of the photogenerated electrons with the overpopulated acceptor doping, may lead to more hole’s traps and recombination in the absorber layer [69]. The cumulative effect of Voc and Jsc causes the PCE of the device to increase with increasing NA up to 3.16×1016 cm−3. The overall PCE of the device decreases after the NA concentration of 3.16×1016 cm−3. Therefore, it becomes important to control the acceptor concentration in the absorber layer to operate the device in the maximum PCE regime. The maximum PCE of about 22.78% for CuSbS2-based solar cell is achieved at the absorber thickness of only 425 nm and acceptor concentration of 3.16×1016 cm−3 (Voc ≈ 966 mV, Jsc ≈ 34.20 mA/cm2, and FF ≈ 68.98%). Consequently, this investigation proposes that CuSbS2-absorber might be appropriate for ultrathin solar PV devices.

3.5.2 Impact of TiO2 ETL and CuAlO2:Mg BSF layer thickness and carrier concentration

In this section, to comprehend the effect of ETL and BSF layer carrier density on PV response of the device, the carrier density of ETL and BSF layer is varied in the range of 1017 to 1020 cm−3 and 1017 to 1019 cm−3, respectively. The simulation is performed with the optimized absorber layer thickness and carrier density obtained previously, whereas, other material parameters are kept at the baseline values. Figure 5 illustrates the change in PV responses of the device with respect to the ETL and BSF layer carrier density variation. It is worth mentioning that we additionally simulated the impact of ETL and BSF layer thickness, results are not shown here, and found that the device performance hardly changes with the thickness of ETL and BSF layer. In particular, all the PV parameters of the device slightly increase with the thickness of the BSF layer. On the contrary, the increase in ETL layer thickness slowly reduces the device performance. So, the thickness of 75 nm and 200 nm are chosen for ETL and BSF layer, respectively, to carry on further study. It is seen from Fig. 5(a) that as the ETL layer carrier concentration increases to 5×1017 cm−3, all PV parameters of the device increase sharply at first, then slowly increase and stabilize at a concentration level exceeding 1019 cm−3. However, the PV responses of the device enhance gradually as the BSF layer carrier concentration increases and finally saturates at the concentration level beyond 1019 cm−3. In this simulation, the maximum PV responses of Voc: 969 mV, Jsc: 34.62 mA/cm2, FF: 68.71%, and PCE: 23.05% of the proposed device are achieved at the ETL and BSF layer carrier densities of 5×1018 cm−3 and 3.75×1018 cm−3, respectively.

 figure: Fig. 5.

Fig. 5. PV responses of CuSbS2-based solar cells as a function of (a) TiO2 ETL, and (b) CuAlO2:Mg BSF layer carrier concentration.

Download Full Size | PDF

3.6 Effect of defects at TiO2/CusbS2 and CuAlO2:Mg/CuSbS2 interface on the device performance

The performance of PV devices is also strongly affected by interfacial defects in the heterojunction structure. In this study, the effect of the defect density at TiO2/CusbS2 and CuAlO2:Mg/CuSbS2 interface on the performance of the proposed CuSbS2 solar cell is investigated. Figure 6 indicates simulation results for the solar cell PV parameters as a function defect density at the TiO2/CuSbS2 and CuAlO2:Mg/CuSbS2 interface in the range from 1012 to 1020 cm-2. It can be seen that all the performance parameters of the device are declined with increasing the defect concentration at both interfaces. In particular, it is evident from the simulated results, all the PV performance parameters of the device are almost stable when the defect density reaches 1017 cm-2 and then gradually decreases as it increases further. This performance drop is due to the increase in the carrier recombination at the interface with increasing the defect density [70,71]. With the increase of interfacial defects, the series resistance of the device also increases significantly [70]. It is worth noting that for the increase of defect density at the TiO2/CuSbS2 interface from 1012 to 1020 cm-2, the conversion efficiency drops from 23.05% to 21.9%, while for the same increase in defect density at the CuAlO2:Mg/CuSbS2 interface, the conversion efficiency drops from 23.05% to 22.46%. Therefore, it is revealed that the performance of the proposed CuSbS2 solar cell with heterojunction structure is more affected by the defect at TiO2/CuSbS2 interface than at the CuAlO2:Mg/CuSbS2 interface [72].

 figure: Fig. 6.

Fig. 6. PV responses of CuSbS2-based solar cells as a function of defect density (a) at CuAlO2:Mg/CuSbS2, and (b) TiO2/CuSbS2 interfaces.

