Next Article in Journal
Investigation of Ionic Polymers’ Stabilizing and Flocculating Properties in Dispersed Activated Carbons Systems
Next Article in Special Issue
Solvent-Free Reaction for Unsymmetrical Organodisulfides with High Purity and Application as Cathode-Active Materials
Previous Article in Journal
Evaluation of Wetting Behaviors of Liquid Sodium on Transition Metals: An Experimental and Molecular Dynamics Simulation Study
Previous Article in Special Issue
Synthesis and Characterization of Type II Ge-Si Clathrate Films for Optoelectronic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of (Ge0.42Sn0.58)S Thin Films via Co-Evaporation and Their Solar Cell Applications

National Institute of Technology (KOSEN), Nagaoka College, Niigata 940-8532, Japan
*
Author to whom correspondence should be addressed.
Materials 2024, 17(3), 692; https://doi.org/10.3390/ma17030692
Submission received: 30 July 2023 / Revised: 28 January 2024 / Accepted: 29 January 2024 / Published: 1 February 2024

Abstract

:
In this study, as a novel approach to thin-film solar cells based on tin sulfide, an environmentally friendly material, we attempted to fabricate (Ge, Sn)S thin films for application in multi-junction solar cells. A (Ge0.42 Sn0.58)S thin film was prepared via co-evaporation. The (Ge0.42 Sn0.58)S thin film formed a (Ge, Sn)S solid solution, as confirmed by X-ray diffraction (XRD) and Raman spectroscopy analyses. The open circuit voltage (Voc), short circuit current density (Jsc), fill factor (FF), and power conversion efficiency (PCE) of (Ge0.42 Sn0.58)S thin-film solar cells were 0.29 V, 6.92 mA/cm2, 0.34, and 0.67%, respectively; moreover, the device showed a band gap of 1.42–1.52 eV. We showed that solar cells can be realized even in a composition range with a relatively higher Ge concentration than the (Ge, Sn)S solar cells reported to date. This result enhances the feasibility of multi-junction SnS-system thin-film solar cells.

1. Introduction

Developing energy resources to replace fossil fuels is an important goal of current energy research. Photovoltaic (PV) power generation is a promising technology for sustainable energy production. As the demand for renewable energy increases, there is a need to increase the number of solar cell installations and to fabricate higher efficiency solar cells. To increase the number of solar cell installations, solar cells need to be light and flexible enough to be installed in previously impossible places, such as on curved surfaces or delicate buildings. This can be achieved by making the material thinner and forming it on a flexible substrate; therefore, compound thin-film semiconductors that can sufficiently absorb light are attracting research attention. Multi-junction compound solar cells consisting of materials with different band gaps have been widely studied and demonstrated to absorb sunlight efficiently. The light-absorbing layers of such multi-junction solar cells should exhibit adjustable bandgaps. Thin-film solar cells fabricated with one such material, Cu(In,Ga)Se2 (CIGS), exhibited a power conversion efficiency of over 23% [1]. However, CIGS contains several rare metals, such as the toxic elements Se, In, and Ga, and the resource constraints and environmental impact of its mass production are concerning. It is therefore vital to develop new alternative materials. Thus, this study focuses on a compound semiconductor—a tin sulfide (SnS)-based compound—which can be an alternative to CIGS regarding resources, cost, and environment.
SnS is a binary chalcogenide compound composed of tin and sulfur. It shows high potential as an environmentally friendly material because it is composed of low-toxicity elements abundant in the Earth’s crust. The structural polymorphism of SnS at room temperature is known to be orthorhombic SnS (Pnma), rock salt SnS (P213), and sphalerite SnS (F-43 m) [2,3,4,5]. This study focuses on the most studied form, orthorhombic SnS. Note that from here on, orthorhombic SnS is denoted as SnS.
SnS thin films can be prepared by various methods, including thermal evaporation [6,7], co-evaporation [8,9,10], sputtering [11,12,13,14], sulfurization [15,16], atomic layer deposition [17,18,19], and chemical bath deposition [20]. This indicates that SnS has a high degree of freedom in fabrication methods. In addition, SnS exhibits a band gap of 1.3 eV [8,9,15,21,22,23] and a high absorption coefficient [21,23]. This means that SnS is suitable for use in the light-absorbing layers of solar cells.
The development history of solar cells using SnS is described briefly below. In 1994, Noguchi et al., produced the first solar cell (ITO/n-CdS/p-SnS/Ag structure) using a SnS film and reported that it exhibited photovoltaic properties [21]. In 2006, Reddy fabricated a SnS thin-film solar cell with a SnO2/p-SnS/CdS/ITO structure and showed a PCE of 1.3% [24]. In 2013, Ikuno et al. fabricated a super straight type SnS thin-film solar cell with an ITO/Zn1−xMgxO/p-SnS/Cu structure and reported a conversion efficiency of 2.1% [25]. A year later, Sinsermsuksakul et al., fabricated a substrate-type Mo/p-SnS/SnO2/Zn(O,S):N/ZnO/ITO structure cell that showed a conversion efficiency of 4.36% [18], significantly improving the performance of SnS thin-film solar cells. Despite the steady improvement in conversion efficiency, it has stagnated in the 4% range [26,27], and research on solar cell applications of SnS is at a low ebb. However, novel approaches have been recently investigated for developing SnS solar cell applications. Here, we present two state-of-the-art approaches for SnS solar cells.
The first is the preparation of SnS with n-type conduction and its application to homojunction solar cells. SnS usually exhibits p-type conduction because tin vacancies (Vsn) in the sulfur-rich state and tin sulfur antisite (SnS) in the tin-rich state behave as acceptor-type defects [13,28]. Yanagi et al. succeeded in preparing bulk SnS with n-type conduction by doping Cl [29], and Suzuki et al. reported that n-type SnS thin films can be prepared by supplying sulfur plasma while sputtering Cl-doped SnS [13]. Furthermore, Kawanishi et al. have achieved the fabrication of homojunction solar cells by employing p-type SnS thin films and n-type SnS single crystals [14].
The second is the preparation of SnS with a bandgap tunable for application to multi-junction solar cells. Regarding fabricating high-efficiency solar cells, multi-junction compound solar cells consisting of materials with different band gaps have been widely studied and were demonstrated to absorb sunlight efficiently. The light-absorbing layers of such multi-junction solar cells should exhibit adjustable bandgaps. Very recently, Kawamura et al. have succeeded in broadening the bandgap of SnS to ~2.25 eV by solid-dissolving alkaline earth metals in SnS [30].
Thus, SnS solar cells have again attracted attention in recent years due to novel approaches that take advantage of their potential capabilities.
In this study, as an approach to wide-bandgap SnS, we attempted to fabricate a (Ge, Sn)S thin-film solar cell by dissolving Ge in SnS. (Ge, Sn)S, a solid solution of SnS and germanium sulfide (GeS), shows a tunable band gap (between 1.3 and 1.6 eV) in its nanocrystals depending on the Ge concentration [31]. However, there have been very few reports on fabricating (Ge, Sn)S thin films; therefore, this topic requires further investigation. A previous study reported the fabrication of (Gex Sn1−x)S (x = 0.18, 0.27) thin films via co-evaporation and their application in thin-film solar cells [32]. Our study reports the fabrication and solar cell applications of (Gex Sn1−x)S thin films containing more Ge than previously reported. We then discuss the Ge solid solution effect by comparing it with the SnS we fabricated in the past [32].

