Next Article in Journal
Preparation of High-Performance Metal-Free UV/Near Infrared-Shielding Films for Human Skin Protection
Next Article in Special Issue
Ultra-Flexible Organic Solar Cell Based on Indium-Zinc-Tin Oxide Transparent Electrode for Power Source of Wearable Devices
Previous Article in Journal
Theoretical Investigation of Proton Diffusion in Dion–Jacobson Layered Perovskite RbBiNb2O7
Previous Article in Special Issue
Efficient Nanocrystal Photovoltaics via Blade Coating Active Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fundamental Aspects and Comprehensive Review on Physical Properties of Chemically Grown Tin-Based Binary Sulfides

by
Sreedevi Gedi
1,
Vasudeva Reddy Minnam Reddy
1,*,
Tulasi Ramakrishna Reddy Kotte
2,
Chinho Park
1 and
Woo Kyoung Kim
1,*
1
School of Chemical Engineering, Yeungnam University, 280 Daehak-ro, Gyeongsan 38541, Korea
2
Department of Physics, Sri Venkateswasra University, Tirupati 517 502, India
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(8), 1955; https://doi.org/10.3390/nano11081955
Submission received: 20 June 2021 / Revised: 21 July 2021 / Accepted: 26 July 2021 / Published: 29 July 2021
(This article belongs to the Special Issue Advances in Nanomaterials for Photovoltaic Applications)

Abstract

:
The rapid research progress in tin-based binary sulfides (SnxSy = o-SnS, c-SnS, SnS2, and Sn2S3) by the solution process has opened a new path not only for photovoltaics to generate clean energy at ultra-low costs but also for photocatalytic and thermoelectric applications. Fascinated by their prosperous developments, a fundamental understanding of the SnxSy thin film growth with respect to the deposition parameters is necessary to enhance the film quality and device performance. Therefore, the present review article initially delivers all-inclusive information such as structural characteristics, optical characteristics, and electrical characteristics of SnxSy. Next, an overview of the chemical bath deposition of SnxSy thin films and the influence of each deposition parameter on the growth and physical properties of SnxSy are interestingly outlined.

Graphical Abstract

1. Introduction

To make a significant contribution to the energy needs of society with low production cost, thin film photovoltaic (TFPV) technology has been developed. Currently, the CdTe/CdS and Cu(In,Ga)Se2(CIGS)/CdS heterojunction TFPV technologies have received worldwide attention because these solar cells achieved record efficiencies of 22.1% [1] and 23.35% [2], respectively. However, their wider impact is hindered due to major concerns raised on the presence of harmful elements (Cd and Se) and scarcity of constituent elements (Te, In and Ga). In view of that, considerable efforts have been made to develop environmentally friendly absorbers and buffers that are free from the aforementioned toxic and inadequacy elements. Along this path, the tin-based binary sulfides (SnxSy) such as tin monosulfide (orthorhombic (ORT)-SnS and cubic (CUB)-SnS), tin disulfide (SnS2), and tin sesquisulfide (Sn2S3) have drawn much attention because these are abundant, inexpensive, and nontoxic [3]. According to the merits of SnxSy (see Table 1), the o-SnS, c-SnS, and Sn2S3 are strongly expected as potential and alternative absorbers to the conventional CdTe and CIGS, and SnS2 is expected as an appropriate and alternative buffer to the regular CdS. Furthermore, their simple composition and promising physical properties made them suitable for other applications such as photocatalytic, thermoelectric, etc. (see Figure 1). Among SnxSy, o-SnS, SnS2, and Sn2S3 occur naturally, whereas c-SnS was synthesized in the laboratory. The historical information and the applications of SnxSy available in the literature are presented in Table 2.
The deposition of SnxSy in thin film form became prominent owing to their wide applications. The selection of deposition technique along with the growth conditions is critical because the properties of SnxSy thin films are susceptible to their growth method. SnxSy films should be prepared by low-cost techniques such as solution processes to further reduce the production cost of TFPV devices. The preparation of SnxSy thin films via chemical methods, especially by chemical bath deposition(CBD), includes a slightly low cost and is, in fact, unchallenging on the preparation side. In addition, CBD provides several experimental flexibilities such as non-vacuum thin-film deposition, a wide selection of various substrates, easy doping of elements, and room temperature film growth, which are suitable for the large-scale fabrication of flexible devices for industrial applications.
The deposition of SnxSy thin films using CBD is relatively new, and their process–property relationships must be understood for the desired application. Further, the formation of single-phase o-SnS, c-SnS, SnS2, and Sn2S3 thin films in CBD is highly dependent on preparative conditions. In addition, identification and separation of the o-SnS, c-SnS, SnS2, and Sn2S3 phases are also critical criteria. However, there is a lack of comprehensive studies on the optimization of growth parameters until now. Therefore, extensive research studies on the growth, deposition mechanism, and preparative parameters that affect phase separation and the physical properties of SnxSy thin films are crucial to a successful device design in the production of clean energy.
In this scenario, the present article provides an overview of the bulk properties of SnxSy and a comprehensive review of the deposition, growth mechanism, and effect of growth parameters on the physical properties of SnxSy thin films. According to the authors’ knowledge, this is the first review of SnxSy thin films by CBD.

2. Physical Properties of o-SnS, c-SnS, SnS2, and Sn2S3

The physical properties such as structural, optical, and electrical properties of SnxSy can significantly influence the device’s performance. The crystal structure of a material can influence its optical and electrical properties, which can affect the material-related device performance [92]. Understanding and obtaining knowledge on the electronic band structure and optical characteristics of SnxSy is essential before using them for device applications because the main optical parameter, band gap energy (Eg), is very sensitive to the crystal structure and defects, which directly influences the performance of a PV device. Electrical characteristics such as conduction type, resistivity, carrier concentration, and mobility of SnxSy play a key role in achieving high-performance photovoltaic devices. These electrical properties critically depend on the formation of defects in SnxSy. Therefore, a good understanding of the physical properties is required for the development of effective devices.