Download Full Size | PDF

3.7 Effect of back-contact metal work function and temperature on the PV response of the optimized CuSbS2 device

The back-contact potential barrier mainly hampers hole transport at back-contact and it deviates the shape of the illuminated JV characteristics curve near to the Voc of thin-film solar cells from ideality. It actually limits the current density of the device under illumination and is known as rollover effect. The presence of rollover effect in the J-V characteristics of the conventional CuSbS2-based solar cell (in Cell-1, Mo is used as back-contact) and Cell-2 (see Fig. 1(b)) imply that a Schottky barrier is developed at the back contact, thus reducing the Voc of the device. Therefore, we can infer that Mo back contact commonly used in CuSbS2-based solar cells is not suitable for high-efficiency CuSbS2-based PV devices and a stable, and low resistance back-contact that is not substantially rectifying is required for high-performance CuSbS2-based solar cells. To study the effect of changing the work function of the back-contact metal i.e. the ohmic or rectifying behavior at the back-contact/BSF layer interface, the simulation is carried out by varying the work function of the anode material from 4.9 eV to 5.7 eV. Throughout the simulation, the thickness and carrier concentration of different layer has been fixed at the optimized value obtained previously. In Fig. 7(a), the normalized values of PV parameters in response to the variation of metal work function at the back-contact are depicted. It is observed that except Jsc, the values of all PV parameters of the device linearly increases with increasing the back- contact work function to 5.4 eV and then become steady on further increasing work function of anode materials. In contrast, the Jscis relatively less sensitive and slightly increases as the barrier height decreases on increasing back-contact metal work function. For anode materials with a work function of less than 5.4 eV, a rectifying Schottky barrier contact is developed at the anode/CuAlO2:Mg interface, resulting in poor hole transport to the anode, reducing FF and PCE, as shown in Fig. 7(a). It is known that the energy barrier (ΨB) at the anode/CuAlO2:Mg interface is given by

$${\Psi _B} = \frac{{{E_g}}}{q} + \chi - {\varphi _m}$$

Here, Eg is the band gap of CuAlO2:Mg (3.46 eV), χ is the electron affinity of CuAlO2:Mg (2.50 eV) and φm is the anode’s work function. For the anode materials with work function higher than 5.4 eV, according to Eq. (17), the energy barrier is less than 0.5 eV, which may not affect the Voc of the device [73], thereby, achieving the maximum performance of the device. This finding indicates that anode materials with relatively high work functions (> 5.4 eV) are required to achieve the maximum efficiency of CuSbS2-based PV devices.

 figure: Fig. 7.

Fig. 7. PV responses of CuSbS2-based solar cells as a function of (a) work function of the back-contact materials, and (b) working temperature.

Download Full Size | PDF

Material properties and PV performance of solar cells are greatly affected by the operating temperature. Generally, solar cells are installed outdoors and exposed to a temperature range of 288 K to 323 K [74], and even higher temperatures in some other applications. The reverse saturation current density (J0) is known to be the key parameter, which is mostly affected by the temperature and band gap of the materials, affecting the PCE of the PV devices [75]. The effect of temperature on the CuSbS2-based solar cell performances is presented in Fig. 7(b). In Fig. 7(b), normalized PV responses of the device are plotted in line with the change in the operating temperature. As expected, the nearly linear degradation in the Voc and FF accompanied by a linear degradation in the PCE is observed on increasing the temperature. The fall in Voc allows electrons to gain enough energy at higher temperatures and recombine with holes before entering into the depletion region and collected at electrode [76]. A slight increase in Jsc is observed, which is due to the decrease of the band gap of semiconductor with increasing temperature. The temperature coefficient CT ((%K−1) of PCE, is an indicator of PV devices, showing how higher operating temperature affects the device’s PEC, and can be expressed under standard test conditions (STC: 298 K) as below [8]:

$${C_T} = \frac{1}{{{\eta _{STC}}}}\frac{{d{\eta _T}}}{{dT}} \times 100\%$$
where, ηSTC is the cell efficiency at STC (298 K) and ηT is the efficiency at any temperature (T). Fitting the PCE curve in Fig. 7(b) using Eq. (18), we found that the temperature coefficient of PCE of the cell is - 0.323%K−1, which is less than conventional Si (- 0.5%K−1), and CIGS (- 0.443%K−1 [77]) solar cells.

3.8 Effect of series resistance on the PV response of the optimized device

Both series (Rs) and shunt (Rsh) resistance have a huge impact on the PV performance of heterojunction solar cells, because the appearance and slope of the J-V characteristics curve are jointly controlled by them. Rs is originated from the electrical resistance related to the contact and bulk resistance of the semiconductors (ETL, BSF, and absorber layer), while the shunt resistance mainly comes from several alternative carrier recombination paths i.e. from defects in bulk semiconductors and interfaces, and also depend on the design of the device (cell edges effect) [78]. Consequently, a larger Rs value will reduce the short-circuit current (Jsc), and a smaller Rsh value will cause the (Voc) to decrease, and as a combined effect, the fill factor drops sharply, resulting in a low PCE. In this section, the role of the Rs and Rsh on the optimized device has been investigated. In this simulation, we have varied Rs from 1 to 10 Ω·cm2, while the Rsh was kept constant at 103 Ω·cm2 and varied shunt resistance from 102 to 105 Ω·cm2, while keeping Rs constant at 2.5 Ω·cm2. The corresponding variations of the PV parameters are shown in Fig. 8. Figure 8(a) shows that as Rs increases, Voc initially increases slightly and then tends to saturate, while in the case of other PV parameters, it continues to decrease as observed. These findings are completely consistent with the reported literature [79].

 figure: Fig. 8.