2. Materials and Methods

Molybdenum disks (Mo, 99.95%, 3-mm thick, Furuuchi Chemical Corporation, Tokyo, Japan), tin shots (99.9999%, Furuuchi Chemical Corporation, Tokyo, Japan), germanium shots (99.999%, Furuuchi Chemical Corporation, Tokyo, Japan), sulfur chunks (S, 99.9999%, Furuuchi Chemical Corporation, Tokyo, Japan), cadmium iodide (CdI2, 99%, FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan), thiourea (CS(NH2)2, 98%, Nacalai Tesque Inc., Kyoto, Japan), ammonia water (NH3, 28 wt%, FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan), ZnO: Al (Al2O3 = 2 wt%, 3-mm thick, 99.99%, Furuuchi Chemical Corporation, Tokyo, Japan), and aluminum (Al, 99.99%, The Nilaco Corporation, Tokyo, Japan) were used as purchased.
First, approximately 0.8 μm Mo was deposited as the bottom electrode on an ultrasonically cleaned Eagle XG (Corning, Corning, NY, USA) substrate using DC sputtering (deposition time; 25 min, applied current; 1.0 A, process pressure; 0.4 Pa, Ar flow; 20 sccm). The deposition was performed under the same sputtering conditions under which the Mo thin film with a resistivity of 3.6 × 10−5 ohm/cm was obtained. Subsequently, SnS and (GexSn1−x)S thin films were deposited by the co-evaporation of Ge, Sn, and S at pressures on the order of 10−4 Pa for 3 h using independent deposition sources. Sulfur is typically supplied as S8 molecules; however, to ensure a higher supply of reactive sulfur, S8 molecules were decomposed into Sx by thermal cracking at 800 °C. Table 1 lists the manufacturing conditions in detail.
(Gex Sn1−x)S thin films were fabricated under lower substrate temperature conditions (150 °C) than those used to synthesize SnS thin films to suppress the re-evaporation of GeS (which exhibits a high vapor pressure).
The as-prepared (Gex Sn1−x)S thin films were annealed under a nitrogen atmosphere (the process pressure was atmospheric pressure, 1 × 105 Pa) in an infrared heating furnace. The temperature profile during annealing is shown in Figure 1. The fabricated thin films were analyzed using X-ray fluorescence spectroscopy (XRF, ZSX Primus IV, Rigaku Corporation, Tokyo, Japan), X-ray diffraction (XRD, MiniFlex, Rigaku Corporation, Tokyo, Japan), and Raman spectroscopy (RMP-510, JASCO Corporation, Tokyo, Japan) at an excitation wavelength of 532 nm, and scanning electron microscopy (SEM, JSM-6060LV, JEOL Ltd., Tokyo, Japan).
Subsequently, an n-type CdS buffer layer approximately 90 nm thick was deposited on the (Gex Sn1−x)S absorption layer by the CBD method at 70 °C for 20 min using a solution consisting of cadmium iodide (CdI2; 3.58 mmol L−1), ammonia water (NH3; 2.83 mol L−1), and thiourea ((NH2)2CS); 0.298 mol L−1). The fabricated (Gex Sn1−x)S /CdS thin films were annealed in air at 200 °C for 20 min [33]. Next, Al-doped zinc oxide (ZnO:Al) was deposited as a window layer by radio frequency sputtering (deposition time; 110 min, RF power; 100 W, process pressure; 0.5 Pa, Ar flow; 20 sccm). The deposition was performed under the same sputtering conditions under which the ZnO:Al thin film was obtained with a thickness of approximately 450 nm, a resistivity of 1 × 10−4 ohm/cm, and a transmittance of >80% in the visible light wavelength range 400–1000 nm. Finally, Al was deposited as the top electrode by thermal evaporation.
The final thin-film solar cells with an Eagle XG/Mo/SnS or (Gex Sn1−x)S/CdS/ZnO:Al/ Al structure were scribed into cells with an approximate size of 4.4 mm × 4.4 mm (aperture area ~0.16 cm2). The photovoltaic characteristics of the fabricated solar cells were evaluated using a solar simulator (SX-UI 500XQ, Ushio Inc., Tokyo, Japan) under AM 1.5 and 100 mW/cm2 irradiation. The external quantum efficiency (EQE) of the fabricated solar cells was evaluated using a solar cell evaluation system (SML-250 J, Bunkoukeiki Co., Ltd., Tokyo, Japan).