2.1. Crystal Structure and Structural Characteristics

In o-SnS, c-SnS, SnS2, and Sn2S3, Sn exhibits ‘2+’ (divalent Sn(II)) or/and ‘4+’ (tetravalent Sn(IV)) oxidation states (Table 3). The capability of Sn to elect different oxidation states is the origin of structural diversity/different structures (Figure 2a) of resulting compounds [93]. o-SnS (α-SnS, space group, Pnma- D 16 2 h ) and c-SnS (π-SnS, space group, P213) are the two polymorphic forms of SnS, resulting from distortions in the crystal lattice depending on growth conditions. The o-SnS consists of a double layer of Sn and S atoms (two-dimensional SnS sheets) with zigzagged chains in which each Sn atom bonds to two S atoms in the b–c plane of the layer with a bond length of 2.671 Å and one additional S atom at a short bond length of 2.633 Å perpendicular to the plane (along a-axis) in the layer stack. The interlayer bond length of Sn-Sn atoms is 3.48 Å, and the distance between the layer stacks is 2.79 Å. The lone pair of 5s2 in the Sn atom occupies the fourth coordination site and weakens interaction along the b-axis. In the case of c-SnS, each Sn atom bonds with three nearest S atoms at 2.7 Å and forms a trigonal pyramidal environment, with the Sn-S bond at the trigonal base and the 5s2 lone pair pointing toward the apex. The stereo chemical activity of Sn(II) 5s2 lone pair creates a highly distorted internal structure of c-SnS. The local coordination in c-SnS is similar to o-SnS; however, a three-dimensional network is formed by a covalent bond [60]. Additionally, SnS also exhibits some of other polymorphs [94] such as, β-SnS (formed at T > 880 K), γ-SnS [95], δ-SnS [96,97], RS-SnS (rock salt-SnS, Fm 3 ¯ m) [98]. One key point is that c-SnS and RS-SnS belong to the cubic crystal system. Further, c-SnS is a simple cubic lattice type, and RS-SnS is a face-centered cubic lattice (important note: the structure of c-SnS was incorrectly assigned as ZB-SnS previously in the literature. To avoid confusion in the literature, the readers should replace ZB-SnS with c-SnS in previous literature). More details can be found in the literature clarifying these assignments [28,93,99,100]. Furthermore, the weak interactions between distorted lone pair of 5s2 in Sn and neighboring S form the metastable SnS crystals or polymorphs of SnS [101].
Next, the SnS2 adopts a layered hexagonal structure with P 3 ¯ ml space group, similar to the structure of the CdI2 system. In this structure, the layers are arranged in the b–c plane, which is perpendicular to the a-axis, and each layer is composed of S, Sn, and S atomic parallel monolayers. Each Sn atom forms bonds in an octahedral environment with six S atoms, similar in rutile SnO2 structure [93]. It has a symmetric edge-sharing Sn(IV)S6 octahedral with 2D planes that are separated by weak van der Waals interaction between 3.6 Å distant S atoms [61]. Finally, the Sn2S3 exhibits an orthorhombic crystal structure similar to o-SnS with the same space group of Pnma. It contains tetravalent (4+) and divalent (2+) Sn atoms in equal proportions, and they form Sn(IV)S6 octahedral 1D chains covered by Sn(II) tetrahedral [51]. The lone pair in Sn(II) occupies one coordination site as in the ground state form of o-SnS. These lone pairs are responsible for weak interchain interactions. The optimized theoretical lattice parameters along with experimental values for all these SnxSy phases are presented in Table 3. The structural characterization of SnxSy thin films is generally performed by X-ray diffraction (XRD). The standard XRD patterns of o-SnS, c-SnS, SnS2, and Sn2S3 (Figure 2b) showed the highest intensity for peaks located at 2θ of: 31.53°, 26.63°/30.84°, 15.03°, and 21.49°, arising from diffraction from (111), (222)/(400), (001), and (130) planes, respectively [25,28,102,103,104].
The o-SnS, c-SnS, SnS2, and Sn2S3 thin films have similarities in XRD patterns (Figure 3a). Thus, clear differentiation of one phase to another is difficult by using diffraction analysis alone. In this respect, it is preferable to identify these phases in pure form using a complementary method such as Raman spectroscopy because the Raman spectrum is sensitive to mainly crystal quality, structural symmetry, and strength of the chemical bond between atoms [105]. The o-SnS has 21 optical vibrational modes with the irreducible representation of Γ = 4Ag + 2B1g + 4B2g + 2B3g + 2Au + 4B1u + 2B2u + 4B3u [106,107]. In these phonons, two are inactive (2Au), seven are infrared-active (3B1u, 3B3u, and 1B2u), and twelve are Raman-active (4Ag, 2B1g, 4B2g, and 2B3g). In the case of c-SnS, there are 189 optic branches, and they can be reduced to 3 in the form of Γ = 16A + 16E + 47T [51]. Next, the SnS2 has six vibrational modes with the irreducible representation, Γ = A1g + Eg + 2A2u + 2 Eu [108]. The six optic modes are divided into two Raman-active modes (A1g and Eg), two infrared-active (A2u and Eu), and two acoustic modes (A2u and Eu). Additionally, the Sn2S3 has 57 optic modes with reduced form of Γ = 10Ag + 5Au + 5B1g + 9B1u + 10B2g + 4B2u + 5B3g + 9B3u [51,109]. The simulated Raman spectra [51] (Figure 3b) clearly showed the significant differences in frequencies and spectral intensities because the o-SnS, c-SnS, SnS2, and Sn2S3 phases have differences in structure and bonding. The Raman spectrum of the o-SnS showed three prominent peaks at 160 cm−1 (narrow mode, B2g), 189 cm−1 (highest intensity mode, Ag), and 220 cm−1 (narrow mode, Ag). The spectrum also showed a weak Ag mode at approximately 92 cm−1, which has a narrow line width up to room temperature. In the case of c-SnS, there are three strong A phonon modes at 174 cm−1, 187 cm−1, and 202 cm−1 and two prominent E modes at 166 cm−1 and 183 cm−1 along with a group of weak modes in between the ranges of 50–125 cm−1 and 200–250 cm−1. Next, the SnS2 showed a single and strong mode at 305 cm−1 (Ag), which has a constant line width with the temperature. Furthermore, the Sn2S3 showed a significantly high-intensity Ag mode at 291 cm−1 and a moderate-intensity Ag mode at 300 cm−1 with a narrow line width. Its spectrum also showed weak modes at 182 cm−1, 210 cm−1, 226 cm−1, 244 cm−1, and 252 cm−1, which are observed at high temperatures. Notably, the Raman mode of Sn2S3 (307 cm−1) overlaps marginally with the active mode of SnS2 (310 cm−1). However, the phase can be easily identified based on the band width of modes. Sn2S3 has a band width that is significantly greater than that of SnS2 (Figure 3b). From the experimental Raman spectra (Figure 3c), o-SnS, c-SnS, SnS2, and Sn2S3 films showed Raman active modes at 93 cm−1, 161 cm−1, 192 cm−1, and 218 cm−1 [110]; 59 cm−1, 71 cm−1, 90 cm−1, 112 cm−1, 123 cm−1, 176 cm−1, 192 cm−1, 202 cm−1, and 202 cm−1 [111]; 224 cm−1 and 310 cm−1 [112]; and 61 cm−1, 91 cm−1, 179 cm−1, 220 cm−1, and 307 cm−1 [113], respectively, which matched well with theoretically calculated data. The Raman results suggested that o-SnS, c-SnS, SnS2, and Sn2S3 phases have distinct modes. Moreover, S-rich impurity phases can easily be found in SnS due to the sharp Raman mode of SnS2 (310 cm−1).

2.2. Electronic Band Structure and Optical Characteristics

According to the electronic band structures of SnxSy (Figure 3d), the valence band maxima (VBM) of o-SnS, c-SnS, and Sn2S3 are formed mostly of S 3p and Sn 5s hybrid states with a tiny contribution from Sn 5p states, whereas SnS2 is primarily composed of S 3p orbitals. The conduction band minimum (CBM) of SnS2 and Sn2S3 is formed by the Sn 5s bands, whereas SnS is mainly composed of Sn 5p orbitals. The VBM of SnS2 is lower than those of o-SnS, c-SnS, and Sn2S3; however, the CBMs of all o-SnS, c-SnS, SnS2, and Sn2S3 are almost aligned. The partial hybridizations with S 3s and Sn 5p states result in SnS polymorphs due to the change in density of state.
On the other hand, the band-edge positions deviated from the special points in the reciprocal space except for the CBM of Sn2S3. However, they fall between the Brillouin zone center and the zone boundaries [64,114]. From the ab initio band-structure calculations and Kohn–Sham density-functional theory, o-SnS and c-SnS exhibited the indirect and direct energy gaps of 1.6 eV and 1.8 eV; and 1.72 eV and 1.74 eV, respectively, which is due to an inherent error in calculating band structure [58]. The energy difference between direct and indirect band gaps is small; thus, the change in the nature of the band gap could be due to the effect of temperature through thermal expansion and electron-phonon coupling [64]. According to the band-structure calculations from the Hartree−Fock exchange HSE06 functional technique, o-SnS, SnS2, and Sn2S3 showed indirect energy gaps of 1.11 eV, 2.24 eV, and 1.09 eV, respectively [115]. The optical characterization of SnxSy thin films is generally studied by a UV-Vis-NIR spectrometer. Although theoretical calculations showed the indirect band gap energy of SnxSy, most of the experimental studies (Figure 3e) proved that SnxSy thin films have direct band gap energies, with the following ranges (Table 4 and Table 5): 1.16–1.79 eV; 1.64–1.75 eV; 2.04–3.30 eV; 0.95–2.03 eV; for o-SnS, c-SnS, SnS2, and Sn2S3, respectively. The band gap energies of these phases depend on various factors such as strain, sulfur impurities, and Sn vacancies [116,117,118,119,120,121,122].

2.3. Conduction Type and Electrical Characteristics

In SnxSy, three types of defects, namely, (i) Sn and S vacancies (VSn and VS), (ii) Sn and S interstitials (Sni and Si), and (iii) Sn on S antisites (SnS) and S on Sn antisites (SSn) are commonly formed, as shown in Figure 4a. According to the defect energy concepts (Figure 4b), the formation energy of vacancies (VSn(II) or Sn(IV), Vs) depends on the coordination number, i.e., it generally increases with increasing coordination number. Thus, VSn(II) has lower formation energy compared to VSn(IV) because the coordination number is three for Sn(II) and six for Sn(IV). As a result, VSn(II) becomes a major defect that acts as an acceptor and contributes to the p-type conducting nature to SnS. In SnS, the primary defects are VSn and VS, whereas Sni and Si have higher energies. SnS exhibits the p-type at the Sn-poor condition, whereas the n-type at the Sn-rich condition. The defect-formation energies in SnS2 differ from those in SnS. The major defects are VS, Si, and SSn(II), which are inert to carrier generation. The defect-formation energies in Sn2S3 are typically interpreted as a mixture of those in SnS and SnS2. On the other hand, the formation energy of interstitials (Sni, Si) associates with the gap of interlayer free spaces, and that gap follows the notation of SnS > Sn2S3 > SnS2. Sni always prefers to locate at the center of the gaps, whereas Si likes to make a covalent bond with neighboring S atoms. The tin interstitial, Sni in both SnS2 and Sn2S3, has lower formation energy compared to SnS, and it acts as a deep donor in SnS2 and Sn2S3 and contributes to an n-type conductivity. All the antisites in SnxSy have higher formation energies because they are correlated with chemical bonds. Therefore, these defects do not play a major role in the conduction type of SnxSy phases. The electrical characterization of SnxSy thin films is generally performed by the popular van der Pauw–Hall method.
From the reported electrical parameters of SnxSy films (Table 3), the o-SnS has a hole density in the order of 1011–1018 cm−3, hole mobility in the range of 4–500 cm2 V−1 s−1, and electrical resistivity in the range of 13–105 Ω cm. In contrast, the c-SnS has a hole density in the order of 1011–1018 cm−3, hole mobility in the range of 10−2–78 cm2 V−1 s−1, and electrical resistivity in the range of 70–107 Ω cm. In the case of the SnS2, it has a carrier concentration of the order of 1013–1017 cm−3, electron mobility in the range of 15–52 cm2 V−1 s−1, and electrical resistivity in the range of 1.11–107 Ω cm, whereas the Sn2S3 has a carrier density in the order of 1014–1016 cm−3 and resistivity in the range of 0.4–105 Ω cm, and a very little information related to Sn2S3 carrier mobility value is available in the literature. The reported variation in electrical parameters is expected due to the differences in the growth process and chemical composition.