Fig. 8. PV responses of CuSbS2-based solar cells as a function of (a) series resistance, and (b) shunt resistance.

Download Full Size | PDF

It has been observed that with every 0.01 Ω increase in Rs, FF degrades by nearly 2.6%, which is consistent with traditional Si (approximately 2.5% for every 0.01 Ω increase in Rs) solar cells. However, it is interesting that although the FF degrades to 2.6%, the efficiency degrades is much lower—only 0.92%, with Rs increases by 0.01 Ω. On the contrary, reducing Rsh below 1000 Ω·cm2 will cause all PV parameters to decrease rapidly (Fig. 8(b)). However, PCE and other PV parameters are less affected by Rsh > 1000 Ω·cm2.

4. Conclusions

This article unveiled a CuSbS2-based heterojunction solar cell with non-toxic buffer and BSF layers and analyzed it by SCAPS 1D simulation software. A (n-p-p+) heterojunction solar cell structure (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) has been designed and analyzed numerically. In this study, poor PV performance of CuSbS2/CdS heterojunction solar cells has been enhanced significantly in the presence of CuAlO2:Mg BSF layer and Au back-contact. This improvement of PV performance of the considered device is attributed to the reduction in back-contact recombination and enhancement of built-in potential. The CuSbS2-device structure has been optimized by tuning the thickness and carrier density of different layers to achieve the maximum PCE of the solar cell. The optimized device shows a maximum PCE of 23.05% when the CuSbS2 absorber layer thickness is only 425nm, indicating that it is suitable for ultra-thin solar PV devices. The study also suggests that the efficiency of novel CuSbS2 solar cells can increase by replacing Au with Mo as back contact material. The study again portrays that although the FF of the proposed device decreases by about 2.6% with an increase of 0.01Ω in Rs, its impact on PCE is really small (only 0.92%, with Rs increases by 0.01 Ω). Based on optimization, the highest efficiency of the solar cell reached to 23.05% (Voc = 969mV, Jsc = 34.61mA/cm2, FF = 68.71%), which is very encouraging compared with traditional CuSbS2/CdS heterojunction solar cells. Nevertheless, this work successfully demonstrates the low-cost, non-toxic, and earth-abundant CuSbS2-based ultrathin solar cell as a potential candidate in the future PV industry.

Acknowledgements

The authors gratefully acknowledge Prof. Marc Burgelman, University of Gent, Belgium, for providing SCAPS 1-D simulation software.

Disclosures

The authors declare no conflicts of interest.

Data availability

The main parameters to carry out the same simulations using SCAPS-1D are all given, and the calculations may easily be reproduced. The readers can also obtain the original data that underlies the present paper directly from the author.

References

1. Colin A. Wolden, Juanita Kurtin, Jason B. Baxter, Ingrid Repins, Sean E. Shaheen, John T. Torvik, Angus A. Rockett, Vasilis M. Fthenakis, and Eray S. Aydil, “Photovoltaic manufacturing: Present status, future prospects, and research needs,” J. Vac. Sci. Technol., A 29(3), 030801 (2011). [CrossRef]  

2. Fraunhofer, Fraunhofer ISE, Photovoltaics Report. Tech. Rep., Freiburg: Fraunhofer ISE, Germany; 2020. https://www.ise.fraunhofer.de/content/dam/ise/de/documents/publications/studies/Photovoltaics-Report.pdf [access on June 2022].

3. C. Candelise, M. Winskel, and R. Gross, “Implications for CdTe and CIGS technologies production costs of indium and tellurium scarcity,” Prog. Photovolt: Res. Appl. 20(6), 816–831 (2012). [CrossRef]  

4. M Tao, “Roadblocks to Terawatt Solar Photovoltaics,” SpringerBriefs in Applied Sciences and Technology (Springer, 2014), pp. 61–79.

5. D Bauer, D Diamond, J Li, D Sandalow, P Telleen, and B. Wanner, U.S., “Department of Energy Critical Materials Strategy.,” Tech. Rep., Office of Scientific and Technical Information (OSTI), (2010).