3. Results and Discussion

The elemental compositions of the fabricated SnS and (Gex Sn1−x)S thin films were analyzed by XRF, as summarized in Table 2. The average composition in a φ 20-mm diameter plane near the center of a 25 mm × 25 mm sample was evaluated, and a summary of the compositional ratios is given. The composition of the thin films was estimated using analysis software (ZSX Primus IV, Rigaku Co., Tokyo, Japan) based on the fundamental parameter method. The error in the composition values of Sn, Ge, and S is less than 1% in repeated measurements of the same sample. The SnS film was thinner than the (Gex Sn1−x)S film, possibly due to the re-evaporation of SnS during fabrication, which was carried out at a substrate temperature of 300 °C. The Ge/(Ge + Sn) composition ratio of the (Gex Sn1−x)S thin film was estimated to be 0.42; the thin film has been subsequently denoted as (Ge0.42 Sn0.58)S. For the (Ge0.42 Sn0.58)S thin film, annealing decreased the film thickness and increased the (Ge + Sn)/S composition ratio, although its Ge/(Ge + Sn) composition ratio remained unchanged. This could be attributed to the re-evaporation of sulfur during annealing. The S-rich composition compared to stoichiometric composition was confirmed for both SnS and (Ge0.42 Sn0.58)S. According to a previously published first-principles study, S-rich SnS is more likely to form tin vacancy (Vsn) defects at shallow energy levels than sulfur vacancy (Vs) and sulfur antisite tin (SnS) defects at deep energy levels [34]. Therefore, it is speculated that the fabricated SnS thin film had relatively fewer Vs, which can serve as recombination centers, and a thin film with excellent potential was fabricated.
Figure 2a shows the XRD patterns of the fabricated SnS and (Ge, Sn)S thin films. XRD measurements were performed using MiniFlex with a NaI scintillation counter used as a detector. This apparatus has a vertical goniometer with a radius of 150 mm and an accuracy of ±0.02°. The measurement parameters were as follows: X-ray = 30 kV/15 mA, divergence slit (DS) = variable, scattering slit (SS) = 4.2°, receiving slit (RS) = 0.3 mm, and Kβ foil filter. The measurement range was 10–90° with a step of 0.01°. The pattern of the fabricated SnS thin film was similar to that of orthorhombic SnS, while the (Ge0.42 Sn0.58)S thin film exhibited peaks at 31.27°, 32.37°, and 32.90°, as shown in Figure 2b. These peaks could be attributed to the (101), (111), and (040) planes of a (Ge, Sn)S solid solution, or to the different phases of Sn2S3, GeS, and GeS2. A plot of the estimated lattice constant of the b-axis for the most intense peak at 32.90° is shown in Figure 2c. The estimated lattice constant b, calculated using the ICDD Powder Diffraction File for SnS and GeS (PDF SnS #00-039-0354 and GeS #00-051-1168), roughly followed the Vegard rule; thus, a (Ge, Sn)S solid solution was formed in the (Ge0.42 Sn0.58)S thin film. Moreover, as shown in Figure 2a, a peak at 12.90° (possibly due to Sn2S3) was observed in the XRD spectrum of the (Ge0.42 Sn0.58)S thin film; here, the peak appeared at a higher angle than its position in the reference (Sn2S3 #00-014-0619) spectrum, probably because a part of Sn2S3 was substituted by Ge to form (Ge, Sn)2S3.
Figure 3 shows the Raman spectra of the SnS [32] and (Ge0.42 Sn0.58)S thin films. The spectral peaks of the SnS thin film could be attributed to orthorhombic SnS; peaks were observed at 95.9 cm−1 (Ag), 164.0 cm−1 (B3g), 192.0 cm−1 (Ag), and 219.5 cm−1 (Ag) (the vibrational modes are indicated in brackets) [35]. No peaks attributable to GeS and SnS2 were observed [36,37]. A higher-wavenumber peak shift was observed in the (Ge0.42 Sn0.58)S spectrum; this could be attributed to the substitution of the Sn sites in SnS by Ge. A similar shift of Raman peaks has been reported by Araki et al. in a similar sulfide solid solution, Cu2(Gex Sn1−x)S3, which is attributed to the substitution of Sn by Ge [38]. Therefore, a (Ge, Sn)S solid solution was formed in the Gex Sn1−xS system in this study, even with a Ge composition corresponding to a high Ge/(Ge + Sn) ratio of 0.42. Raman spectroscopy did not indicate any of the different phases (such as GeS2 or Sn2S3) indicated by XRD analysis; thus, it is suggested that (Ge, Sn)S predominantly existed near the surface of the (Ge0.42 Sn0.58)S thin film.