3. Influence of Deposition Parameters on SnxSy Thin Film Growth and Properties

In CBD, the selection of tin source precursor, sulfur source precursor, complexing agent, and their concentrations is crucial to prepare the high-quality SnxSy thin films using CBD. Moreover, the selection of suitable activation conditions such as solution/bath temperature, solution/bath pH (acidic or basic medium), deposition time, and stirring speed is also important [36,142] because they significantly affect the phase formation, growth, and properties of SnxSy films. In addition to the above parameters, the nature of the substrate and its cleaning procedure also affect the phase formation, growth, and properties of SnxSy films. Therefore, the understanding of the influence of all those parameters on the growth process of SnxSy films and their physical properties is necessary to deposit the quality films for device applications. In Table 4 and Table 5, the deposition parameters used for different thin films of tin sulfides made from chemical methods were summarized.

3.1. Overview of CBD Process of SnxSy Thin Films

The CBD refers to “a typical synthesis employing mild conditions [143]”. As schematically illustrated in Figure 5a, the experimental setup of CBD consists of the following parts: (i) magnetic stirrer with thermostat (to stir the mixed reactant solution continuously), (ii) oil bath (to maintain the desired temperature), (iii) substrate holder (to keep the substrates stable), (iv) stock chemical solutions to compose the reaction bath (mixture of different reagent solutions and its level always remains below the outer oil level), and (v) cleaned substrates [144]. The deposition of o-SnS, c-SnS, SnS2, and Sn2S3 by CBD was reported in 1987 [145], 2006 [146], 1990 [36,130], and 2012 [34], respectively. SnxSy films were deposited using various Sn precursors such as tin (II) chloride dihydrate (SnCl2∙2H2O), tin(IV) chloride pentahydrate (SnCl4∙5H2O), and tin ingots; various S precursors such as sodium sulfide (Na2S), ammonium sulfide (NH4)2S, sodium thiosulfate (Na2S2O3), thioacetamide (C2H5NS), and thiourea (CH4N2S); and various complexing agents such as triethanolamine (C6H15NO3), ammonia (NH3)/ammonium hydroxide (NH4OH), ammonium fluoride (NH4F), ammonium citrate (C6H17N3O7), trisodium citrate (Na3C6H5O7), citric acid (C6H8O7), tartaric acid (C4H6O6), ethylenediaminetetraacetic acid (C10H16N2O8), and disodium ethylenediaminetetraacetate (C10H14N2Na2O8). Among the above-mentioned chemicals, tin (II) chloride, thioacetamide, and triethanolamine, along with ammonia, were widely used as Sn precursor, S precursor, and complexing agents, respectively (Figure 5b). Other types of Sn precursors such as tin ingots [130,147] and tin (IV) chloride [131] were employed to deposit the SnS2 films. Except for the above reports, tin (II) chloride was used as an Sn source. In the case of S precursors, sodium thiosulfate was used as a second alternative to the regularly used thioacetamide.
The preparation of SnxSy thin films by CBD occurs when a substrate is immersed in the solution mixture of Sn ion (Sn2+ or Sn4+)-source, S ion (S2−)-source, and an appropriate complexing agent. In the deposition process, the Sn2+/Sn4+ ions are complexed through the coordinated bond formation by the complexing agent, which controls the rate of reaction [148]. At super saturation condition (Ionic product, Qip > Solubility product Ksp), SnxSy films can be deposited (Figure 5c). However, simply maintaining supersaturation condition in the bath will not provide acceptable quality SnxSy films; managing the solubility product of tin hydroxides is required because when an Sn precursor is dissolved in water, it rapidly binds with hydroxide ions, creating Sn(OH)2 and Sn(OH)4. The differences in Ksp values between SnS (1 × 10−25 mol2 dm−6), irrespective of the polymorphs, and SnS2 (1 × 10−46 mol3 dm−9) are very close to those between their hydroxides (Sn(OH)2 (1 × 10−28 mol3 dm−9) and Sn(OH)4 (1 × 10−56 mol5 dm−15)). Therefore, it is vital to monitor supersaturation with respect to an individual phase as well as the growth kinetics. In addition, the Ksp of SnxSy is affected by the concentration of precursor, solvent type, bath temperature, and bath pH [148,149]. Therefore, the optimum condition for the deposition of SnxSy thin film can be achieved by manipulating the above deposition parameters.
According to previous reports, the formation of o-SnS, c-SnS, SnS2, and Sn2S3 thin films is achieved through either an ion-by-ion mechanism (Figure 6a) or a simple cluster (hydroxide) mechanism (Figure 6b) based on the reaction process and parameters maintained in the bath [150]. The formation reaction of SnxSy thin films through the ion-by-ion mechanism and cluster (hydroxide) mechanism is as follows:
Ion-by-ion mechanism:
xSnp+ + ySq− → SnxSy
[∵ p = 2 or 4, q = 2; x = 1 or 2, y = 1, 2, or 3; SnxSy = SnS, SnS2, or Sn2S3]
Cluster (hydroxide) mechanism:
xSnp+ + y(OH)q− → Sn(OH)n
Sn(OH)n + ySq− → SnxSy + n(OH)
[∵ p = 2 or 4, q = 2; n = 2 or 4; x = 1 or 2, y = 1, 2, or 3; SnxSy = SnS, SnS2, or Sn2S3]
The physical properties of the CBD-deposited o-SnS, c-SnS, SnS2, and Sn2S3 thin films can be affected by the growth mechanism, level of supersaturation, and surface energy of the complexing agents [17]. Therefore, it is crucial to understand the actual mechanism undertaken in the solution for tuning the properties of the deposited films. Moreover, in the process of o-SnS, c-SnS, SnS2, and Sn2S3 thin film deposition, controlling the reaction to reduce or remove the spontaneous precipitation is essential, which can only be achieved by complexing the tin ions using an appropriate complexing agent (L). The kinetics of o-SnS, c-SnS, SnS2, and Sn2S3 thin film formation can be comprehended through the following reactions.
The complexing reactions in an aqueous Sn precursor solution are as follows:
SnCl2∙2H2O + L ⇔ Sn(L)2+ + 2Cl + 2H2O
SnCl4∙5H2O + L ⇔ Sn(L)4+ + 4Cl + 5H2O
The free Sn2+/Sn4+ ions are slowly released by the tin complex in a controlled way. As the tin complex dissociates, then
Sn(L)2+ ⇌ Sn2+ + L
Sn(L)4+ ⇌ Sn4+ + L
Here, the concentration of complex tin ions in the solution, Sn(L)2+ or Sn(L)4+, can be controlled by adjusting the concentration of the complexing agent and bath temperature [151]. If these ions can be generated, then the deposition of SnxSy thin films can be achieved. On the other hand, controlling the reaction by a slow and uniform generation of sulfur ions in the solution is also a significant factor when thin films are deposited. Thioacetamide (C2H5NS) is one of the most frequently employed S precursors. The hydrolysis of the S precursor can produce H2S and then S2− ions by the following reactions [152]:
CH3CSNH2 + H2O ⇌ CH3CONH2 + H2S
When the reaction attains an equilibrium [153], the following reactions are expected at a temperature of 25 °C:
H2S + H2O ⇌ H3O+ + HS (K0 = 10−7)
HS ⇌ H+ + S2− (K1 = 10−17)
HS + OH ⇌ H2O + S2− (K2 = 10−3)
At low pH values (<2.5), the reaction is controlled by the rate of hydrolysis of S precursor leading to the formation of hydrogen sulfide (H2S), whereas at higher pH values (>2.5), the reaction is controlled by the formation and decomposition of the tin–thioacetamide complex. Therefore, the pH of the bath and metal–thioacetamide complexes are also considered as the growth rate- and growth mechanism-determining components in the film formation [154].
The Sn ions react with the S ions and initiate the formation of tin sulfides (o-SnS, c-SnS, SnS2, and Sn2S3). The generation rate of Sn and S ions is controlled primarily by the source concentration, pH, and solution temperature. When the precursor concentration is changed, multiphase or other single-phase films can be formed by the following reactions [155,156]:
Sn2+ + S2− → SnS (Ksp = 1 × 10−25 at 25 °C)
Sn4+ + 2S2− → SnS2 (Ksp = 1 × 10−46 at 25 °C)
Sn2+ + Sn4+ + 3S2− → Sn2S3
Sn(II)S + Sn(I V)S2 → Sn(II)(IV)2S3 (or) SnS + SnS2 → Sn2S3

