A potential solar absorber material, sputtered kesterite Cu2ZnSnS4 (CZTS) thin film, has been extensively studied in recent years due to its advantageous properties, including the earth abundance of its constituent elements, nontoxicity, suitable band gap, and high absorption coefficient. 2000 nm CZTS thin films were deposited on soda lime glass by a sputtering technique. The prepared films underwent a postannealing treatment for crystallization in which different temperatures and pressures were applied to understand its impact on film growth, phase formation, and stoichiometry. The annealed samples were subsequently characterized by Raman and UV-visible (UV-Vis) spectroscopy, energy-dispersive X-ray spectroscopy (EDX), X-ray powder diffraction (XRD), scanning electron microscopy (SEM), and atomic force microscopy (AFM). The thickness of each film was measured using a surface profilometer and from a cross-sectional image obtained by SEM. The XRD pattern for each film showed characteristic (112), (220), and (312) peaks, and the phase purity was confirmed via Raman studies. Film surface morphology and roughness were studied by AFM. The root mean square roughness was found to increase with annealing temperature and base pressure. The chemical compositions of the prepared samples were analyzed by EDX, and the films showed desired stoichiometry. UV-Vis absorption spectroscopy indicated that the direct band gap energies (Eg) of the films were 1.47 eV–1.51 eV, within the optimum range for use in solar cells. These attractive properties of the sputtered CZTS thin film should heighten interest in its use as a solar absorber layer in the next-generation photovoltaic cells, suggesting that it possesses substantial commercial promise.

Photovoltaic solar power generation has grown exponentially in recent decades due to increasing energy consumption. Researchers in this field have focused on mastering low-cost and high efficient photovoltaic device manufacture. The most widely used Si-based solar cells exhibit high conversion efficiencies of up to 24.5%.1 Thin film solar cells are becoming increasingly prominent due to their high power conversion efficiencies (PCEs) and direct and tunable band gap energies (Eg), and because less material is required to construct them compared to conventional Si-based solar cells. The PCE of an amorphous silicon (a-Si) thin film can reach 13.6%,2 whereas solar cells based on copper indium selenide (CIS), copper indium gallium selenide (CIGS), and cadmium telluride (CdTe) offer higher PCEs of up to 22%.3 Another potential solar absorber material is the p-type chalcogenide semiconductor Cu2ZnSnS4 (referred to as CZTS), which is derived from the chalcopyrite structure of CIGS by replacing In and Ga with less expensive and more earth-abundant elements such as Zn and Sn.4–7 This material also has the advantage that it does not contain cadmium, which is toxic. Solar cells based on CZTS show great promise, as they exhibit high absorption coefficients8 (>104 cm−1) and a tunable band gap (Eg ∼ 1.45 eV–1.6 eV) that can boost the PCE. Among the various possible structures of CZTS, kesterite shows greater stability than stannite and wurzite, as kesterite possesses lower energy than those of the other two structures.9 

Different methods of growing CZTS thin films have been reported, including vacuum and nonvacuum (i.e., solution-based) deposition approaches. The use of vacuum-based technologies such as thermal evaporation,10 sputtering,11,12 electron beam evaporation,13 and pulsed laser deposition (PLD)14,15 increases manufacturing costs but results in high-quality thin film. Manufacturing costs can be lowered through the use of solution-based techniques such as sol-gel dip16 and spin17 coating, chemical bath deposition,18 screen printing,19,20 spray pyrolysis21 and electrodeposition,22 and successive ionic layer adsorption and reaction (SILAR).23 

In 2018, Yan et al.24 reported that heterojunction heat treatment of a CZTS-based solar cell permitted a PCE of >10%. Green et al.25 achieved a PCE of 11% and an open circuit voltage of 730.6 mV using a CZTS-based solar cell. A record PCE of around 12.6% was attained for a Cu2ZnSnS4-based solar cell fabricated by Wang et al. using a pure hydrazine solution approach.26 Saha and Alam reported that a maximum efficiency of 17.59% at 550 nm and an open circuit voltage of 940 mV were observed for a silver mixed CZTS (ACZTS) cell with a CdS/ACZTS/CZTS/ITO structure, according to optoelectronic simulations.27 Nevertheless, the PCEs of CZTS cells are yet to reach the efficiencies of solar cells based on CIGS and CdTe. According to the Shockley–Queisser theory,28 a maximum PCE of 32.2% is possible using CZTS. Further research into CZTS solar cells is, therefore, needed to improve their PCEs, as this would lead to a more prominent role for CZTS solar cells in energy generation in the future.

The application of kesterite in solar cells has attracted considerable attention despite the commercialization of direct band gap absorbers such as CdTe, CIS, and CIGS/Se. Great efforts have been directed into the design of an optimized device structure that can accommodate current CZTS absorbers, which have relatively high defect levels and short carrier life times. The usage of reduced materials necessitates the application of an ultrathin crystalline absorber film. The issue of optical losses29 and implementation of light trapping30,31 can play a vital role for higher PCE of the CZTS film.