Figure 4a shows the SEM surface morphology of SnS. Block-like and plate-like crystal grains were observed on the SnS surface; these grains are commonly observed in SnS [39]. Figure 4b shows the surface morphology of (Ge0.42 Sn0.58)S. The observed grains are elongated platelets. The observed 1–5 µm grains, which are considerably larger than those of SnS, are thought to be due to the crystal growth of microcrystal grains by annealing. Sinsermsuksakul et al. reported that as the film thickness increased, the {010} plane switched its crystallographic orientation from parallel to perpendicular to the substrate [17]. This suggests that the difference in the morphology of SnS and (Ge0.42 Sn0.58)S may be partly due to the difference in film thickness. From the aforementioned XRD results, the (Ge0.42 Sn0.58)S thin film was not preferentially oriented in one direction, but the orientation of each grain was considered to be different, as peaks that can be attributed to (111) and (040) are observed in the film. They have also pointed out that the plane perpendicular to the substrate is a desirable structure for solar cells because it significantly improves the mobility in the thickness direction compared to the plane parallel to the substrate [17]. The differences in contrast between the individual grains observed in Figure 4b may be due to differences in conductivity resulting from differences in orientation.
Figure 5a and Table 3 show the J-V curves and photovoltaic characteristics of SnS and (Ge0.42 Sn0.58)S thin-film solar cells. The solar cells fabricated using SnS thin films showed an open circuit voltage (Voc) of 0.053 V, short circuit current density (Jsc) of 9.00 mA/cm2, fill factor (FF) of 0.32, and power conversion efficiency (PCE) of 0.15%. Contrarily, solar cells fabricated using the (Ge0.42 Sn0.58)S thin film showed a Voc of 0.29 V, Jsc of 6.92 mA/cm2, FF of 0.34, and PCE of 0.67%. The (Ge0.42 Sn0.58)S device showed a higher Voc than the SnS device, with similar values of FF. The efficient formation of a p-n junction interface between the n-type CdS and p-type (Ge, Sn)S possibly improved the Voc of the (Ge0.42 Sn0.58)S device; the Voc value of the device fabricated here is greater than those of previously reported solar cells based on (Ge0.27 Sn0.73)S thin films [32]. The Voc deficit Vocfef = VocEg/q (V), the difference between the band gap and the open circuit voltage estimated from the absorption edge of the EQE, which will be discussed below, was Vocdef = 1.217 V for x = 0 (Eg = 1.27 eV). For x = 0.42 (Eg = 1.42 or 1.52 eV), Vocdef = 1.13 or 1.23 V. This large Vocdef indicates that significant recombination occurs in both samples. As the Vocdef is almost the same magnitude for both samples, the change in Voc can be attributed to the increase in band gap due to the introduction of Ge. This could be due to an increase in the band gap of the optical absorption layer due to an increase in the Ge concentration and the formation of large crystal grains due to the annealing process. Furthermore, the Jsc of the solar cell based on (Ge0.42 Sn0.58)S thin films was slightly lower than that based on SnS thin films, possibly due to the presence of different phases, such as GeS2 and Sn2S3, which promoted recombination. Figure 5b shows the EQE of the fabricated SnS and (Ge0.42 Sn0.58)S thin-film solar cells. The highest EQE was recorded at about 530 nm for both devices, possibly because only the carriers near the light absorption layer (SnS or (Ge0.42 Sn0.58)S)/CdS interface were extracted; carriers that were excited deep inside the light absorption layer (near the Mo back electrode) underwent recombination due to the short diffusion length of the carriers, and could not be extracted. The energy E versus [E ln(1-EQE)]2 plots for the SnS and (Ge0.42 Sn0.58)S thin-film solar cells are shown in Figure 5c and d, respectively. The band gap of SnS was estimated to be 1.28 eV by extrapolation, similar to the previously reported value [8,9,15,21,22,23]. The (Ge0.42 Sn0.58)S spectrum shows a tail at the absorption edge, and its band gap was estimated to be in the range of 1.42–1.52 eV, indicating a non-uniform Ge concentration in (Ge, Sn)S, or absorption due to subgap defect levels.