3.2. Sn and S Precursors and Their Concentration Effect

The selection of Sn precursor and its concentration plays a vital role in the growth, phase formation, crystallinity, preferred orientation, morphology, band gap, and other properties of SnxSy thin films [251]. This is because the releasing rate of Sn ions strongly depends on the selection of Sn precursors. As mentioned in Section 3.1, the SnCl2∙2H2O (T(II)C) has been considerably utilized as an Sn precursor for the deposition of SnxSy films. According to the literature (Table 4 and Table 5), until recently, there have been no reports related to the study of different types of Sn precursors on the formation of SnxSy films and the Sn precursor concentration effects on the formation of SnS2 and Sn2S3 films and their properties. However, there have been very few quantitative analyses of Sn precursor concentration effect on the formation of o-SnS films and their properties. The primary report related to the effect of Sn precursor concentration ([T(II)C] = 0.06–0.12 M) on the growth of o-SnS films was made in 2012 [191]. A lower T(II)C concentration stimulates the formation of multi phases with a dominant SnS2 phase, whereas a higher T(II)C concentration reduces the crystallinity. The T(II)C concentration of 0.1 M is beneficial for the deposition of pure, good crystalline o-SnS with (111) preferred orientation (Figure 7a). A small variation in T(II)C concentration (at 0.15 M) changes the preferred orientation of o-SnS from (111) to (200) [194]. Moreover, the change in T(II)C concentration can increase the grain size and decrease the band gap (1.95–1.5 eV) (Figure 7b,c) [191]. Therefore, the manipulation of preferred orientation, crystallinity, and band gap can be achieved by the change in Sn precursor concentration.
In addition to the suitable Sn precursor selection, the choice of S precursor and its concentration are highly desirable to obtain good quality SnxSy films. In CBD, the releasing rate (or reaction rate) of S ions greatly affect the growth kinetics and phase formation, and it can be controlled by the S precursor concentration. According to the previous reports (Table 4 and Table 5), TA and ST have been chiefly used as S ion sources (Figure 5b). In those, TA is preferable compared to ST because it works in both acidic and alkaline bath conditions. The influence of TA concentration on o-SnS film growth (thickness) was initially reported in 1987 [145]. An extremely low or high TA concentration produces the o-SnS films of smaller terminal thickness, whereas a moderate TA concentration promotes the growth of maximum thickness (Figure 7d). The reason for the lower film thickness obtained at a lower S precursor concentration is the insufficient number of S ions in the reaction bath that can combine with all the available Sn ions. At a higher S precursor concentration, the releasing rate of S ions is high enough to stimulate the precipitation process, which also results in a lower film thickness [145].
Furthermore, the S precursor concentration can influence the morphology and phase formation of o-SnS films (Figure 7e). A higher TA concentration stimulates the formation of multi phases such as Sn2S3 and Sn3S4 (Sn2S3 + SnS → Sn3S4) [209] and a lower TA concentration assists the growth of single-phase o-SnS films, but with lower crystallinity [142,198]. A TA concentration of 0.1 M is preferable for the deposition of a single-phase, polycrystalline o-SnS with (101) preferred orientation [198], and a ST concentration of 0.75 M is advisable for (111)/(040) preferred orientation (Figure 7f). The effect of changes in the S source concentration on the band gap of o-SnS films is controversial until the present. A reduction in band gap from 1.70 eV to 1.25 eV with increasing TA concentrations was reported in [200], although no significant change in band gap was found with TA concentration in [198]. On the other hand, no studies in the literature have focused on the influence of S precursor concentration on the properties of c-SnS, SnS2, and Sn2S3 films.

3.3. Complexing Agents and Their Concentration Effect

As stated in Section 3.1, in order to develop influential SnxSy films, the control of the availability of Sn ions in the reaction bath is essential. It can be successfully attained by the addition of an appropriate concentration of a complexing agent [128,148]. Moreover, adhesion, morphology, crystallinity, and the deposition rate of SnxSy films can be significantly affected by the concentration of the complexing agent [128]. Therefore, knowledge of the behavior of complexing agents in the bath can help to obtain good quality SnxSy films. The behavior of complexing agents is described in terms of their stability constants (Ks), which is the equilibrium constant for the formation of a complex in a solution [252]. It is defined for the equilibrium between an Sn ion (Sn2+/4+) and a ligand (L) as [148]
K S = a Sn 2 + / 4 +   L a Sn 2 + / 4 + +   a L
where a is the activity of subscripted species and can be approximated by its concentration. A large value of Ks implies a strong binding affinity for the metal (Sn) ion, while a small value of Ks implies a weak binding affinity [148]. Generally, complexing agents can prevent the formation of powder/bulk precipitation of tin hydroxides in the reaction bath, and they can easily maintain the supersaturating condition. If a complexing agent has a weak binding affinity, it does not arrest the bulk precipitation of tin hydroxides. On the other hand, if it has an extremely strong binding affinity, it restricts the deposition of the desired film [253]. Therefore, in order to prevent powder/bulk precipitation of tin hydroxides, the complexing binding affinity must be intermediate.
Various complexing agents have been explored to control Sn ions depending on the bath conditions during the deposition of SnxSy films (Table 4 and Table 5). However, there are only a few reports on the study of complexing agent concentration. Initially, the influence of TEA complexing agent concentration on the thickness of o-SnS films was made in 1987 [145]. An optimized TEA complexing agent concentration controls the formation of o-SnS films, yielding a thick o-SnS film (Figure 7g). In addition to the growth (thickness), the change in TEA complexing agent concentration can also influence the phase formation and crystallinity of o-SnS, c-SnS, and SnS2 films. The lower tartaric acid (TTA) complexing agent concentration creates the weak tin complexation, leading to partial homogeneous precipitation, resulting in low-crystalline o-SnS films. As the complexing agent concentration increases, the improved tin complexation controls the reaction, yielding the formation of better crystalline films o-SnS. Over the limit, the availability of free Sn ions is reduced due to strong complexation, resulting in the formation of sulfur-rich tin phases such as Sn2S3 and SnS2 (Figure 7h) [211]. Single-phase, polycrystalline o-SnS films with (111) preferred orientation are produced at 1.85 M of TEA [191] and 1.4 M of TTA [211], while a c-SnS (222)/(400) is formed at 0.125 M of EDTA [26] concentrations (Figure 7i). The crystallinity of o-SnS films can be improved by replacing the lower stability (Sn2+-TEA) complexing agent with the higher stability (Sn2+-EDTA) one [198,254], due to the fact that EDTA (hexaligand) may generate a ligand more quickly than TEA (triligand) [255]. An increase in citric acid and ammonia concentration also improves the crystallinity in the case of SnS2 films (Figure 7j) [41,131]. The concentration of the complexing agent similarly influences the morphological and optical properties of the o-SnS, c-SnS, and SnS2 films. The direct optical energy gap for o-SnS films reduces with increasing complexing agent concentration (TSC, 0.06–0.08 M; TEA, 12.5–13 M; TTA, 0.6–1.4 M) from 2.16 eV to 1.17 eV [50,182,214], but rises from 1.67 eV to 1.73 eV [26] for c-SnS films with EDTA (0.075–0.125 M) (Figure 7l). The change in complexing agent concentration (TSC, TTA) improves the compactness and morphology of o-SnS films (Figure 7m). This may improve their electrical properties, such as electrical mobility (~228 cm2V−1s−1) and carrier concentration (~4.1 × 1015 cm−3). No previous study has examined the effect of complexing agents on the formation and physical properties of Sn2S3 films.

3.4. Solution pH Effect

In CBD, solution pH/bath pH (a measure of the acidity or basicity of a solution) is an important parameter because it directly affects the growth mechanism as well as reaction rate. Therefore, it can influence the formation of phases and physical properties of films [27]. In addition, the bath pH must be at a specific optimum value to maintain supersaturation condition (Qip > Ksp, Figure 5c) for the formation of SnxSy films. The preparation of SnxSy films was reported both in acidic (pH < 7) and alkaline (pH > 7) baths (see Table 4 and Table 5). When the bath pH is varied between 1 and 14, the concentration of OH- ions increases, which results in a reduction in the concentration of free Sn2+ or Sn4+ ions in the solution. Thus, the hydroxide mechanism can predominate during film development, resulting in the creation of Sn(OH)2 or 4 in addition to SnxSy. A higher bath pH, on the other hand, encourages the hydrolysis of a sulfur source precursor.
A few researchers have investigated the bath pH effect on o-SnS and c-SnS films growth and their physical properties (Table 4). The bath pH effect on the adhesion and growth rate of o-SnS films was first reported in 1989 [36]. According to this report, the good adhesion of o-SnS films on glass can be obtained with a bath pH >3. The growth rate is low at pH~7 and high at pH~10 for o-SnS films due to the formation of Sn(OH)2 or 4 precipitate from the hydrolysis of an Sn precursor because a part of Sn(OH)2 or 4 precipitate turns into Na2SnO2, which dissolves back in the solution. The change in growth rate by bath pH leads to the variation in grain size of o-SnS films [183]. The bath pH can also influence the growth mechanism, which leads to phase transformation [184,185]. An o-SnS forms at a lower pH of 6.5 via the cluster-by-cluster mechanism, whereas c-SnS forms at a higher pH of 7.0 through the ion-by-ion mechanism (Figure 7n,o) [27].
The phase transition caused by the change in bath pH leads to a change in the morphologies of the film surfaces (Figure 7o) and the energy band gaps (o-SnS: 1.51 eV and c-SnS: 1.64 eV) (Figure 7p) [27,184,185]. The increase in solution pH results in the decrease of free Sn2+ ion concentration as well as the concentration of OH ions, which are favorable for the hydrolysis of the S ion source [256], leading to the increase in the concentration of S2− ions. Thus, the interaction of Sn2+ and S2− ions can form the c-SnS via an ion-by-ion mechanism because the potential barrier of heterogeneous nucleation is lower than that of homogeneous nucleation [25,185]. However, these mechanisms are speculated, and direct in situ measurement evidence such as in situ quartz crystal microbalance and electrochemical impedance is lacking. Therefore, such studies are required for understanding the growth mechanism of SnxSy films [257]. On the other hand, no studies in the literature have examined the effect of bath pH on the properties of SnS2 and Sn2S3 films.