6. G. Martin A, D. Ewan D, H. Ebinger Jochen, Y. Masahiro, K. Nikos, and H. Xiaojing, “Solar cell efficiency tables (version 59),” Progress in Photovoltaics: Research and Applications 30(1), 3–12 (2022). [CrossRef]  

7. A. Rahman M, “Enhancing the photovoltaic performance of Cd-free Cu2ZnSnS4 heterojunction solar cells using SnS HTL and TiO2 ETL,” Sol. Energy 215, 64–76 (2021). [CrossRef]  

8. A. Rahman M, “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. 3(2), 253 (2021). [CrossRef]  

9. D. Colombara, L. Peter, K. Rogers, J. Painter, and S. Roncallo, “Formation of CuSbS2 and CuSbSe2 thin films via chalcogenisation of Sb-Cu metal precursors,” Thin Solid Films 519(21), 7438–7443 (2011). [CrossRef]  

10. J. Zhou, G. Q. Bian, Q. Y. Zhu, Y. Zhang, C. Y. Li, and J. Dai, “Solvothermal crystal growth of CuSbQ2 (Q = S, Se) and the correlation between macroscopic morphology and microscopic structure,” J. Solid State Chem. 182(2), 259–264 (2009). [CrossRef]  

11. A. Rabhi, M. Kanzari, and B. Rezig, “Optical and structural properties of CuSbS2 thin films grown by thermal evaporation method,” Thin Solid Films 517(7), 2477–2480 (2009). [CrossRef]  

12. Y. Rodriguez-lazcano, M. Nair, and P. Nair, “CuSbS2 thin film formed through annealing chemically deposited Sb2S3-CuS thin films,” J. Cryst. Growth 223(3), 399–406 (2001). [CrossRef]  

13. S. H. Pawar, A. J. Pawar, and P. N. Bhosale, “Spray pyrolytic deposition of CuBiS2 thin films,” Bull. Mater. Sci. 8(3), 423–426 (1986). [CrossRef]  

14. P. Sonawane, P. Wani, L. Patil, and T. Seth, “Growth of CuBiS2 thin films by chemical bath deposition technique from an acidic bath,” Mater. Chem. Phys. 84(2-3), 221–227 (2004). [CrossRef]  

15. L. Yu and A. Zunger, “Identification of potential photovoltaic absorbers based on first-principles spectroscopic screening of materials,” Phys. Rev. Lett. 108(6), 068701 (2012). [CrossRef]  

16. J. T. R. Dufton, A. Walsh, P. M. Panchmatia, L. M. Peter, D. Colombara, and M. S. Islam, “Structural and electronic properties of CuSbS2 and CuBiS2: potential absorber materials for thin-film solar cells,” Phys. Chem. Chem. Phys. 14(20), 7229 (2012). [CrossRef]  

17. M. Kumar and C. Persson, “CuSbS2 and CuBiS2 as potential absorber materials for thin-film solar cells,” J. Renewable Sustainable Energy 5(3), 031616 (2013). [CrossRef]  

18. J. Heo, G. Laurita, S. Muir, M. A. Subramanian, and D. A. Keszler, “Enhanced thermoelectric performance of synthetic tetrahedrites,” Chem. Mater. 26(6), 2047–2051 (2014). [CrossRef]  

19. D. Chen, Y. Zhao, Y. Chen, T. Lu, Y. Wang, J. Zhou, and Z. Liang, “Thermoelectric enhancement of ternary copper chalcogenide nanocrystals by magnetic nickel doping,” Adv. Electron. Mater. 2(6), 1500473 (2016). [CrossRef]  

20. S. Dekhil, H. Dahman, S. Rabaoui, N. Yaacoub, and L. E. Mir, “Investigation on microstructural and optical properties of CuSbS2 nanoparticles synthesized by hydrothermal technique,” J. Mater. Sci.: Mater. Electron. 28(16), 11631–11635 (2017). [CrossRef]  

21. V. Vinayakumar, S. Shaji, D. Avellaneda, T. K. Das, G. A. Roy, C. J. A. A. Martinez, and B. Krishnan, “CuSbS2 thin films by rapid thermal processing of Sb2S3-Cu stack layers for photovoltaic application,” Sol. Energy Mater. Sol. Cells 164, 19–27 (2017). [CrossRef]  

22. L. Shi, C. Wu, J. Li, and J. Ding, “Selective synthesis and photoelectric properties of Cu3SbS4 and CuSbS2 nanocrystals,” J. Alloys Compd. 694, 132–135 (2017). [CrossRef]  

23. L. Wan, C. Ma, K. Hu, R. Zhou, X. Mao, S. Pan, L. HelenaWong, and J. Xu, “Two-stage co-evaporated CuSbS2 thin films for solar cells,” J. Alloys Compd. 680, 182–190 (2016). [CrossRef]  

24. J. A. R. Aquino, D. L. R. Vela, S. Shaji, D. A. Avellaneda, and B. Krishnan, “Spray pyrolysed thin films of copper antimony sulfide as photovoltaic absorber,” Phys. Status Solidi C 13(1), 24–29 (2016). [CrossRef]  