4. Conclusions

This study reports the synthesis of (Ge0.42 Sn0.58)S thin films using co-evaporation. XRD and Raman spectroscopy confirmed that a (Ge, Sn)S solid solution was formed in the (Ge0.42 Sn0.58)S thin film. The open circuit voltage (Voc), short circuit current density (Jsc), FF, and PCE of solar cells fabricated using the (Ge0.42 Sn0.58)S thin film were 0.29 V, 6.92 mA/cm2, 0.34, and 0.67%, respectively. Moreover, the bandgap of the (Ge0.42 Sn0.58)S thin-film solar cell was estimated to be 1.42–1.52 eV. In summary, in this study, we report the synthesis of a (Ge0.42 Sn0.58)S thin film with a Ge content of 0.42—confirmed to be a (Ge, Sn)S solid solution—and its application in solar cells. Among the (Ge, Sn)S solar cells reported thus far, it was clarified that solar cells can be realized even in a composition region with a relatively high Ge concentration. However, the photoelectric conversion characteristics are still inadequate, and improvements such as the suppression of thin-film crystal defects and the formation of appropriate heterojunction interfaces are required to improve the efficiency further. In addition, more detailed research on (Ge, Sn)S is required, such as the fabrication of (Ge, Sn)S thin films with even higher Ge concentrations and the realization of (Ge, Sn)S thin-film solar cells.