3.5. Solution Temperature (Tb) Effect

In CBD, solution temperature/bath temperature (Tb) also played a crucial role in the preparation of thin films with high quality and desired features. It critically enhances the rate of dissociation of the precursors and thus strongly affects the thickness, growth rate, type of nucleation, crystalline phase, crystallite size, morphology, and optoelectrical properties of thin films. The change in film growth rate as a function of Tb can be determined through the Arrhenius equation [258],
k ( T ) = Ae E a RT
where k(T) is the temperature-dependent growth rate for the given deposition conditions, A is a pre-exponential constant related to the initial reagent concentration, Ea is the activation energy (kJ/mol), and R is the gas constant (R = 8.3145 J mol−1 K−1).
The deposition of SnxSy films has been reported in the Tb range of room temperature (Tr)—90 °C (Table 4 and Table 5). The effect of Tb on the formation of o-SnS, c-SnS films, and their properties (Table 4) was studied extensively. The Tb changes the growth rate due to the variation in the deposition mechanism, i.e., the ion-by-ion mechanism, which is less thermally activated with low activation energy (at lower bath temperatures). In contrast, cluster-by-cluster is believed to occur at relatively higher temperatures [259]. Thus, the thickness of a film has a close relationship with the Tb. First, it increases significantly with the Tb due to the increase in bath supersaturation [145,213] and reaches saturation point very quickly because the hydrolysis of the S precursor is greatly improved by the increase in Tb [182] (Figure 8a). Then, it decreases down to a terminal point because of the ion–ion condensation process and high homogeneous precipitation rate [145,146]. In addition, the Tb significantly affects the microstructures of o-SnS and c-SnS films.
Generally, the films prepared at lower Tb have smaller grains, and those grains increase in size with Tb due to the covering of voids by secondary nucleation (Figure 8b) [50,212]. The change in grain shape may indicate a change in the growth mechanism [146]. The Tb can also influence the composition of c-SnS films. The c-SnS films show a non-stoichiometric composition at higher and lower Tb values due to the relatively faster and slower release of Sn2+ ions from the tin complex due to the variation of thermal energy in the solution [50]. Single-phase, polycrystalline o-SnS films with (111) preferred orientation and c-SnS films with (222)/(400) preferred orientation are produced separately at Tb of 70 °C [212] and 65 °C [50] using a different source of materials (Table 4), respectively.
On the other hand, the Tb shows an impact on phase transformations when other bath parameters remained constant. The films predominantly exhibit the o-SnS phase above the Tb range of 30–40 °C, whereas the c-SnS phase is below this range (20–30 °C) for particular deposition conditions [146]. The Tb can directly affect the crystallinity of both o-SnS and c-SnS films. The crystallinity of both films is improved with Tb due to the supply of sufficient thermal energy for further crystallization (Figure 8c) [50,212]. Thus, an average crystallite size is improved with Tb—however, up to a certain extent [212]. As the Tb improves the kinetic energy of the reactants and accelerates the interaction between all ions in the reaction bath, the nuclei formation (crystallite grow) is enhanced on the surface of the substrate [260]. However, at higher Tb, the crystallite size is decreased due to the dissolution of grown film.
The Tb also influences the optical characteristics of the o-SnS and c-SnS films. In o-SnS films, the Tb improves the sharpness of the absorption edge with a high optical absorption coefficient (>104 cm−1) [17] (Figure 8d), which is suitable for PV devices. The optical energy gap of o-SnS and c-SnS films decreases from 1.41 eV to 1.30 eV and from 1.74 eV to 1.68 eV, respectively, with the increase in Tb (30–70 °C) [17,50]. As mentioned above, Tb can improve the grain size and simultaneously reduce height (smoothness of the surface) and the number of grain boundaries [261]; this minimizes imperfections in the film and enhances the quality of the film, which can lead to change in density of localized states within the energy gap [262]. Therefore, band gap tuning is easily possible in CBD deposited o-SnS and c-SnS films regarding Tb, which is essential for designing highly efficient solar cells [263]. The optical parameters, namely, refractive index (n), extinction coefficient (k), and real/imaginary dielectric constants of o-SnS films, are in ranges of 2.72–3.24, 0.24–0.13, and 7.34–10.48/0.85–1.32 [17], respectively (Figure 8e). Here, the variation in the optical parameters may be arrived from the change in strain and packing density with Tb [264].
As previously mentioned, Tb improves the crystallinity along with grain size and thickness. Thus, the scattering of charge carriers by grain boundaries decreases with respect to Tb, which makes a significant change in the electrical characteristics of films [265]. These possible reasons may improve the carrier density and a consequent reduction in resistivity in both o-SnS and c-SnS films. The reduction of the dispersing effects of carriers can lead to an increase in the mobility of carriers in those films (~55 cm2 V−1 s−1 for o-SnS at 70 °C [212] and 28 cm2 V−1 s−1 for c-SnS at 45 °C [50]) (Figure 8f). However, the above-mentioned description confirms the importance of Tb in the CBD process until there are no reports on the Tb influence on both SnS2 and Sn2S3 films.

3.6. Deposition Time Effect

Deposition time (td) is the most important bath parameter among the various deposition conditions. It affects the growth rate/thickness and properties of SnxSy films. In CBD, the td period can be divided into three steps, namely, (i) nucleation or initiation, which requires high activation energy; (ii) linear growth, which includes heterogeneous growth of nuclei; and (iii) termination or saturation, in which chemical reagents become depleted and the reaction begins to slow down and eventually stops [266]. Thus, the rate of formation of nuclei can be in terms of td by the following Avrami equation (a conventional diffusion-controlled reaction model) [267]:
α = 1 e ( kt d ) n
where α is the fractional decomposition (or reaction), k is a rate constant, and n is the Avrami exponent.
There are a considerable number of reports on the study of the td effect on o-SnS films, but there are only a few reports for c-SnS [210,225], SnS2 [56], and Sn2S3 [34] films (Table 4 and Table 5). Typically, a td from a few minutes to several hours has been reported to prepare these SnxSy films (Table 4 and Table 5). The growth of SnxSy films with td was simply described in terms of thickness. In the initial state, the change in film thickness is insignificant because of the requirement of long incubation time for nucleation [198], and the thickness increases linearly due to the availability of sufficient amounts of Sn2+ or Sn4+ and S2− ions. Next, the film thickness increases faster, then decreases at a longer deposition time, and attains a maximum value as a terminal/final thickness. Here, the attained terminal thickness is not only td-dependent but also Tb-related [145]. Thus, a terminal thickness should be considered when the reaction undergoes at a constant temperature. A terminal thickness in the range of 120–900 nm can be obtained for different td varying from 1 h to 24 h at a constant range of Tb (Tr—75 °C) for o-SnS and c-SnS films [128,145,175,198,225], and a thickness of 152 nm can be attained at a td of 90 min for SnS2 films (Figure 8g) [56]. The variation in the thickness (growth) of these films with respect to td can be explained by considering two competing processes taking place in the deposition bath. One process includes heterogeneous precipitation, which leads to film growth (thickness improves). The other involves the dissolution of the preformed film, which results in the decrease of film thickness.
The td has a significant impact on the surface morphology, crystallinity, crystallite size, and phase purity of SnxSy films. As the td increases, the size and quantity of grains (or aggregations) can be improved to form a more homogeneous film [177] (Figure 8h). This indicates an occurrence of nucleation growth with td [56,190]. If td exceeds the optimum value, a non-uniform film with porous nature might be formed due to the dissolution of pre-adhered grains in the film [190]. This phenomenon can be experimentally observed for o-SnS, SnS2, and Sn2S3 films when the td varies between 2 and 10 h [177], 30–120 min [56], and 20–24 h [34], respectively. The td can considerably improve the crystallinity of the SnxSy films and simultaneously enhance the crystallite size. However, beyond the limit of td, the crystallinity becomes poor, and the crystallite size decreases (Figure 8i) [56,146,177,190,225]. The reduction of crystallite size is due to the lowering of the van der Waals force in between crystallites because the substrate remained in the solution longer than necessary [177]. In addition to the crystallinity of films, the td also influences the phase purity of a film. At low td, the released Sn2+ or Sn4+ ions are relatively low in the reaction bath compared to the available S2− ions. These available S2− ions are not balanced by the all released Sn2+ or Sn4+ ions, leading to the development of other secondary phases, whereas at longer td they are counterbalanced, promoting the growth of the pure phase [56].
As mentioned previously, the td directly influences the thickness of SnxSy films. Thus, it tremendously shows an impact on their optical transmittance/absorbance. Always, shorter td periods generate the thinnest film of high transmittance, which might be affected by abundant porosities [175,177]. Simultaneously, the more extended td periods produce thick films of high absorption [34,175,190], essential for solar cell application. The longer td period also improves the size of crystallites that can affect the optical absorption and the band gap energy of films [190].
The quantum size effect and changing barrier height (or variation in grain size) are also responsible for the variation in the band gap of films with td at other identical growth conditions [22,177]. The increase in td period reduces the band gap of o-SnS, SnS2, and Sn2S3 films from 1.83 eV to 1.30 eV [177], 2.95 eV to 2.80 eV [56], and 2.12 to 2.03 eV [34], respectively. In contrast, the longer td period generates greater compression impacts with the thickness in o-SnS films, which may enhance the band gap (0.82–1.22 eV) [175]. In addition to the optical gap, td shows a significant effect on the optical constants such as refractive index (n, SnS2:2.57–2.63, Sn2S3: 4.89–7.18) and extinction coefficient (k, SnS2: 0.69–0.61, Sn2S3: 0.0015–0.0019) (Figure 8j,k) [34,56]. However, no previous studies had included the variations in optical constants of o-SnS and c-SnS films with td. On the other hand, the td reduces the electrical resistivity and improves the carrier density and mobility of carriers in the case of both o-SnS and SnS2 films [56,177] due to the improved crystallinity and suppression of secondary phases with td period. In contrast, in the case of c-SnS and Sn2S3 films, there are no reports available in the literature.