The novelty of the present work is the postannealing parameter optimization after successful growth of quaternary CZTS thin film on a transparent glass substrate using a single sputtering system at 200 °C with a delicate CZTS target. When designing a highly efficient solar cell, the appropriate absorber layer thickness for light trapping can be determined by comparing the absorption length with the carrier diffusion length in the layer.32 This layer thickness affects the overall device design.29 The present study demonstrates the effectiveness of adjusting the annealing parameters (the pressure and the temperature) to obtain the CZTS film of the desired thickness for use in solar cells. Sulfurization is an indispensable step as sulfur is apt to be deficient in most cases, which also makes the system expensive. To obtain the desired stoichiometry, sulfurization is obvious for some deposition processes. In addition, the selenization method could have some advantages in obtaining smooth, uniform, and compact films. However, both selenization and sulfurization processes are skipped here to employ a low cost method. As it is a quaternary compound, CZTS contains some binary and ternary phases that need to overcome to obtain the pure phase only. Moreover, there are numerous issues to address, such as the grain size, surface smoothness, roughness, stoichiometric ratio, volatility of tin, and pinholes in the film surface when seeking to obtain a highly textured CZTS film with good coverage. In the investigation reported here, the crystal structure, surface morphology, composition, and optical properties of postannealed CZTS thin film samples were rigorously studied to better understand and optimize the growth and phase formation of CZTS thin films. This knowledge should facilitate attempts to fabricate CZTS thin films with the optimal stoichiometry and photovoltaic properties for use in solar cells.

Soda lime glass (SLG) substrates (3 × 3 cm2) were washed with detergent, brushed, and sequentially cleaned with methanol, ethanol, acetone, and deionized water in an ultrasonic bath for 10 min each. The glass substrates were then rinsed with deionized water and dried. The quaternary CZTS was deposited by a radiofrequency (RF) magnetron sputtering system (NSC-4000, NANO-MASTER, Inc., USA) operated at 80 W for 2 h. The target used in this study was manufactured by Kurt J. Lesker Co. (USA), and was 2 in diameter and 0.12 5 thick. The composition of the target was 2:1:1:4 C:Zn:Sn:S, and its purity was 99.99%. A voltage was applied between the target material (cathode) and the substrate (anode) to be coated with the target material. Electrons from the surface of the target ionized the process gas, resulting in the formation of a plasma. The chamber pressure was reduced as much as possible to stop the background gases from reacting with the film or sputter target. Careful control of the partial pressures of the reactive gases led to the growth of a thin film on the substrate via sputtering. The target to the substrate distance was set to 7 cm. When the base pressure reached 6.3 × 10−6 Torr, precursors were deposited on the SLG inside a sputter coater at a working pressure of 4.5 mTorr. An Ar flow of 5 SCCM was maintained throughout the process. The substrate holder was rotated at 20 rpm to maximize film uniformity and target utilization. The temperature of the chamber during deposition was maintained at 200 °C.

A successful heat treatment strategy for CZTS consisting of a two-stage process in which precursors were prepared at a substrate temperature of <300 °C and then annealed at high temperature (around 500 °C) has been reported.15 In the present work, postannealing was conducted at two different temperatures, 470 °C and 560 °C, under an N2 atmosphere in a tube furnace (GSL-1100X, MTI Corporation, USA). Samples were maintained at the selected temperature at four different base pressures (150 Torr, 250 Torr, 350 Torr, or 450 Torr) for 30 min. Afterwards, the samples were left to cool within the annealing chamber.

X-ray powder diffraction (XRD) and Raman spectroscopy were performed on a bare CZTS absorber layer to characterize the microstructure of the CZTS film. XRD patterns were recorded using an EMMA X-ray diffractometer (GBC Corporation, Australia) using Cu-Kα radiation with a wavelength of 1.5406 Å. The unit was operated at 35 kV and 28 mA. Raman spectra (excitation wavelength: 785 nm) were obtained using a MacroRAM Raman spectrometer (Horiba Scientific, Japan) with a tunable output power of 7 mW–450 mW in the backscattering configuration. This spectrometer was equipped with a 685 gr/mm holographic grating for superior stray light rejection. An UV-visible (UV-Vis) spectrophotometer (UH4150, Hitachi, Japan) was used to study the optical absorbance at visible wavelengths. The thickness of the film was measured with a surface profilometer (Dektak, Bruker, USA), and the precursor morphology was examined by atomic force microscopy (AFM; Flex AFM, C3000, Nanosurf, Switzerland) and scanning electron microscopy (SEM; EVO18, Carl Zeiss, UK). The chemical composition of the film was measured by energy-dispersive X-ray spectroscopy (EDX) using a Team EDS system (EDAX, AMETEK, USA; beam voltage: 15 kV) attached to the SEM unit.

Table I lists the annealing conditions and the average film thickness of the samples measured using both a surface profilometer and SEM. There was no significant variation in the thickness among the samples annealed under different conditions. Figure 1 shows a SEM image of the cross-section of a CZTS thin film.

TABLE I.

CZTS thin film samples annealed under different conditions.

Annealing Base pressure for Thickness via surface Thickness via
Sample temperature (°C) annealing (Torr) profilometer (μm) SEM (μm)
S-560-150  560  150  2.112  2.156 
S-560-250  250 
S-560-350  350 
S-560-450  450 
S-470-150  470  150 
S-470-250  250 
S-470-350  350 
S-470-450  450 
Annealing Base pressure for Thickness via surface Thickness via
Sample temperature (°C) annealing (Torr) profilometer (μm) SEM (μm)
S-560-150  560  150  2.112  2.156 
S-560-250  250 
S-560-350  350 
S-560-450  450 
S-470-150  470  150 
S-470-250  250 
S-470-350  350 
S-470-450  450 
FIG. 1.

Example of the utilization of a SEM image to measure the cross-section of a CZTS thin film (sample S-560-450) annealed under a N2 atmosphere.

FIG. 1.

Example of the utilization of a SEM image to measure the cross-section of a CZTS thin film (sample S-560-450) annealed under a N2 atmosphere.