Author Contributions

Investigation and writing—original draft preparation, D.M.; writing—review, editing, and supervision, H.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially supported by JSPS KAKENHI (Grant Number JP 19H02663).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Green, M.A.; Dunlop, E.D.; Yoshita, M.; Kopidakis, N.; Bothe, K.; Siefer, G.; Hao, X. Solar cell efficiency tables (version 62). Prog. Photovolt. Res. Appl. 2023, 31, 651–663. [Google Scholar] [CrossRef]
  2. Nair, P.K.; Garcia-Angelmo, A.R.; Nair, M.T.S. Cubic and orthorhombic SnS thin-film absorbers for tin sulfide solar cells. Phys. Status Solidi A 2016, 213, 170–177. [Google Scholar] [CrossRef]
  3. Garcia-Angelmo, A.R.; Romano-Trujillo, R.; Campos-Alvarez, J.; Gomez-Daza, O.; Nair, M.T.S.; Nair, P.K. Thin film solar cell of SnS absorber with cubic crystalline structure. Phys. Status Solidi A 2015, 212, 2332–2340. [Google Scholar] [CrossRef]
  4. Ahmet, I.Y.; Guc, M.; Sánchez, Y.; Neuschitzer, M.; Izquierdo-Roca, V.; Saucedo, E.; Johnson, A.L. Evaluation of AA-CVD deposited phase pure polymorphs of SnS for thin films solar cells. RSC Adv. 2019, 9, 14899. [Google Scholar] [CrossRef] [PubMed]
  5. Norton, K.J.; Alam, F.; Lewis, D.J. A Review of the Synthesis, Properties, and Applications of Bulk and Two-Dimensional Tin (II) Sulfide (SnS). Appl. Sci. 2021, 11, 2062. [Google Scholar] [CrossRef]
  6. Steinmann, V.; Jaramillo, R.; Hartman, K.; Chakraborty, R.; Brandt, R.E.; Poindexter, J.R.; Lee, Y.S.; Sun, L.; Polizzotti, A.; Park, H.H.; et al. 3.88% Efficient Tin Sulfide Solar Cells Using Congruent Thermal Evaporation. Adv. Mater. 2014, 26, 7488–7492. [Google Scholar] [CrossRef] [PubMed]
  7. Kawano, Y.; Chantana, J.; Minemoto, T. Impact of growth temperature on the properties of SnS film prepared by thermal evaporation and its photovoltaic performance. Curr. Appl. Phys. 2015, 15, 897–901. [Google Scholar] [CrossRef]
  8. Yago, A.; Kibishi, T.; Akaki, Y.; Nakamura, S.; Oomae, H.; Katagiri, H.; Araki, H. Influence of Sn/S composition ratio on SnS thin-film solar cells produced via co-evaporation method. Jpn. J. Appl. Phys. 2018, 57, 02CE08. [Google Scholar] [CrossRef]
  9. Yago, A.; Sasagawa, S.; Akaki, Y.; Nakamura, S.; Oomae, H.; Katagiri, H.; Araki, H. Comparison of buffer layers on SnS thin-film solar cells prepared by co-evaporation. Phys. Status Solidi C 2017, 14, 1600194. [Google Scholar] [CrossRef]
  10. Cifuentes, C.; Botero, M.; Romero, E.; Calderón, C.; Gordillo, G. Optical and structural studies on SnS films grown by co-evaporation. Braz. J. Phys. 2006, 36, 1046–1049. [Google Scholar] [CrossRef]
  11. Arepalli, V.K.; Shin, Y.; Kim, J. Photovoltaic behavior of the room temperature grown RF-Sputtered SnS thin films. Opt. Mater. 2019, 88, 594–600. [Google Scholar] [CrossRef]
  12. Zhao, L.; Di, Y.; Yan, C.; Liu, F.; Cheng, Z.; Jiang, L.; Hao, X.; Lai, Y.; Li, J. In situ growth of SnS absorbing layer by reactive sputtering for thin film solar cells. RSC Adv. 2016, 6, 4108–4115. [Google Scholar] [CrossRef]
  13. Suzuki, I.; Kawanishi, S.; Bauers, S.R.; Zakutayev, A.; Lin, Z.; Tsukuda, S.; Shibata, H.; Kim, M.; Yanagi, H.; Omata, T. n-type electrical conduction in SnS thin films. Phys. Rev. Mater. 2021, 5, 125405. [Google Scholar] [CrossRef]
  14. Kawanishi, S.; Suzuki, I.; Bauers, S.R.; Zakutayev, A.; Shibata, H.; Yanagi, H.; Omata, T. SnS Homojunction Solar Cell with n-Type Single Crystal and p-Type Thin Film. Sol. RRL 2021, 5, 2000708. [Google Scholar] [CrossRef]
  15. Sugiyama, M.; Reddy, K.T.R.; Revathi, R.; Shimamoto, Y.; Murata, Y. Band offset of SnS solar cell structure measured by X-ray photoelectron spectroscopy. Thin Solid Films 2011, 519, 7429–7431. [Google Scholar] [CrossRef]
  16. Minemura, T.; Miyauchi, K.; Noguchi, K.; Ohtsuka, K.; Nakanishi, H.; Sugiyama, M. Preparation of SnS films by low-temperature sulfurization. Phys. Status Solidi C 2009, 6, 1221–1224. [Google Scholar] [CrossRef]
  17. Sinsermsuksakul, P.; Heo, J.; Noh, W.; Hock, A.S.; Gordon, R.G. Atomic Layer Deposition of Tin Monosulfide Thin Films. Adv. Energy Mater. 2011, 1, 1116–1125. [Google Scholar] [CrossRef]
  18. Sinsermsuksakul, P.; Sun, L.; Lee, S.W.; Park, H.H.; Kim, S.B.; Yang, C.; Gordon, R.G. Overcoming Efficiency Limitations of SnS-Based Solar Cells. Adv. Energy Mater. 2014, 4, 1400496. [Google Scholar] [CrossRef]
  19. Choi, Y.; Park, H.; Lee, N.; Kim, B.; Lee, J.; Lee, G.; Jeon, H. Deposition of the tin sulfide thin films using ALD and a vacuum annealing process for tuning the phase transition. J. Alloys Compd. 2022, 896, 162806. [Google Scholar] [CrossRef]