25. R. E. Ornelas-Acosta, D. Avellaneda, S. Shaji, G. A. Castillo, T. K. D. Roy, and B. Krishnan, “CuSbS2 thin films by heating Sb2S3/Cu layers for PV applications,” J. Mater. Sci.: Mater. Electron. 25(10), 4356–4362 (2014). [CrossRef]  

26. S. Banu, S. J. Ahn, S. K. Ahn, K. Yoon, and A. Cho, “Fabrication and characterization of cost-efficient CuSbS2 thin film solar cells using hybrid inks,” Solar Energy Materials and Solar Cells 151, 14–23 (2016). [CrossRef]  

27. W. Septina, S. Ikeda, Y. Iga, T. Harada, and M. Matsumura, “Thin film solar cell based on CuSbS2 absorber fabricated from an electrochemically deposited metal stack,” Thin Solid Films 550, 700–704 (2014). [CrossRef]  

28. S. C. Riha, A. A. Koegel, J. D. Emery, M. J. Pellin, and A. B. F. Martinson, “Low-temperature atomic layer deposition of CuSbS2 for thin-film photovoltaics,” ACS Appl. Mater. Interfaces 9(5), 4667–4673 (2017). [CrossRef]  

29. S. Dekhil, H. Dahman, F. Ghribi, H. Mortada, N. Yaacoub, and L. E. Mir, “Study of CuSbS2 thin films nanofibers prepared by spin coating technique using ultra-pure water as a solvent,” Mater. Res. Express 6(8), 086450 (2019). [CrossRef]  

30. B. Yang, L. Wang, J. Han, Y. Zhou, H. Song, S. Chen, J. Zhong, L. Lv, D. Niu, and J. Tang, “CuSbS2 as a promising earth-abundant photovoltaic absorber material: a combined theoretical and experimental study,” Chem. Mater. 26(10), 3135–3143 (2014). [CrossRef]  

31. V. J. Embden, J. O. Mendes, J. J. Jasieniak, A. S. R. Chesman, and E. D. Gaspera, “Solution-processed CuSbS2 thin films and superstrate solar cells with CdS/In2S3 buffer layers,” ACS Appl. Energy Mater. 3(8), 7885–7895 (2020). [CrossRef]  

32. Y. Rodriguez-Lazcano, M. T. S. Nair, and P. K. Nair, “Photovoltaic p-i-n Structure of Sb2S3 and CuSbS2 absorber films obtained via chemical bath deposition,” J. Electrochem. Soc. 152(8), G635 (2005). [CrossRef]  

33. Francisco Willian de Souza Lucas, Adam W. Welch, Lauryn L. Baranowski, Patricia C. Dippo, Hannes Hempel, Thomas Unold, Rainer Eichberger, Beatrix Blank, Uwe Rau, Lucia H. Mascaro, and Andriy Zakutayev, “Effects of thermochemical treatment on CuSbS2 photovoltaic absorber quality and solar cell reproducibility,” J. Phys. Chem. C 120(33), 18377–18385 (2016). [CrossRef]  

34. C. Macíasa, S. Lugob, Á. Beníteza, I. Lópeza, B. Kharissova, A. Vázqueza, and Y. Peña, “Thin film solar cell based on CuSbS2 absorber prepared by chemical bath deposition (CBD),” Mater. Res. Bull. 87, 161–166 (2017). [CrossRef]  

35. A. W. Welch, L. L. Baranowski, P. Zawadzki, C. DeHart, S. Johnston, S. Lany, C. A. Wolden, and A. Zakutayev, “Accelerated development of CuSbS2 thin film photovoltaic device prototypes,” Prog. Photovolt: Res. Appl. 24(7), 929–939 (2016). [CrossRef]  

36. Y. Zhang, J. Huang, C. Yan, K. Sun, X. Cui, F. Liu, Z. Liu, X. Zhang, X. Liu, J. A. Stride, M. A. Green, and X. Hao, “High open-circuit voltage CuSbS2 solar cells achieved through the formation of epitaxial growth of CdS/CuSbS2 hetero-interface by post-annealing treatment,” Progress in Photovoltaics: Research and Applications 27(1), 37–43 (2019). [CrossRef]  

37. DK Sadanand Dwivedi, “Modeling of photovoltaic solar cell based on CuSbS2 absorber for the enhancement of performance,” IEEE Trans. Electron Devices 68(3), 1121–1128 (2021). [CrossRef]  

38. T. J. Whittles, T. D. Veal, C. N. Savory, A. W. Welch, F. W. de Souza Lucas, J. T. Gibbon, M. Birkett, R. J. Potter, D. O. Scanlon, A. Zakutayev, and V. R. Dhanak, “Core levels, band alignments, and valence-band states in CuSbS2 for solar cell applications,” ACS Appl. Mater. Interfaces 9(48), 41916–41926 (2017). [CrossRef]  