Close modal

The deposited CZTS films exhibited an amorphous nature. Polycrystalline CZTS thin films with the crystal structure of kesterite were obtained by annealing. The annealing process causes the structure of the crystal to shift from a nonequilibrium state to a state closer to thermodynamic equilibrium. The XRD profiles of CZTS thin films annealed under different conditions are shown in Fig. 2 (JCPDS-004751). The XRD profiles of all the samples exhibit a particularly intense (112) peak as well as less intense (200), (220), and (312) peaks. It is worth noting that the peak positions do not shift depending on the annealing conditions, indicating that the CZTS phase is stable.

FIG. 2.

XRD profiles of CZTS samples annealed under different conditions.

FIG. 2.

XRD profiles of CZTS samples annealed under different conditions.

Close modal

Bragg peaks from the Cu2−xS (x = 1) phase were observed in the XRD patterns of the thin films annealed at 470 °C (JCPDS-06-0464). The presence of a Cu2−xS impurity phase indicates a significantly increased Cu fraction in the film, although this promotes the growth of single-phase CZTS. Sun et al. reported the appearance of this impurity in CZTS films deposited at 350 °C–400 °C.33 Tang et al. reported a diffraction peak from a CuxS phase in a film that had been sulfurized at 450 °C.34 This peak disappeared when the CZTS film was annealed at 500 °C, 525 °C, or 550 °C. Similarly, in the present investigation, Bragg peaks from a CuS phase were seen for the samples annealed at a higher temperature (560 °C). It is possible that at this annealing temperature, the Cu:Zn:Sn:S ratio has reached ideal stoichiometric condition, which diminishes impure phases such as CuS. Thus, the Bragg peaks disappear when the higher annealing temperature is applied.35 

Figure 3 shows that Bragg peaks36 for the (112), (200), (220), and (312) planes appear in the XRD patterns of the CZTS thin film samples.16,17 Two more characteristic peaks of kesterite that are visible in Fig. 2 at 2θ = 38.07° and 52.06°can also be seen in Fig. 3. This figure also shows that the (112) peak of S-560-450 is sharper than that of S-470-450, indicating that the peak sharpness increases with annealing temperature, in agreement with previous reports.33,37

FIG. 3.

X-ray diffraction patterns of two CZTS thin film samples (JCPDS 00-4751) annealed at the same base pressure (450 Torr), but at different temperatures: 470 °C (top) and 560 °C (bottom).

FIG. 3.

X-ray diffraction patterns of two CZTS thin film samples (JCPDS 00-4751) annealed at the same base pressure (450 Torr), but at different temperatures: 470 °C (top) and 560 °C (bottom).

Close modal

In addition, the full width at half maximum (FWHM) of a diffraction peak increases with decreasing crystallite size. The average crystallite size can be estimated from the FWHM of the most intense peak (112) using the Scherrer formula38 (assuming that the nanocrystals are spherical) as

(1)

where β is the FWHM in radians, λ is the wavelength, and k is the Scherrer constant. The calculated average crystallite size, D, of each thin film sample is shown in Table II. It is clear that the samples annealed at the higher temperature (560 °C; samples S-560-150,S-560-250, S-560-350, and S-560-450) have lower FWHMs and larger crystallite sizes than the samples annealed at the lower temperature (470 °C). The crystallization temperature plays an important role in the morphology of the final structure. Increasing the temperature causes greater mobility and surface diffusion, which allows the deposited material to migrate and coalesce into larger islands, leading to a larger average crystallite size and a lower FWHM. The dislocation density δ is also lower for the samples annealed at a higher temperature, which indicates that these samples have fewer lattice imperfections and decreased strain.39,40 However, the crystallite size does not appear to vary significantly with increasing base pressure.

TABLE II.

Microstructural parameters of the CZTS thin film samples in the (112) plane.

d spacing, FWHM, Crystallite size, Dislocation density, Strain,
Sample Δ (Å) β (deg) D (nm) δ (1015 line/m2) ε (×10−3)
S-560-150  3.162  0.38  21.32  2.20  6.60 
S-560-250  3.096  0.41  19.78  2.55  6.97 
S-560-350  3.092  0.466  17.41  3.30  7.91 
S-560-450  3.116  0.404  20.07  2.48  6.91 
S-470-150  3.159  0.524  15.46  4.18  9.10 
S-470-250  3.147  0.488  16.61  3.63  8.43 
S-470-350  3.161  0.514  15.76  4.03  8.93 
S-470-450  3.119  0.564  14.38  4.84  9.66 
d spacing, FWHM, Crystallite size, Dislocation density, Strain,
Sample Δ (Å) β (deg) D (nm) δ (1015 line/m2) ε (×10−3)
S-560-150  3.162  0.38  21.32  2.20  6.60 
S-560-250  3.096  0.41  19.78  2.55  6.97 
S-560-350  3.092  0.466  17.41  3.30  7.91 
S-560-450  3.116  0.404  20.07  2.48  6.91 
S-470-150  3.159  0.524  15.46  4.18  9.10 
S-470-250  3.147  0.488  16.61  3.63  8.43 
S-470-350  3.161  0.514  15.76  4.03  8.93 
S-470-450  3.119  0.564  14.38  4.84  9.66 

The dislocation density,41 the number of dislocations in a unit volume of a crystalline material, is calculated as

(2)

Crystal dislocation increases the strain in nanostructured materials. In the Williamson–Hall isotropic strain model (W-HISM),42 the crystallite size and lattice strain are independent of each other. According to the Williamson–Hall anisotropic uniform deformation energy density model (W-HUDEDM), there are more volume defects at grain boundaries when the crystallites are very small. The internal pressure exerted by the surface tension due to volume defects creates a stress field. This additional stress at grain boundaries causes lattice strain. For simplicity, the strain, ε, of the thin film is measured as41 