  20. Higareda-Sánchez, A.; Mis-Fernández, R.; Rimmaudo, I.; Camacho-Espinosa, E.; Peña, J.L. Evaluation of pH and deposition mechanisms effect on tin sulfide thin films deposited by chemical bath deposition. Superlattices Microstruct. 2021, 151, 106831. [Google Scholar] [CrossRef]
  21. Araki, H.; Abe, M.; Yago, A.; Katagiri, H.; Akaki, Y. Preparation of SnS Thin Films Using Close Space Sublimation. J. Nanoelectron. Optoelectron. 2019, 12, 920–924. [Google Scholar] [CrossRef]
  22. Lin, D.-Y.; Hsu, H.-P.; Liu, K.-H.; Wu, P.-H.; Shih, Y.-T.; Wu, Y.-F.; Wang, Y.-P.; Lin, C.-F. Enhanced Optical Response of SnS/SnS2 Layered Heterostructure. Sensors 2023, 23, 4976. [Google Scholar] [CrossRef]
  23. Noguchi, H.; Setiyadi, A.; Tanamura, H.; Nagatomo, T.; Omoto, O. Characterization of vacuum-evaporated tin sulfide film for solar cell materials. Sol. Energy Mater. Sol. Cells 1994, 35, 325–331. [Google Scholar] [CrossRef]
  24. Reddy, K.T.R.; Reddy, N.K.; Miles, R.W. Photovoltaic properties of SnS based solar cells. Sol. Energy Mater. Sol. Cells 2006, 90, 3041–3046. [Google Scholar] [CrossRef]
  25. Ikuno, T.; Suzuki, R.; Kitazumi, K.; Takahashi, N.; Kato, N.; Higuchi, K. SnS thin film solar cells with Zn1−xMgxO buffer layers. Appl. Phys. Lett. 2013, 102, 193901. [Google Scholar] [CrossRef]
  26. Cho, J.Y.; Kim, S.Y.; Nandi, R.; Jang, J.; Yun, H.-S.; Enkhbayar, E.; Kim, J.H.; Lee, D.-K.; Chung, C.-H.; Kim, J.H.; et al. Achieving over 4% efficiency for SnS/CdS thin-film solar cells by improving the heterojunction interface quality. J. Mater. Chem. A 2020, 8, 20658–20665. [Google Scholar] [CrossRef]
  27. Yun, H.-S.; Park, B.; Choi, Y.C.; Im, J.; Shin, T.J.; Seok, S. Efficient Nanostructured TiO2/SnS Heterojunction. Sol. Cells Adv. Energy Mater. 2019, 9, 1901343. [Google Scholar] [CrossRef]
  28. Andrade-Arvizua, J.A.; García-Sánchez, M.F.; Courel-Piedrahitaa, M.; Pulgarín-Agudeloa, F.; Santiago-Jaimes, E.; Valencia-Resendiz, E.; Arce-Plazac, A.; Vigil-Galána, O. Suited growth parameters inducing type of conductivity conversions on chemical spray pyrolysis synthesized SnS thin films. J. Anal. Appl. Pyrolysis 2016, 121, 347–359. [Google Scholar] [CrossRef]
  29. Yanagi, H.; Iguchi, Y.; Sugiyama, T.; Kamiya, T.; Hosono, H. N-type conduction in SnS by anion substitution with Cl. Appl. Phys. Express 2016, 9, 051201. [Google Scholar] [CrossRef]
  30. Kawamura, F.; Song, Y.; Murata, H.; Tampo, H.; Nagai, T.; Koida, T.; Imura, M.; Yamada, N. Tunability of the bandgap of SnS by variation of the cell volume by alloying with A.E. elements. Sci. Rep. 2022, 12, 7434. [Google Scholar] [CrossRef]
  31. Im, H.Y.; Myung, Y.; Park, K.; Jung, C.S.; Lim, Y.R.; Jang, D.M.; Park, J. Ternary alloy nanocrystals of tin and germanium chalcogenides. RSC Adv. 2014, 4, 15695–15701. [Google Scholar] [CrossRef]
  32. Motai, D.; Tasaki, T.; Araki, H. Fabrication of solar cell using Ge-Sn-S thin film prepared by co-evaporation. Jpn. J. Appl. Phys. 2023, 62, SK1037. [Google Scholar] [CrossRef]
  33. Motai, D.; Ohashi, R.; Araki, H. Effect of annealing temperature on p–n junction formation in Cu2SnS3 thin-film solar cells fabricated via the co-evaporation of elemental precursors. Jpn. J. Appl. Phys. 2022, 61, SB1043. [Google Scholar] [CrossRef]
  34. Malone, B.D.; Gali, A.; Kaxiras, E. First principles study of point defects in SnS. Phys. Chem. Chem. Phys. 2014, 16, 26176–26183. [Google Scholar] [CrossRef]
  35. Li, M.; Wu, Y.; Li, T.; Chen, Y.; Ding, H.; Lin, Y.; Pan, N.; Wang, X. Revealing anisotropy and thickness dependence of Raman spectra for SnS flakes. RSC Adv. 2017, 7, 48759–48765. [Google Scholar] [CrossRef]
  36. Tan, D.; Lim, H.E.; Wang, F.; Mohamed, N.B.; Mouri, S.; Zhang, W.; Miyauchi, Y.; Ohfuchi, M.; Matsuda, K. Anisotropic optical and electronic properties of two-dimensional layered germanium sulfide. Nano Res. 2017, 10, 546–555. [Google Scholar] [CrossRef]
  37. Stetsenko, M.О.; Voznyi, A.A.; Kosyak, V.V.; Rudenko, S.P.; Maksimenko, L.S.; Serdega, B.K.; Opanasuk, A.S. Plasmonic Effects in Tin Disulfide Nanostructured Thin Films Obtained by the Close-Spaced Vacuum Sublimation. Plasmonics 2016, 12, 1213–1220. [Google Scholar] [CrossRef]
  38. Araki, H.; Yamano, M.; Nishida, G.; Takeuchi, A.; Aihara, N.; Tanaka, K. Synthesis and characterization of Cu2Sn1−xGexS3. Phys. Stat. Sol. 2014, 14, 1600199. [Google Scholar] [CrossRef]
  39. Banai, R.E.; Horn, M.W.; Browson, J.R.S. A review of tin (II) monosulfide and its potential as a photovoltaic absorber. Sol. Energy Mater. Sol. Cells 2016, 150, 112–129. [Google Scholar] [CrossRef]
Figure 1. Temperature profile during the annealing of (Gex Sn1−x)S thin films.
Figure 1. Temperature profile during the annealing of (Gex Sn1−x)S thin films.
Materials 17 00692 g001