3.7. Other Parameters

3.7.1. Substrate Nature and Its Cleaning Process Effect

Generally, thin films require proper mechanical support that provides sufficient adhesion. These supports are commonly called substrates. Substrates have a significant effect on the film properties in practice [268]. Therefore, the choice of a suitable substrate with a specific form for a thin film with a particular application is critical since the substrate must be structurally and chemically compatible with the thin film material in terms of thermal and mechanical stability [269,270]. Moreover, the substrate nature strongly affects the preferred orientation of a thin film, which plays a major role in device performance [271]. Therefore, currently, the exploration of feasible substrates has become an active research area. The CBD has the benefit of allowing thin film deposition on unevenly shaped surfaces. However, the substrate nature greatly affects the deposition process and film quality. Usually, substrates with rough surfaces have better anchoring of the initial deposit in the tiny valleys. Substrates such as glass, tin oxide (TO), indium tin oxide (ITO), and silica/quartz are relatively reactive, owing to the presence of hydroxyl surface groups. Furthermore, when the lattice of the deposited material matches well with that of the substrate, the free energy change is smaller; this facilitates fast nucleation with good morphology and structure. Although the substrate nature has more impact on the process of CBD and the deposited thin film characteristics, there are only a few studies on this area in the case of o-SnS and SnS2 films and no reports for c-SnS and Sn2S3 films (Table 4 and Table 5). The reports related to the effects of molybdenum (Mo), ITO, and TO and borosilicate glass substrates [204] on the properties of o-SnS films and the glass, TO, and titanium (Ti) substrates [130] on SnS2 films are available in the literature. At 0.01 M of Sn and S sources concentrations, both Mo and TO substrates generate o-SnS films with a better surface coverage, whereas the borosilicate glass and ITO substrates produce a discontinuous film with separate agglomerated o-SnS particles. When the concentration of sources is 0.03 M, all substrates except the borosilicate glass form a complete and uniform coverage of o-SnS films. At a high concentration of 0.09 M, all substrates produce a complete coverage of o-SnS films but with a lower adhesive nature [204]. In the case of SnS2, the amorphous and n-type nature films formed on the glass and Ti substrates, respectively.
In addition to the substrate nature, the cleaning process of the substrate also significantly affects the quality of thin films. Improper cleaning of substrates results in the formation of pinholes in the film, which creates major issues on the fabrication of large-area devices and produces short circuits in solar cells [272]. Unfortunately, there is a lack of research on this area for CBD deposited o-SnS, c-SnS, SnS2, and Sn2S3 thin films.

3.7.2. Stirring Speed and Humidity Effect

In the CBD, chemical solutions with a homogeneous distribution of precursors are necessary before starting the process. Continuous mixing of the reaction solution is mandatory for realizing a uniform thin film deposition [273]. This could be achieved by stirring the solutions at appropriate speeds. At the beginning of the deposition, the stirring speed does not have a significant impact on the growth rate of thin films. However, for longer deposition times, it directly affects the growth rate. In addition, stirring with uneven speed may produce a variation in thin film uniformity and improper diffusion of complex ions toward the substrate [274], and stirring provokes precipitation and reduces the final thickness of the film. Therefore, care must be taken in stirring the solution to obtain the desired quality of thin films.
On the other hand, environmental humidity also influences the formation and physical properties of CBD processed films [273] since the CBD can be performed in an open environment where the gas–liquid interface is influenced by moisture. Even after the deposition of films, they considerably degrade because of their colloidal nature [275]. Therefore, the maintenance of environmental humidity is vital for the deposition of defect-free films. Although the control of stirring speed and environmental humidity is essential for producing quality SnxSy films, there is no systematic study on these effects in the literature.

3.8. Summary

SnxSy are binary metal chalcogenides that have attracted considerable attention due to their abundant, low cost, and nontoxic constituent elements. In comparison to other vacuum and chemical approaches, they may be simply synthesized utilizing a simple non-vacuum CBD methodology. Sn precursors, S precursors, and complexing agents are ideally T(II)C, TA, and TEA, respectively. The following lines are made based on the examination of published data (Table 4 and Table 5) and the explanation in Section 3.1, Section 3.2, Section 3.3, Section 3.4, Section 3.5, Section 3.6 and Section 3.7. Changes in Sn precursor concentration, complexing agent concentration, and Tb can be used to manipulate high-intensity plans and crystallinity. Maintaining complexing agent concentration, bath pH, and td, may regulate phase transition and growth rate. Controlling S precursor and complexing agent concentrations results in good morphological, optical, and electrical characteristics. As a result, optimizing each deposition parameter is critical for producing high-quality SnxSy thin films for a variety of applications. However, no previous research has looked at the effect of S precursor concentration on c-SnS, SnS2, and Sn2S3 films; complexing agent concentration on Sn2S3 films; bath pH on the properties of SnS2 and Sn2S3 films; Tb on both SnS2 and Sn2S3 films; and td on c-SnS and Sn2S3 films. Furthermore, for all o-SnS, c-SnS, SnS2, and Sn2S3 thin films, there is a dearth of research on the substrate nature-cleaning procedure, stirring speed, and humidity influence.
According to the description in this part, it is confirmed that further research is required to improve the quality of SnxSy films and more studies are necessary related to the optimization of all deposition parameters. Hence, research focusing on this area is essential.

4. Conclusions

SnxSy thin films deposited with CBD are a relatively recent development, and their process–property correlations must be understood for the desired application. Further, the fabrication of single-phase o-SnS, c-SnS, SnS2, and Sn2S3 thin films in CBD is very condition-dependent. Additionally, it is crucial to identify and separate the o-SnS, c-SnS, SnS2, and Sn2S3 phases. However, until recently, there has been a dearth of detailed studies on the optimization of growth parameters. The present review outlined the background and basic properties of SnxSy (o-SnS, c-SnS, SnS2, and Sn2S3) along with the principle, nucleation, growth, and growth mechanism of SnxSy thin films by CBD. Furthermore, the influence of growth parameters such as precursor concentration (tin source, sulfur source, and complexing agent), bath pH, bath temperature (Tb), deposition time (td) on the phase formation, and physical properties of SnxSy thin films were comprehensively described. As a result, the reader should be able to prepare single-phase tin sulfide materials with ease after studying the present article. Hence, the present review should motivate readers to conduct extensive investigations on SnxSy films to develop cost-effective, eco-friendly, and earth-abundant tin sulfide materials to meet all future energy requirements. The connection between the physical properties of SnxSy thin films and their photovoltaic application will be discussed in our subsequent article.