(3)

It is difficult to distinguish experimentally between tetragonal CTS, cubic CTS, sphalerite (ZnS), and CZTS as they present very similar lattice parameters.36,43 To detect traditional issues of phase segregation (secondary phases), the Raman spectrum in the wave number region 140 cm−1–495 cm−1 was obtained for a thin film sample at room temperature; this spectrum is shown in Fig. 4. It presents well-defined peaks at 288 cm−1, 334 cm−1, and 361 cm−1, corresponding to the reported characteristic peaks for polycrystalline kesterite films.44 Moreover, it is obvious that there are no extra peaks due to the presence of other compounds such as SnS2 (314 cm−1), Cu2SnS3 (352 cm−1 and 374 cm−1), cubic ZnS (352 cm−1 and 275 cm−1), and orthorhombic Cu2SnS3 (318 cm−1), which confirms that the sample consists of single-phase CZTS.45,46

FIG. 4.

Raman spectrum of a CZTS thin film sample (S-560-450) annealed at 560 °C.

FIG. 4.

Raman spectrum of a CZTS thin film sample (S-560-450) annealed at 560 °C.

Close modal

The morphology of the thin film is one of the most important properties to understand the surface phenomena. Figure 5 shows AFM micrographs of a 25 µm2 region of the surface of each annealed CZTS thin film sample. These images reveal that the sputtered films have similar surface topographies and thicknesses and that the Volmer–Weber mode was the dominant growth mode. The films show far better coverage than a film deposited by a sol–gel technique.17 The variation in the root mean square (rms) surface roughness of the thin film derived from the AFM image with annealing pressure and temperature is shown in Fig. 6. This figure indicates that the surface roughness increases significantly with increasing annealing temperature. The surface roughness barely changes with increasing base pressure when the annealing temperature is 470 °C, but it changes drastically upon increasing the base pressure from 350 Torr to 450 Torr, when the annealing temperature is 560 °C. Changing the annealing temperature shifts the phase transition of the CZTS thin film.37 The average grain size increases with increasing annealing temperature.47 At high temperatures, grain coalescence and reorganization occur due to increased surface mobility. Figure 5(i) shows the topography of the CZTS thin film sample S-470-150, based on AFM data. Table II shows that when the temperature is higher than 500 °C, the grains in the film become larger.

FIG. 5.

AFM images of CZTS thin film samples: (a) S-560-150, (b) S-560-250, (c) S-560-350, (d) S-560-450, (e) S-470-150, (f) S-470-250, (g) S-470-350, and (h) S-470-450. (i) AFM topographic image of the CZTS thin film sample S-470-150.

FIG. 5.

AFM images of CZTS thin film samples: (a) S-560-150, (b) S-560-250, (c) S-560-350, (d) S-560-450, (e) S-470-150, (f) S-470-250, (g) S-470-350, and (h) S-470-450. (i) AFM topographic image of the CZTS thin film sample S-470-150.

Close modal
FIG. 6.

Variation in the rms roughness with annealing pressure and temperature (560 °C and 470 °C).

FIG. 6.

Variation in the rms roughness with annealing pressure and temperature (560 °C and 470 °C).

Close modal

The analysis of the AFM images shown in Fig. 6 confirmed that the grain size and surface roughness of the CZTS thin film increased with increasing annealing temperature. However, some voids were observed on the surface of the film. This was due to the significant loss of Sn at the higher annealing temperature.48 Nevertheless, annealing appears to be useful for enhancing the crystallinity and phase formation of the film and is, therefore, beneficial for device quality. Our results indicate that CZTS thin films should be annealed at a temperature of at least 500 °C because an increased grain size leads to a lower recombination rate, which enhances the PCE.

The atomic percentages (at. %) for the CZTS thin film samples were determined using EDX and are shown in Table III. The theoretical stoichiometry of CZTS is Cu:Zn:Sn:S = 2:1:1:4. It is evident from Table III that the films are Cu rich and Zn poor compared to the expected stoichiometry. Chen et al. reported that Cu-rich and Zn-poor conditions are suitable for the growth of single-phase CZTS films.49,50 Cu is a lighter element and, therefore, has a higher flow speed than Zn, which leads to the enrichment of Cu in the film.33 Zn and S are more volatile and have lower melting points than the other two elements in CZTS, which results in Zn- and S-deficiency in CZTS thin films.51 However, even without sulfurization, the sulfur-to-metal ratios of the present samples are compatible with those stated in the previous reports.45,46 The films annealed at 560 °C (S-560-150, S-560-250, S-560-350, S-560-450) are clearly Sn deficient compared to the other samples. This undesirable phenomenon is likely to be due to the increasing loss of volatile elements or compounds at higher crystallization temperatures—an effect that has a significant impact on the final composition. An additional disadvantage of applying a relatively high annealing temperature is the possibility of diffusion through substrate interfaces, which may result in the formation of solid solutions. According to Weber et al.,48 the evaporation rates from different phases increase according to the sequence Cu2ZnSnS4 < Cu4SnS4 < Cu2SnS3 < SnS. SnS is lost from the CZTS thin film at a significant rate at temperatures of 550 °C and above. However, Todorov et al. reported the opposite composition i.e., Zn-rich and Cu-poor CZTS thin film yielded the highest conversion efficiency.52 

TABLE III.

Chemical composition data for the annealed CZTS thin films.