Figure 2. (a) XRD patterns of the fabricated SnS [32] and (Ge0.42 Sn0.58)S thin films. The relevant PDF cards are shown at the bottom of the figure as a reference. (b) Magnified views of the XRD patterns are shown in the range of 30–35° (Red line: SnS thin film, blue line: (Ge0.42 Sn0.58)S thin film). (c) Dependence of the b-axis lattice constant on Ge concentration. The lattice constant b is calculated from the peak assigned to the (040) plane.
Figure 2. (a) XRD patterns of the fabricated SnS [32] and (Ge0.42 Sn0.58)S thin films. The relevant PDF cards are shown at the bottom of the figure as a reference. (b) Magnified views of the XRD patterns are shown in the range of 30–35° (Red line: SnS thin film, blue line: (Ge0.42 Sn0.58)S thin film). (c) Dependence of the b-axis lattice constant on Ge concentration. The lattice constant b is calculated from the peak assigned to the (040) plane.
Materials 17 00692 g002aMaterials 17 00692 g002b
Figure 3. Raman spectra of the fabricated SnS [32] and (Ge0.42 Sn0.58)S (red and blue lines, respectively) thin films. The reference Raman spectra of SnS [35], GeS [36], and SnS2 [37] are shown by black lines at the bottom of the figure.
Figure 3. Raman spectra of the fabricated SnS [32] and (Ge0.42 Sn0.58)S (red and blue lines, respectively) thin films. The reference Raman spectra of SnS [35], GeS [36], and SnS2 [37] are shown by black lines at the bottom of the figure.
Materials 17 00692 g003
Figure 4. Surface morphology of (a) SEM images of surface morphology at different locations on the same sample as SnS in Ref. [32] and (b) (Ge0.42 Sn0.58)S (annealed), observed using SEM. A white scale bar is shown at the bottom of the figure.
Figure 4. Surface morphology of (a) SEM images of surface morphology at different locations on the same sample as SnS in Ref. [32] and (b) (Ge0.42 Sn0.58)S (annealed), observed using SEM. A white scale bar is shown at the bottom of the figure.
Materials 17 00692 g004
Figure 5. J-V curves and EQEs for the fabricated solar cells; the red and blue lines correspond to the SnS and (Ge0.42 Sn0.58)S thin-film solar cells, respectively. (a) J-V curves of solar cells fabricated with the SnS and (Ge0.42 Sn0.58)S thin films. (b) The EQEs of solar cells fabricated with the SnS and (Ge0.42 Sn0.58)S thin films. (c) Energy (E) versus [E ln(1-EQE)]2 plot and extrapolated line (black line) in the EQE plot of the SnS thin-film solar cell. (d) E versus [E ln(1-EQE)]2 plots and extrapolation lines for (Ge0.42 Sn0.58)S thin-film solar cells, similar to (c).
Figure 5. J-V curves and EQEs for the fabricated solar cells; the red and blue lines correspond to the SnS and (Ge0.42 Sn0.58)S thin-film solar cells, respectively. (a) J-V curves of solar cells fabricated with the SnS and (Ge0.42 Sn0.58)S thin films. (b) The EQEs of solar cells fabricated with the SnS and (Ge0.42 Sn0.58)S thin films. (c) Energy (E) versus [E ln(1-EQE)]2 plot and extrapolated line (black line) in the EQE plot of the SnS thin-film solar cell. (d) E versus [E ln(1-EQE)]2 plots and extrapolation lines for (Ge0.42 Sn0.58)S thin-film solar cells, similar to (c).
Materials 17 00692 g005
Table 1. Film fabrication conditions for the co-evaporation process.
Table 1. Film fabrication conditions for the co-evaporation process.
SnS [32](Gex Sn1−x)S
Deposition time [hours]33
Ge cell temperature [°C]-1200
Sn cell temperature [°C]10151010
S temperature [°C]150150
Substrate temperature [°C]300150
S-Valve opening31.5
Table 2. Film thicknesses and constituent-element compositions estimated by XRF.
Table 2. Film thicknesses and constituent-element compositions estimated by XRF.
Thickness [µm]Ge/(Ge + Sn)(Ge + Sn)/S
SnS [32]0.68700.98
(Gex Sn1−x)S (As deposited)1.0870.42 0.95
(Gex Sn1−x)S (Annealed)1.0330.420.98
Table 3. Photovoltaic characteristics of solar cells fabricated using the as-prepared SnS and (Ge0.42 Sn0.58)S thin films.
Table 3. Photovoltaic characteristics of solar cells fabricated using the as-prepared SnS and (Ge0.42 Sn0.58)S thin films.
Area
[cm2]
Voc
[V]
Jsc [mA/cm2]FFPCE
[%]
SnS [32]0.15770.0539.00 0.320.15
(Ge0.42 Sn0.58)S0.15140.296.920.340.67
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Motai, D.; Araki, H. Fabrication of (Ge0.42Sn0.58)S Thin Films via Co-Evaporation and Their Solar Cell Applications. Materials 2024, 17, 692. https://doi.org/10.3390/ma17030692

AMA Style

Motai D, Araki H. Fabrication of (Ge0.42Sn0.58)S Thin Films via Co-Evaporation and Their Solar Cell Applications. Materials. 2024; 17(3):692. https://doi.org/10.3390/ma17030692

Chicago/Turabian Style

Motai, Daiki, and Hideaki Araki. 2024. "Fabrication of (Ge0.42Sn0.58)S Thin Films via Co-Evaporation and Their Solar Cell Applications" Materials 17, no. 3: 692. https://doi.org/10.3390/ma17030692

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

Article Metrics

Back to TopTop