39. T. Minemoto, T. Matsui, H. Takakura, Y. Hamakawa, T. Negami, Y. Hashimoto, T. Uenoyama, and M. Kitagawa, “Theoretical analysis of the effect of conduction band offset of window/CIS layers on performance of CIS solar cells using device simulation,” Sol. Energy Mater. Sol. Cells 67(1-4), 83–88 (2001). [CrossRef]  

40. H. Han, X. Zhao, and J. Liu, “Enhancement in photoelectric conversion properties of the dye-sensitized nanocrystalline solar cells based on the hybrid TiO2 electrode,” J. Electrochem. Soc. 152(1), A164–A166 (2005). [CrossRef]  

41. B. Tan and Y. Wu, “Dye-sensitized solar cells based on Anatase TiO2 nanoparticle/nanowire composites,” J. Phys. Chem. B 110(32), 15932–15938 (2006). [CrossRef]  

42. B. H. Lee, M. Y. Song, S. Y. Jang, S. M. Jo, S. Y. Kwak, and D. Y. Kim, “Charge transport characteristics of high efficiency dye-sensitized solar cells based on electrospun TiO2 nanorod photoelectrodes,” J. Phys. Chem. C 113(51), 21453–21457 (2009). [CrossRef]  

43. Y. Zhang, Z. Liu, D. Zang, and L. Feng, “Structural and opto-electrical properties of Cu-Al-O thin films prepared by magnetron sputtering method,” Vacuum 99, 160–165 (2014). [CrossRef]  

44. R. Liu, O. Yongfeng Li, B. Yao, Z. Ding, Y. Jiang, L. Meng, R. Deng, L. Zhang, Z. Zhang, H. Zhao, and L. Liu, “Shallow acceptor state in Mg-Doped CuAlO2 and its effect on electrical and optical properties: an experimental and first-principles study,” ACS Appl. Mater. Interfaces 9(14), 12608–12616 (2017). [CrossRef]  

45. D. A. Clugston and P. A. Basore, “PC1D Version 5- 32-bit solar cell modeling on personal computers,” 26th IEEE Photovoltaics Specialist Conference, Anaheim California, pp. 207–210, (1997).

46. M. Elbar and S. Tobbeche, “Numerical simulation of CGS/CIGS single and tandem thin-film solar cells using the Silvaco-Atlas software,” Energy Procedia 74, 1220–1227 (2015). [CrossRef]  

47. S. Geißendörfer, J. Lacombe, G. Letay, K. von Maydell, and C. Agert, “Simulation of Amorphous and Microcrystalline Thin Film Silicon Solar Cells with Sentaurus TCAD,” 5th World Conference on Photovoltaic Energy Conversion, Valencia, Spain, pp-3133–3137, 6-10 September (2010).

48. Y. Liu, Y. Sun, and A. Rockett, “A new simulation software of solar cells—wxAMPS,” Sol. Energy Mater. Sol. Cells 98, 124–128 (2012). [CrossRef]  

49. Hong Zhu, Ali Kaan Kalkan, Jingya Hou, and Stephen J. Fonash, “Applications of AMPS-1D for solar cell simulation,” in: Proceedings of the National Center for Photovoltaics (NCPV) 15th Program Review Meeting, Denver, Colorado, USA, pp. 309–314 (1999).

50. M. Burgelman, J. Verschraegen, S. Degrave, and P. Nollet, “Modeling thin-film PV devices,” Prog. Photovolt: Res. Appl. 12(23), 143–153 (2004). [CrossRef]  

51. P. K. Patel, “Device simulation of highly efficient eco-friendly CH3NH3SnI3 perovskite solar cell,” Sci. Rep. 11(1), 3082 (2021). [CrossRef]  

52. R. Teimouri and R. Mohammadpour, “Potential application of CuSbS2 as the hole transport material in perovskite solar cell: A simulation study,” Superlattices Microstruct. 118, 116–122 (2018). [CrossRef]  

53. M. Shasti and A. Mortezaali, “Numerical study of Cu2O, SrCu2O2, and CuAlO2 as hole-transport materials for application in perovskite solar cells,” Phys. Status Solidi A 216(18), 1900337 (2019). [CrossRef]  

54. A. Kuddus, M. F. Rahman, S. Ahmmed, J. Hossain, and A. B. M. Ismail, “Role of facile synthesized V2O5 as hole transport layer for CdS/CdTe heterojunction solar cell: Validation of simulation using experimental data,” Superlattices Microstruct. 132, 106168 (2019). [CrossRef]  

55. T. Shawky, M. H. Aly, and M. Fedawy, “Performance analysis and simulation of c-Si/SiGe based solar cell,” IEEE Access 9, 75283–75292 (2021). [CrossRef]  

56. Francisco Willian de Souza Lucas, Haowei Peng, Steve Johnston, Patricia C. Dippo, Stephan Lany, Lucia H. Mascaro, and Andriy Zakutayev, “Characterization of defects in copper antimony disulfide,” J. Mater. Chem. A 5(41), 21986–21993 (2017). [CrossRef]  

57. H. B. Michaelson, “The work function of the elements and its periodicity,” J. Appl. Phys. 48(11), 4729–4733 (1977). [CrossRef]  

58. T. Rath, A. J. MacLachlan, M. D. Brown, and S. A. Haque, “Structural, optical and charge generation properties of chalcostibite and tetrahedrite copper antimony sulfide thin films prepared from metal xanthates,” J. Mater. Chem. A 3(47), 24155–24162 (2015). [CrossRef]  

59. AA Hassan and NFA Al-Rushed, “Structural and Optical Properties of CuAlO2 Thin Film Prepared by Spray Pyrolysis,” Eng. and Tech. Journal 33, Part-B(4), 602–611 (2015).