Sample Cu (at. %) Zn (at. %) Sn (at. %) S (at. %) Zn/Sn Cu/(Zn + Sn) S/(Cu + Zn + Sn)
S-560-150  28.98  10.04  10.2  50.78  0.98  1.43  1.03 
S-560-250  28.23  10.44  9.57  51.76  1.09  1.41  1.07 
S-560-350  29.2  8.66  11.4  50.74  0.76  1.46  1.03 
S-560-450  29.66  9.12  9.34  51.88  0.98  1.61  1.08 
S-470-150  27.01  12.07  12.38  48.54  0.97  1.10  0.94 
S-470-250  28.21  11.81  11.09  48.89  1.06  1.23  0.96 
S-470-350  28.3  8.01  11.52  52.17  0.70  1.45  1.09 
S-470-450  28.41  10.25  12.73  48.61  0.81  1.24  0.95 
Sample Cu (at. %) Zn (at. %) Sn (at. %) S (at. %) Zn/Sn Cu/(Zn + Sn) S/(Cu + Zn + Sn)
S-560-150  28.98  10.04  10.2  50.78  0.98  1.43  1.03 
S-560-250  28.23  10.44  9.57  51.76  1.09  1.41  1.07 
S-560-350  29.2  8.66  11.4  50.74  0.76  1.46  1.03 
S-560-450  29.66  9.12  9.34  51.88  0.98  1.61  1.08 
S-470-150  27.01  12.07  12.38  48.54  0.97  1.10  0.94 
S-470-250  28.21  11.81  11.09  48.89  1.06  1.23  0.96 
S-470-350  28.3  8.01  11.52  52.17  0.70  1.45  1.09 
S-470-450  28.41  10.25  12.73  48.61  0.81  1.24  0.95 

The UV-Vis absorption spectra of the CZTS thin film samples depicted in Fig. 7 indicate that the samples show high optical absorbance and a large absorption coefficient, α, in the visible region (ignoring reflection and transmission losses). The nature of the transition associated with absorption can be determined using the classical equation21 for the absorption coefficient

(4)

Here, n = 1/2 for an allowed direct transition and n = 2 for an allowed indirect transition. The absorption coefficient, α, can be calculated from the absorbance A using53 

(5)

where I is the transmitted intensity, Io is the incident intensity, A is the absorbance, and d is the thickness of the film. The absorption coefficients of the samples were found to be larger than 104 cm−1, which is consistent with previous reports.17,21,23 Studying the optical absorption of the samples allows us to explore properties such as the band structure, refractive index, and high-frequency dielectric constant,46 all of which are important influences on solar cell performance.

FIG. 7.

Absorption coefficient (α) vs photon energy plots and Tauc plots (inset) for CZTS thin films annealed at (a) 560 °C and (b) 470 °C.

FIG. 7.

Absorption coefficient (α) vs photon energy plots and Tauc plots (inset) for CZTS thin films annealed at (a) 560 °C and (b) 470 °C.

Close modal

The optical band gap determines the portion of the solar spectrum that a photovoltaic cell will absorb. The band gap energy Eg for direct transitions can be calculated by extrapolating the linear portion of the (αhν)2 vs plot, and the band gap energies of the samples calculated in this manner are shown in Table IV. These energies are observed to range between 1.47 eV and 1.51 eV. This variation in the band gap energy is largely explained by differences in the homogeneity and crystallinity of the thin films, which in turn leads to the differences in the crystallite size between the samples.54 Note that the calculated band gap values are close to the optimum band gap for solar cells.55 

TABLE IV.

Calculated values of the band gap energy (Eg), refractive index(n), and dielectric constant for the CZTS thin film samples annealed under different conditions.

High-frequency
Band gap Refractive dielectric Dielectric
Sample Eg (eV) index n constant ε constant εo
S-560-150  1.5  2.913  8.485  13.900 
S-560-250  1.48  2.923  8.542  13.962 
S-560-350  1.5  2.913  8.485  13.900 
S-560-450  1.5  2.913  8.485  13.900 
S-470-150  1.49  2.918  8.514  13.931 
S-470-250  1.47  2.928  8.571  13.992 
S-470-350  1.47  2.928  8.571  13.992 
S-470-450  1.51  2.908  8.457  13.869 
High-frequency
Band gap Refractive dielectric Dielectric
Sample Eg (eV) index n constant ε constant εo
S-560-150  1.5  2.913  8.485  13.900 
S-560-250  1.48  2.923  8.542  13.962 
S-560-350  1.5  2.913  8.485  13.900 
S-560-450  1.5  2.913  8.485  13.900 
S-470-150  1.49  2.918  8.514  13.931 
S-470-250  1.47  2.928  8.571  13.992 
S-470-350  1.47  2.928  8.571  13.992 
S-470-450  1.51  2.908  8.457  13.869 

The refractive index (n) of the thin film is a major determinant of the possibility of total internal reflection inside the solar cell.31 The refractive index of each of the CZTS thin films examined in this work was calculated using the Moss relation50 

(6)

where k is a constant with a value of 108 eV.

The dielectric properties of a material relate to its capacity to impede the electron movement when it is polarized under the influence of an external electric field. Materials with suitable dielectric constants are needed to develop efficient solar cells.