Author Contributions

Conceptualization, V.R.M.R.; writing—original draft preparation, S.G.; review, V.R.M.R., T.R.R.K., and C.P.; Writing final version—review and editing, W.K.K.; supervision, W.K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Priority Research Centers Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (2014R1A6A1031189) and the 2019 Yeungnam University Research Grant.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbbreviationChemical Name
AAmmonia
AAAcetic acid
ACAmmonium citrate
ACEAcetone
AFAmmonium fluoride
AHAmmonium hydroxide
ALDAtomic layer deposition
ASAmmonium sulfide
BTBaking temperature
CACitric acid
CALPHADCALculation of PHAse diagram
CBDChemical bath deposition
CBMConduction band minimum
CBOConduction band offset
CSSClose space sublimation
CUBCubic
DDTDodecanethiol
DIWDeionized water
DWDistilled water
EDSEnergy-dispersive X-ray spectroscopy
EDTAEthylenediaminetetraacetic acid
ELElectrolyte
FFFill factor
GGlass
GAGlacial acetic acid
GIXRDGrazing incidence X-ray diffraction
HCLHydrochloric acid
HEXHexagonal
HHHydrazine hydrate
HWVDHot wall vapor deposition
ITOIndium tin oxide
JCPDSJoint committee on powder diffraction standards
LiLithium
MeOHMethanol
MoMolybdenum
NaSodium
Na2EDTADisodium ethylenediaminetetraacetate
NTANitriloacetic acid
ODEOoctadecene
OLAOleylamine
ORTOrthorhombic
PGPropylene glycol
PLPhotoluminescence
QEQuantum efficiency
RSRock salt
SAEDSelected area electron diffraction
SCRSpace charge region
SDSSodium sulfide
SiSilicon
SILARSuccessive ionic layer adsorption and reaction
SIMSSecondary ion mass spectrometry
SSStainless steel
SnSTin monosulfide
SnS2Tin disulfide
Sn2S3Tin sesquisulfide
STSodium thiosulfate
TAThioacetamide
T(II)CTin (II) chloride dehydrate
TC(IV)Tin(IV) chloride pentahydrate
TEATriethanolamine
TEMTransmission electron microscopy
TiTitanium
TOTin oxide
TOPTrioctylphosphine oxide
TrRoom temperature
TSCTrisodium citrate
TTATartaric acid
TUThiourea
UAEDUltrasound-assisted electrodeposition
VBMValence band maximum
XRDX-ray diffraction
ZBZinc blended