60. A. El Kissani, H. Ait Dads, S. Oucharrou, F. Welatta, H. Elaakib, L. Nkhaili, A. Narjis, A. Khalfi, K. El Assail, and A. Outzourhit, “A facile route for synthesis of cadmium sulfide thin films,” Thin Solid Films 664, 66–69 (2018). [CrossRef]  

61. Y. Bouachiba, A. Bouabellou, F. Hanini, F. Kermiche, A. Taabouche, and K. Boukheddaden, “Structural and optical properties of TiO2 thin films grown by sol-gel dip coating process,” Mater Sci-Pol 32(1), 1–6 (2014). [CrossRef]  

62. G. K. Gupta and A. Dixit, “Simulation studies on photovoltaic response of ultrathin CuSb(S/Se2) ternary compound semiconductors absorber-based single junction solar cells,” Int. J. Energy Res. 44(5), 3724–3736 (2020). [CrossRef]  

63. W. Shockley and H. J. Queisser, “Detailed Balance Limit of Efficiency of p-n Junction Solar Cells,” J. Appl. Phys. 32(3), 510–519 (1961). [CrossRef]  

64. Pravin Sadanand, Shambhavi Kumar Singh, Pooja Rai, D.K. Lohia, and Dwivedi, “Comparative study of the CZTS, CuSbS2 and CuSbSe2 solar photovoltaic cell with an earth-abundant non-toxic buffer layer,” Sol. Energy 222, 175–185 (2021). [CrossRef]  

65. CM Wolfe, N Holonyak, and GE Stillman, “Physical properties of semiconductors,” First Edition, Prentice Hall, 1989.

66. Y. Alivov, V. Singh, Y. Ding, L. J. Cerkovnik, and P. Nagpal, “Doping of wide-band gap titanium-dioxide nanotubes: optical, electronic and magnetic properties,” Nanoscale 6(18), 10839–10849 (2014). [CrossRef]  

67. S. Paul, S. Grover, I. L. Repins, B. M. Keyes, M. A. Contreras, K. Ramanathan, R. Noufi, Z. Zhao, F. Liao, and J. V. Li, “Analysis of Back-Contact Interface Recombination in Thin-Film Solar Cells,” IEEE J. Photovoltaics 8(3), 871–878 (2018). [CrossRef]  

68. Y. Cao, X. Zhu, H. Chen, X. Zhang, J. Zhou, Z. Hu, and J. Pang, “Towards high efficiency inverted Sb2Se3 thin film solar cells,” Sol. Energy Mater. Sol. Cells 200, 109945 (2019). [CrossRef]  

69. M. DjinkwiWanda, S. Ouédraogo, F. Tchoffo, F. Zougmoré, and J. M. B. Ndjaka, “Numerical Investigations and Analysis of Cu2ZnSnS4 Based Solar Cells by SCAPS-1D,” Int. J. Photoenergy 2016, 2152018 (2016).

70. J. Tao, X. Hu, J. Xue, Y. Wang, G. Weng, S. Chen, Z. Zhu, and J. Chu, “Investigation of electronic transport mechanisms in Sb2Se3 thin-film solar cells,” Sol. Energy Mater. Sol. Cells 197, 1–6 (2019). [CrossRef]  

71. Y. Xiao, H. Wang, and H. Kuang, “Numerical simulation and performance optimization of Sb2S3 solar cell with a hole transport layer,” Opt. Mater. 108, 110414 (2020). [CrossRef]  

72. Z. Gu, F. Chen, X. Zhang, Y. Liu, C. Fan, G. Wu, H. Li, and H. Chen, “Novel planar heterostructure perovskite solar cells with CdS nanorods array as electron transport layer,” Sol. Energy Mater. Sol. Cells 140, 396–404 (2015). [CrossRef]  

73. S. Demtsu and J. Sites, “Effect of back-contact barrier on thin-film CdTe solar cells,” Thin Solid Films 510(1-2), 320–324 (2006). [CrossRef]  

74. S Sze and KK Ng, Physics of Semiconductor Devices (John Wiley and Sons, Inc., 1981).

75. P. Singh and N. Ravindra, “Temperature dependence of solar cell performance-an analysis,” Sol. Energy Mater. Sol. Cells 101, 36–45 (2012). [CrossRef]  