The high-frequency dielectric constant ε of each sample was determined using the relation17,23

(7)

The static dielectric constant εo of each thin film17,23 was calculated using the relation

(8)

The band gap, refractive index, and dielectric constant data shown in Table IV for the CZTS thin films annealed under different conditions are all in good agreement with previously reported results.23 

Here, we report the production of thin (∼2000 nm) Se-free kesterite CZTS films without notable phase segregation or voids via sputtering. Sample properties were analyzed after postannealing treatment to find out the optimum annealing conditions for the CZTS thin film. The Raman spectra of the samples verified that they were indeed kesterite based on the presence of a large peak at 334 cm−1 and additional small peaks at 288 cm−1 and 361 cm−1. The samples annealed at 560 °C also had larger crystallites than those annealed at 470 °C, as demonstrated by the sharper peaks in the XRD spectra of thin films annealed at 560 °C. The grains in the samples were compact, and agglomeration increased with annealing temperature, which increased the absorption coefficient. AFM analysis revealed that the rms roughness of the thin film annealed at 560 °C changed drastically upon increasing the base pressure from 350 Torr to 450 Torr. EDX analysis indicated that the CZTS samples produced without sulfurization were Cu and S rich but Zn and Sn poor. Significant SnS losses were found to occur at temperatures above 500 °C, resulting in the generation of voids on the film surface and affecting the stoichiometry of the film. UV-Vis absorption spectra indicated that the direct band gap energies (Eg) of the thin films ranged from 1.47 eV to 1.51 eV. Nevertheless, the notably improved CZTS film was realized in this work through postannealing parameter optimization, which could facilitate enhanced manufacturing throughput. We can conclude that samples annealed at a higher temperature (560 °C) show better crystallinity, as predicted, and the base pressure plays a vital role in the case of surface morphology. Future work in this area will focus on improving the morphology, chemical composition, and electric properties of the thin films, thereby enhancing their optical absorbance and stoichiometry to achieve further improvements in CZTS solar absorber quality.

The logistic support of the Solar Energy Technology Research Laboratory of IFRD, BCSIR, is gratefully acknowledged.

1.
J.
Zhao
,
A.
Wang
, and
M. A.
Green
, “
24·5% efficiency silicon PERT cells on MCZ substrates and 24·5% efficiency PERL cells on FZ substrates
,”
Prog. Photovoltaics
7
(
6
),
471
474
(
1999
).
2.
H.
Sai
 et al., “
Triple-junction thin-film silicon solar cell fabricated on periodically textured substrate with a stabilized efficiency of 13.6%
,”
Appl. Phys. Lett.
106
,
213902
(
2015
).
3.
M. A.
Green
 et al., “
Solar cell efficiency tables (version 50)
,”
Prog. Photovoltaics
25
(
7
),
668
676
(
2017
).
4.
M.
Jiang
and
X.
Yan
, “
Cu2ZnSnS4 thin film solar cells: Present status and future prospects
,” in
Solar Cells: Research and Application Perspectives
, edited by
A.
Morales-Acevedo
(
Intech
,
London
,
2013
), pp.
107
143
.
5.
X.
Song
,
X.
Ji
,
M.
Li
,
W.
Lin
,
X.
Luo
, and
H.
Zhang
, “
A review on development prospect of CZTS based thin film solar cells
,”
Int. J. Photoenergy
2014
,
613173
.
6.
A.
Haddout
,
A.
Raidou
, and
M.
Fahoume
, “
A review on the numerical modeling of CdS/CZTS-based solar cells
,”
Appl. Phys. A
125
,
124
(
2019
).
7.
A.
Kowsar
 et al., “
Progress in major thin-film solar cells: Growth technologies, layer materials and efficiencies
,”
Int. J. Renewable Energy Res.
9
(
2
),
579
597
(
2019
), see https://www.ijrer.org/ijrer/index.php/ijrer/article/view/9054/pdf.
8.
D. B.
Mitzi
,
O.
Gunawan
,
T. K.
Todorov
,
K.
Wang
, and
S.
Guha
, “
The path towards a high-performance solution-processed kesterite solar cell
,”
Sol. Energy Mater. Sol. Cells
95
,
1421
1436
(
2011
).
9.
T.
Maeda
,
S.
Nakamura
, and
T.
Wada
, “
Electronic structure and phase stability of In-free photovoltaic semiconductors, Cu2ZnSnSe4 and Cu2ZnSnS4 by first-principles calculation
,”
MRS. Proc.
1165
,
1165-M04-M03
(
2011
).
10.
B.
Shin
,
O.
Gunawan
,
Y.
Zhu
,
N. A.
Bojarczuk
,
S. J.
Chey
, and
S.
Guha
, “
Thin film solar cell with 8.4% power conversion efficiency using an earth-abundant Cu2ZnSnS4 absorber
,”
Prog. Photovoltaics
21
(
1
),
72
76
(
2013
).
11.
F.
Jiang
,
H.
Shen
, and
W.
Wang
, “
Optical and electrical properties of Cu2ZnSnS4 film prepared by sulfurization method
,”
J. Electron. Mater.
41
(
8
),
2204
2209
(
2012
).
12.
N.
Muhunthan
,
O. P.
Singh
,
S.
Singh
, and
V. N.
Singh
, “
Growth of CZTS thin films by cosputtering of metal targets and sulfurization in H2S
,”
Int. J. Photoenergy
2013
,
752012
.
13.
H.
Katagiri
 et al., “
Development of thin film solar cell based on Cu2ZnSnS4 thin films
,”
Sol. Energy Mater. Sol. Cells
65
,
141
148
(
2001
).
14.
A. V.
Moholkar
 et al., “
Development of CZTS thin films solar cells by pulsed laser deposition: Influence of pulse repetition rate
,”
Sol. Energy
85
,
1354
1363
(
2011
).
15.
A.
Cazzaniga
 et al., “
Ultra-thin Cu2ZnSnS4 solar cell by pulsed laser deposition
,”
Sol. Energy Mater. Sol. Cells
166
,
91
99
(
2017
).
16.
S.
Dilruba
,
T. P.
Ananna
,
A.
Sharmin
,
M. S.
Bashar
, and
Z. H.
Mahmood
, “
Properties of Cu2ZnSnS4 thin films fabricated by dip coating technique for solar cell application
,”
AIJRSTEM
26
(
1
),
170
178
(
2019
).
17.
A.
Sharmin
,
M. S.
Bashar
,
S.
Tabassum
, and
Z. H.
Mahmood
, “
Low cost and sol-gel processed earth abundant Cu2ZnSnS4thin film as an absorber layer for solar cell: Annealing without sulfurization
,”
IJTFST
8
(
2
),
65
74
(
2019
), see http://www.naturalspublishing.com/Article.asp?ArtcID=19727.
18.
N. M.
Shinde
,
C. D.
Lokhande
,
J. H.
Kim
, and
J. H.
Moon
, “
Low cost and large area novel chemical synthesis of Cu2ZnSnS4 (CZTS) thin films
,”
J. Photochem. Photobiol., A
235
,
14
20
(
2012
).
19.
X.
Sheng
,
L.
Wang
,
Y.
Tian
,
Y.
Luo
,
L.
Chang
, and
D.
Yang
, “
Low-cost fabrication of Cu2ZnSnS4 thin films for solar cell absorber layers
,”
J. Mater. Sci.: Mater. Electron.
24
(
2
),
548
552
(
2013
).
20.
W.
Wang
,
H.
Shen
,
F.
Jiang
,
X.
He
, and
Z.
Yue
, “
Low-cost chemical fabrication of Cu2ZnSnS4 microparticles and film
,”
J. Mater. Sci.: Mater. Electron.
24
(
6
),
1813
1817
(
2013
).
21.
N. M.
Shinde
,
R. J.
Deokate
, and
C. D.
Lokhan
, “
Properties of spray deposited Cu2ZnSnS4 (CZTS) thin films
,”
J. Anal. Appl. Pyrolysis
100
,
12
16
(
2013
).
22.
M.
Jeon
,
T.
Shimizu
, and
S.
Shingubara
, “
Cu2ZnSnS4 thin films and nanowires prepared by different single-step electrodeposition method in quaternary electrolyte
,”
Mater. Lett.
65
(
15–16
),
2364
2367
(
2011
).
23.
J.
Henry
,
K.
Mohanraj
, and
G.
Sivakumar
, “
Electrical and optical properties of CZTS thin films prepared by SILAR method
,”
J. Asian Ceram. Soc.
4
(
1
),
81
84
(
2016
).
24.
C.
Yan
 et al., “
Cu2ZnSnS4 solar cells with over 10% power conversion efficiency enabled by heterojunction heat treatment
,”
Nat. Energy
3
(
9
),
764
772
(
2018
).
25.
M. A.
Green
,
Y.
Hishikawa
,
E. D.
Dunlop
,
D. H.
Levi
,
J.
Hohl-Ebinger
, and
A. W. Y.
Ho-Baillie
, “
Solar cell efficiency tables (version 51)
,”
Prog. Photovoltaics
26
,
3
12
(
2018
).
26.
W.
Wang
,
M. T.
Winkler
,
O.
Gunawan
,
T.
Gokmen
,
T. K.
Todorov
,
Y.
Zhu
, and
D. B.
Mitzi
, “
Device characteristics of CZTSSe thin film solar cells with 12.6% efficiency
,”
Adv. Energy Mater.
4
(
7
),
1301465
(
2014
).
27.
U.
Saha
and
M. K.
Alam
, “
Boosting the efficiency of single junction kesterite solar cell using Ag mixed Cu2ZnSnS4 active layer
,”
RSC Adv.
8
,
4905
4913
(
2018
).