76. J. Li, H. Wang, M. Luo, J. Tang, C. Chen, W. Liu, F. Liu, Y. Sun, J. Hand, and Y. Zhang, “10% Efficiency Cu2ZnSn(S,Se)4 thin film solar cells fabricated by magnetron sputtering with enlarged depletion region width,” Sol. Energy Mater. Sol. Cells 149, 242–249 (2016). [CrossRef]  

77. M. Fathi, M. Abderrezek, F. Djahli, and M. Ayad, “Study of thin film solar cells in high temperature condition,” Energy Procedia 74, 1410–1417 (2015). [CrossRef]  

78. SR Rummel and TJ McMahon, “Effect of cell shunt resistance on PV module performance at reduced light levels,” In: AIP Conference Proceedings353, 581–586 (1996).

79. Most. Marzia Khatun, Adil Sunny, and Sheikh Rashel Al Ahmed, “Numerical investigation on performance improvement of WS2 thin-film solar cell with copper iodide as hole transport layer,” Sol. Energy 224, 956–965 (2021). [CrossRef]  

Data availability

The main parameters to carry out the same simulations using SCAPS-1D are all given, and the calculations may easily be reproduced. The readers can also obtain the original data that underlies the present paper directly from the author.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1.
Fig. 1. (a) The schematic structure of the proposed (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) n-p-p+-heterojunction solar cell, and (b) J-V characteristics under illumination for CuSbS2-based single and heterojunction solar cells with different device configurations.
Fig. 2.
Fig. 2. Schematic energy band diagram of the proposed (Al/ITO/n-TiO2/p-CuSbS2/p+-CuAlO2:Mg/Au) heterojunction solar cell.
Fig. 3.
Fig. 3. Intensity-dependent Voc with different excitation wavelength of white light (WL), λexc= 450 nm, and λexc= 800 nm at constant temperature (300 K) for (a) Cell-1 and, (b) Cell-4.
Fig. 4.
Fig. 4. PV responses of CuSbS2-based solar cells as a function of absorber layer (a) thickness and (b) acceptor concentration.
Fig. 5.
Fig. 5. PV responses of CuSbS2-based solar cells as a function of (a) TiO2 ETL, and (b) CuAlO2:Mg BSF layer carrier concentration.
Fig. 6.
Fig. 6. PV responses of CuSbS2-based solar cells as a function of defect density (a) at CuAlO2:Mg/CuSbS2, and (b) TiO2/CuSbS2 interfaces.
Fig. 7.
Fig. 7. PV responses of CuSbS2-based solar cells as a function of (a) work function of the back-contact materials, and (b) working temperature.
Fig. 8.
Fig. 8. PV responses of CuSbS2-based solar cells as a function of (a) series resistance, and (b) shunt resistance.

Tables (4)

Tables Icon

Table 1. Baseline material parameters, used for simulating CuSbS2-based heterojunction ultrathin film solar cells [5154].a

Tables Icon

Table 2. Electrical properties of front and back contact materials used in simulation

Tables Icon

Table 3. PV performance parameters for CuSbS2-based solar cells with different cell configurations

Tables Icon

Table 4. Extracted values of K1 and K2, and recombination coefficients for CuSbS2-based solar cells of different configurations. R 0 d and R o i , b are calculated at bias voltage of 0.9 V.

Equations (18)

Equations on this page are rendered with MathJax. Learn more.

Ψ n p = Δ E C Δ E V 2 q + k T q l n N D N A n i n n i p + k T 2 q l n N v n N c p N v p N c n
Ψ p + p = Δ E C Δ E V 2 q + k T q l n N a N A n i p n i p + + k T 2 q l n N v p N c p + N v p + N c p
Ψ = Ψ n p + Ψ p + p
R i , f = R o i , f e q V k B T
R d = R o d e q V 2 k B T
R b = R o b e q V k B T
R i , b = R o i , b e q V k B T
V o c = 2 k T q ln [ K 1 ( G a K 2 + 1 1 )
K 1 = 1 2 R 0 d R 0 i , f + R 0 b + R 0 i , b
K 2 = 4 W ( R 0 i , f + R 0 b + R 0 i , b ) ( R 0 d ) 2
R 0 d = 2 W K 1 K 2
K 1 S W = 1 2 R o d ( R o i , f + R o b )
K 2 S W = 4 W ( R o i , f + R o b ) ( R o d ) 2
K 1 L W = 1 2 R o d ( R o i , f + R o b + R o i , b )
K 2 L W = 4 W ( R o i , f + R o b + R o i , b ) ( R o d ) 2
R o i , b = 1 2 [ 1 K 1 L W 1 K 1 S W ] × R o d
Ψ B = E g q + χ φ m
C T = 1 η S T C d η T d T × 100 %
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.