28.
W.
Shockley
and
H. J.
Queisser
, “
Detailed balance limit of efficiency of p-n junction solar cells
,”
J. Appl. Phys.
32
,
510
519
(
1961
).
29.
B.
Vermang
 et al., “
Employing Si solar cell technology to increase efficiency of ultra-thin Cu(In,Ga)Se2 solar cells
,”
Prog. Photovoltaics
22
,
1023
1029
(
2014
).
30.
C.
Van Lare
,
G.
Yin
,
A.
Polman
, and
M.
Schmid
, “
Light coupling and trapping in ultrathin Cu(In,Ga)Se2 solar cells using dielectric scattering patterns
,”
ACS Nano
9
,
9603
9613
(
2015
).
31.
R. S.
Dubey
 et al., “
Performance enhancement of thin film silicon solar cells based on distributed Bragg reflector & diffraction grating
,”
AIP Adv.
4
,
127121
(
2014
).
32.
F.
Liu
 et al., “
Beyond 8% ultrathin kesterite Cu2ZnSnS4 solar cell by interface reaction route controlling and self organized nanopattern at the contact
,”
NPG Asia Mater.
9
,
e401
(
2017
).
33.
L.
Sun
,
J.
He
,
H.
Kong
,
F.
Yue
,
P.
Yang
, and
J.
Chu
, “
Structure, composition and optical properties of Cu2ZnSnS4 thin films deposited by pulsed laser deposition method
,”
Sol. Energy Mater. Sol. Cells
95
(
10
),
2907
2913
(
2011
).
34.
D.
Tang
 et al., “
An alternative route towards low-cost Cu2ZnSnS4 thin film solar cells
,”
Surf. Coat. Tech.
232
,
53
59
(
2013
).
35.
X.
Lin
 et al., “
Structural and optical properties of Cu2ZnSnS4 thin film absorbers from ZnS and Cu3SnS4 nanoparticle precursors
,”
Thin Solid Films
535
,
10
13
(
2013
).
36.
N. M.
Shinde
 et al., “
Aqueous chemical growth of Cu2ZnSnS4 (CZTS) thin films: Air annealing and photoelectrochemical properties
,”
Mater. Res. Bull.
48
,
1760
1766
(
2013
).
37.
Sk.
Shahenoor Basha
and
M. C.
Rao
, “
Effect of annealing temperature on structural and morphological studies of electrodeposited CZTS thin films
,”
Ceram. Int.
44
(
1
),
648
656
(
2018
).
38.
P. K.
Nair
,
J.
Cardoso
,
O.
Gomez Daza
, and
M. T. S.
Nair
, “
Polyethersulfone foils as stable transparent substrates for conductive copper sulfide thin film coatings
,”
Thin Solid Films
401
(
1-2
),
243
250
(
2001
).
39.
R.
Rafique
,
K. N.
Tonny
,
A.
Sharmin
, and
Z. H.
Mahmood
, “
Study on the effect of varying film thickness on the transparent conductive nature of aluminum doped zinc oxide deposited by dip coating
,”
Mater. Focus
7
(
5
),
707
713
(
2018
).
40.
K. N.
Tonny
,
R.
Rafique
,
A.
Sharmin
,
M. S.
Bashar
, and
Z. H.
Mahmood
, “
Electrical, optical and structural properties of transparent conducting Al doped ZnO (AZO) deposited by sol-gel spin coating
,”
AIP Adv.
8
,
065307
(
2018
).
41.
A.
Sharmin
,
S.
Tabassum
,
M. S.
Bashar
, and
Z. H.
Mahmood
, “
Depositions and characterization of sol gel processed Al-doped ZnO (AZO) as transparent conducting oxide (TCO) for solar cell application
,”
J. Thoer. Appl. Phys.
13
(
2
),
123
132
(
2019
).
42.
P. M.
Shafi
and
A. C.
Bose
, “
Impact of crystalline defects and size on X-ray line broadening: A phenomenological approach for tetragonal SnO2 nanocrystals
,”
AIP Adv.
5
,
057137
(
2015
).
43.
A.
Irkhina
 et al., “
Metal acetate based synthesis of small sized Cu2ZnSnS4 nanocrystals: Effect of injection temperature and synthesis time
,”
RSC Adv.
7
,
11752
11760
(
2017
).
44.
N. M.
Shinde
 et al., “
Development of polyaniline/Cu2ZnSnS4 (CZTS) thin film based heterostructure as room temperature LPG sensor
,”
Sens. Actuators, A
193
,
79
86
(
2013
).
45.
A. G.
Kannan
,
T. E.
Manjulavallia
, and
J.
Chandrasekaran
, “
Influence of solvent on the properties of CZTS nanoparticles
,”
Procedia Eng.
141
,
15
22
(
2016
).
46.
R.
Sun
 et al., “
Cu2ZnSNSSe4 solar cells with 9.6% efficiency via selenizing Cu-Zn-Sn-S precursor sputtered from a quaternary target
,”
Sol. Energy Mater. Sol. Cells
174
,
42
49
(
2018
).
47.
G. D.
Surgina
 et al., “
Effect of annealing on structural and optical properties of Cu2ZnSnS4 thin films grown by pulsed laser deposition
,”
Thin Solid Films
594
(
A
),
74
79
(
2015
).
48.
A.
Weber
,
R.
Mainz
, and
H. W.
Schock
, “
On the Sn loss from thin films of the material system Cu-Zn-Sn-S in high vacuum
,”
J. Appl. Phys.
107
,
013516
(
2010
).
49.
S.
Chen
,
J. H.
Yang
,
X. G.
Gong
,
A.
Walsh
, and
S. H.
Wei
, “
Intrinsic point defects and complexes in the quartenary kesterite semiconductor Cu2ZnSnS4
,”
Phys. Rev. B
819
,
245204
(
2010
).
50.
S.
Chen
,
X. G.
Gong
,
A.
Walsh
, and
S. H.
Wei
, “
Defect physics of the kesterite thin-film solar cell absorber Cu2ZnSnS4
,”
Appl. Phys. Lett.
96
(
2
),
021902
(
2010
).
51.
V.
Kheraj
 et al., “
Synthesis and characterization of copper zinc tin sulphide(CZTS) compound for absorber material in solar cells
,”
J. Cryst. Growth
362
,
174
177
(
2013
).
52.
T. K.
Todorov
,
K. B.
Reuter
, and
D. B.
Mitzi
, “
High-efficiency solar cell with earth-abundant liquid-processed absorber
,”
Adv. Mater.
22
,
E156
E159
(
2010
).
53.
J. H.
Lambert
,
D. A.
Lightner
,
H. F.
Shurvelland
, and
R. G.
Cooks
,
Introduction to Organic Spectroscopy
(
Macmillan
,
New York
,
1987
).
54.
M. C.
Rao
and
Sk.
Shahenoor Basha
, “
Structural and electrical properties of CZTS thin films by electrodeposition
,”
Results Phys.
9
,
996
1006
(
2018
).
55.
S. M.
Power
 et al., “
Single step electrosynthesis of Cu2ZnSnS4 (CZTS) thin film for solar cell application
,”
Electrochim. Acta
55
(
12
),
4057
4061
(
2010
).