References

  1. De Elisa, R. Cadmium telluride solar cells: Selenium diffusion unveiled. Nat. Energy 2016, 1, 16143–16146. [Google Scholar]
  2. Solar Frontier Achieves World Record Thin-Film Solar Cell Efficiency of 23.35%. Available online: http://www.solar-frontier.com/eng/news/2019/0117_press.html (accessed on 14 July 2021).
  3. Chen, X.; Hou, Y.; Zhang, B.; Yang, X.H.; Yang, H.G. Low-cost SnSx counter electrodes for dye-sensitized solar cells. Chem. Commun. 2013, 49, 5793–5795. [Google Scholar] [CrossRef] [PubMed]
  4. PV Division | Business Information | KITAGAWA SEIKI Co., Ltd. Available online: http://www.kitagawaseiki.co.jp/en/jigyo_pv.html (accessed on 14 July 2021).
  5. Ge, J.; Zhang, Y.; Heo, Y.-J.; Park, S.-J. Advanced Design and Synthesis of Composite Photocatalysts for the Remediation of Wastewater: A Review. Catalysts 2019, 9, 122. [Google Scholar] [CrossRef] [Green Version]
  6. Liao, C.-H.; Huang, C.-W.; Wu, J.C.S. Hydrogen Production from Semiconductor-based Photocatalysis via Water Splitting. Catalysts 2012, 2, 490–516. [Google Scholar] [CrossRef] [Green Version]
  7. Zhao, Y.; Yang, Q.; Chang, Y.; Pang, W.; Zhang, H.; Duan, X. Novel Gas Sensor Arrays Based on High-Q SAM-Modified Piezotransduced Single-Crystal Silicon Bulk Acoustic Resonators. Sensors 2017, 17, 1507. [Google Scholar] [CrossRef]
  8. Campaña, A.; Florez, S.; Noguera, M.; Fuentes, O.; Ruiz Puentes, P.; Cruz, J.; Osma, J. Enzyme-Based Electrochemical Biosensors for Microfluidic Platforms to Detect Pharmaceutical Residues in Wastewater. Biosensors 2019, 9, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Lee, S.-H.; Park, J.-U.; Kim, G.; Jee, D.-W.; Kim, J.H.; Kim, S. Rigorous Study on Hump Phenomena in Surrounding Channel Nanowire (SCNW) Tunnel Field-Effect Transistor (TFET). Appl. Sci. 2020, 10, 3596. [Google Scholar] [CrossRef]
  10. Mohammed, M.; Chun, D.-M. Electrochemical Performance of Few-Layer Graphene Nano-Flake Supercapacitors Prepared by the Vacuum Kinetic Spray Method. Coatings 2018, 8, 302. [Google Scholar] [CrossRef] [Green Version]
  11. Wang, Y.; Mao, Y.; Ji, Q.; Yang, M.; Yang, Z.; Lin, H. Electrostatic Discharge Characteristics of SiGe Source/Drain PNN Tunnel FET. Electronics 2021, 10, 454. [Google Scholar] [CrossRef]
  12. Li, S.; Lam, K.H.; Cheng, K.W.E. The Thermoelectric Analysis of Different Heat Flux Conduction Materials for Power Generation Board. Energies 2017, 10, 1781. [Google Scholar] [CrossRef] [Green Version]
  13. Kumar, S.G.; Rao, K.S.R.K. Physics and chemistry of CdTe/CdS thin film heterojunction photovoltaic devices: Fundamental and critical aspects. Energy Environ. Sci. 2014, 7, 45–102. [Google Scholar] [CrossRef]
  14. Alzoubi, T.; Moustafa, M. Numerical optimization of absorber and CdS buffer layers in CIGS solar cells using SCAPS. Int. J. Smart Grid Clean Energy 2019, 8, 291–298. [Google Scholar] [CrossRef]
  15. Bag, S.; Gunawan, O.; Gokmen, T.; Zhu, Y.; Todorov, T.K.; Mitzi, D.B. Low band gap liquid-processed CZTSe solar cell with 10.1% efficiency. Energy Environ. Sci. 2012, 5, 7060–7065. [Google Scholar] [CrossRef]
  16. Jiang, F.; Shen, H.; Gao, C.; Liu, B.; Lin, L.; Shen, Z. Preparation and properties of SnS film grown by two-stage process. Appl. Surf. Sci. 2011, 257, 4901–4905. [Google Scholar] [CrossRef]
  17. Sreedevi, G.; Vasudeva Reddy, M.; Park, C.; Jeon, C.W.; Ramakrishna Reddy, K.T. Comprehensive optical studies on SnS layers synthesized by chemical bath deposition. Opt. Mater. 2015, 42, 468–475. [Google Scholar] [CrossRef]
  18. Koteeswara Reddy, N.; Ramakrishna Reddy, K.T. Preparation and characterisation of sprayed tin sulphide films grown at different precursor concentrations. Mater. Chem. Phys. 2007, 102, 13–18. [Google Scholar] [CrossRef]
  19. Jiang, F.; Shen, H.; Wang, W.; Zhang, L. Preparation of SnS film by sulfurization and SnS/a-Si heterojunction solar cells. J. Electrochem. Soc. 2012, 159, H235. [Google Scholar] [CrossRef]
  20. Devika, M.; Koteeswara Reddy, N.; Venkatramana Reddy, S.; Ramesh, K.; Gunasekhar, K.R. Influence of rapid thermal annealing (RTA) on the structural and electrical properties of SnS films. J. Mater. Sci. Mater. Electron. 2009, 20, 1129–1134. [Google Scholar] [CrossRef]
  21. 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]
  22. Ogah, O.E.; Zoppi, G.; Forbes, I.; Miles, R.W. Thin films of tin sulphide for use in thin film solar cell devices. Thin Solid Films 2009, 517, 2485–2488. [Google Scholar] [CrossRef]
  23. Ragina, A.J.; Murali, K.V.; Preetha, K.C.; Deepa, K.; Remadevi, T.L. A Study of optical parameters of tin sulphide thin films using the swanepoel method. In AIP Conference Proceedings; American Institute of Physics: College Park, MD, USA, 2011; Volume 1391, pp. 752–754. [Google Scholar]
  24. Mahdi, M.S.; Ibrahim, K.; Hmood, A.; Ahmed, N.M.; Mustafa, F.I.; Azzez, S.A. High performance near infrared photodetector based on cubic crystal structure SnS thin film on a glass substrate. Mater. Lett. 2017, 200, 10–13. [Google Scholar] [CrossRef]
  25. 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 Appl. Mater. Sci. 2016, 213, 170–177. [Google Scholar] [CrossRef]
  26. Chalapathi, U.; Poornaprakash, B.; Park, S.-H. Growth and properties of cubic SnS films prepared by chemical bath deposition using EDTA as the complexing agent. J. Alloys Compd. 2016, 689, 938–944. [Google Scholar] [CrossRef]
  27. Mahdi, M.S.; Ibrahim, K.; Ahmed, N.M.; Hmood, A.; Azzez, S.A.; Mustafa, F.I.; Bououdina, M. Influence of pH value on structural, optical and photoresponse properties of SnS films grown via chemical bath deposition. Mater. Lett. 2018, 210, 279–282. [Google Scholar] [CrossRef]
  28. Garcia-Angelmo, A.R.; Romano-Trujillo, R.; Campos-Álvarez, 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 2015, 212, 2332–2340. [Google Scholar] [CrossRef]
  29. Chalapathi, U.; Poornaprakash, B.; Park, S.H. Effect of post-deposition annealing on the growth and properties of cubic SnS films. Superlattices Microstruct. 2017, 103, 221–229. [Google Scholar] [CrossRef]
  30. Alpen, U.V.; Fenner, J.; Gmelin, E. Semiconductors of the type MeIIMeIVS3. Strateg. Surv. 1975, 76, 50–54. [Google Scholar] [CrossRef]
  31. Lopez, S.; Ortiz, A. Spray pyrolysis deposition of SnxSy thin films. Semicond. Sci. Technol. 1994, 9, 2130–2133. [Google Scholar] [CrossRef]
  32. Godoy-Rosas, R.; Barraza-Felix, S.; Ramirez-Bon, R.; Ochoa-Landin, R.; Pineda-Leon, H.A.; Flores-Acosta, M.; Ruvalcaba-Manzo, S.G.; Acosta-Enriquez, M.C.; Castillo, S.J. Synthesis and characterization of Sn2S3 as nanoparticles, powders and thin films, using soft chemistry reactions. Chalcogenide Lett. 2017, 14, 365–371. [Google Scholar]
  33. Koteeswara Reddy, N.; Ramakrishna Reddy, K.T. Optical behaviour of sprayed tin sulphide thin films. Mater. Res. Bull. 2006, 41, 414–422. [Google Scholar] [CrossRef]
  34. Guneri, E.; Gode, F.; Boyarbay, B.; Gumus, C. Structural and optical studies of chemically deposited Sn2S3 thin films. Mater. Res. Bull. 2012, 47, 3738–3742. [Google Scholar] [CrossRef]
  35. Kanevce, A.; Reese, M.O.; Barnes, T.M.; Jensen, S.A.; Metzger, W.K. The roles of carrier concentration and interface, bulk, and grain-boundary recombination for 25% efficient CdTe solar cells. J. Appl. Phys. 2017, 121, 214506. [Google Scholar] [CrossRef]
  36. Ristov, M.; Sinadinovski, G.; Grozdanov, I.; Mitreski, M. Chemical deposition of Tin(II) sulphide thin films. Thin Solid Films 1989, 173, 53–58. [Google Scholar] [CrossRef]
  37. Acharya, S.; Srivastava, O.N. Electronic behaviour of SnS2 crystals. Phys. Status Solidi 1981, 65, 717–723. [Google Scholar] [CrossRef]
  38. Varkey, A.J. Preparation of tin disulphide thin films by solution growth. Int. J. Mater. Prod. Technol. 1997, 12, 490–495. [Google Scholar] [CrossRef]
  39. Shi, C.; Chen, Z.; Shi, G.; Sun, R.; Zhan, X.; Shen, X. Influence of annealing on characteristics of tin disulfide thin films by vacuum thermal evaporation. Thin Solid Films 2012, 520, 4898–4901. [Google Scholar] [CrossRef]
  40. Fadavieslam, M.R.; Shahtahmasebi, N.; Rezaee-Roknabadi, M.; Bagheri-Mohagheghi, M.M. Effect of deposition conditions on the physical properties of SnxSy thin films prepared by the spray pyrolysis technique. J. Semicond. 2011, 32, 113002. [Google Scholar] [CrossRef]
  41. Ramakrishna Reddy, K.T.; Sreedevi, G.; Ramya, K.; Miles, R. Physical properties of nano-crystalline SnS2 layers grown by chemical bath deposition. Energy Procedia 2012, 15, 340–346. [Google Scholar] [CrossRef] [Green Version]
  42. Ghorpade, U.; Suryawanshi, M.; Shin, S.W.; Gurav, K.; Patil, P.; Pawar, S.; Hong, C.W.; Kim, J.H.; Kolekar, S. Towards environmentally benign approaches for the synthesis of CZTSSe nanocrystals by a hot injection method: A status review. Chem. Commun. 2014, 50, 11258. [Google Scholar] [CrossRef]
  43. Avellaneda, D.; Nair, M.T.S.; Nair, P.K. Polymorphic tin sulfide thin films of zinc blende and orthorhombic structures by chemical deposition. J. Electrochem. Soc. 2008, 155, D517. [Google Scholar] [CrossRef]
  44. Gode, F.; Guneri, E.; Baglayan, O. Effect of tri-sodium citrate concentration on structural, optical and electrical properties of chemically deposited tin sulfide films. Appl. Surf. Sci. 2014, 318, 227–233. [Google Scholar] [CrossRef]
  45. Koteeswara Reddy, N.; Ramakrishna Reddy, K.T. Electrical properties of spray pyrolytic tin sulfide films. Solid State Electron. 2005, 49, 902–906. [Google Scholar] [CrossRef]
  46. 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] [Green Version]
  47. Wangperawong, A.; Herron, S.M.; Runser, R.R.; Hagglund, C.; Tanskanen, J.T.; Lee, H.B.R.; Clemens, B.M.; Bent, S.F. Vapor transport deposition and epitaxy of orthorhombic SnS on glass and NaCl substrates. Appl. Phys. Lett. 2013, 103, 052105. [Google Scholar] [CrossRef]
  48. Djessas, K.; Masse, G. SnS thin films grown by close-spaced vapor transport. J. Mater. Sci. Lett. 2000, 19, 2135–2137. [Google Scholar] [CrossRef]
  49. Albers, W.; Haas, C.; Vink, H.J.; Wasscher, J.D. Investigations on SnS. J. Appl. Phys. 1961, 32, 2220–2225. [Google Scholar] [CrossRef]
  50. Chalapathi, U.; Poornaprakash, B.; Park, S.H. Chemically deposited cubic SnS thin films for solar cell applications. Sol. Energy 2016, 139, 238–248. [Google Scholar] [CrossRef]
  51. Skelton, J.M.; Burton, L.A.; Jackson, A.J.; Oba, F.; Parker, S.C.; Walsh, A. Lattice dynamics of the tin sulphides SnS2, SnS and Sn2S3: Vibrational spectra and thermal transport. Phys. Chem. Chem. Phys. 2017, 19, 12452–12465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kherchachi, I.B.; Attaf, A.; Saidi, H.; Bouhdjar, A.; Bendjdidi, H.; Youcef, B.; Azizi, R. The synthesis, characterization and phase stability of tin sulfides (SnS2, SnS and Sn2S3) films deposited by ultrasonic spray. Main Group Chem. 2016, 15, 231–242. [Google Scholar] [CrossRef]
  53. Chen, B.; Xu, X.; Wang, F.; Liu, J.; Ji, J. Electrochemical preparation and characterization of three-dimensional nanostructured Sn2S3 semiconductor films with nanorod network. Mater. Lett. 2011, 65, 400–402. [Google Scholar] [CrossRef]
  54. Julien, C.; Eddrief, M.; Samaras, I.; Balkanski, M. Optical and electrical characterizations of SnSe, SnS2 and SnSe2 single crystals. Mater. Sci. Eng. B 1992, 15, 70–72. [Google Scholar] [CrossRef]
  55. Madelung, O. IV-VII2 compounds. In Semiconductors: Data Handbook; Springer: Berlin/Heidelberg, Germany, 2004; pp. 606–612. [Google Scholar]
  56. Sreedevi, G.; Vasudeva Reddy, M.R.; Babu, P.; Chinho, P.; Chan Wook, J.; Ramakrishna Reddy, K.T. Studies on chemical bath deposited SnS2 films for Cd-free thin film solar cells. Ceram. Int. 2017, 43, 3713–3719. [Google Scholar] [CrossRef]
  57. Chen, S.; Gong, X.G.; Walsh, A.; Wei, S.-H. Defect physics of the kesterite thin-film solar cell absorber Cu2ZnSnS4. Appl. Phys. Lett. 2010, 96, 021902. [Google Scholar] [CrossRef]
  58. Ettema, A.R.H.F.; de Groot, R.A.; Haas, C.; Turner, T.S. Electronic structure of SnS deduced from photoelectron spectra and band-structure calculations. Phys. Rev. B 1992, 46, 7363–7373. [Google Scholar] [CrossRef] [Green Version]
  59. Lippens, P.E.; El Khalifi, M.; Womes, M. Electronic structures of SnS and SnS2. Phys. Status Solidi 2016, 254, 1600194. [Google Scholar] [CrossRef]
  60. Skelton, J.M.; Burton, L.A.; Oba, F.; Walsh, A. Chemical and lattice stability of the tin sulfides. J. Phys. Chem. C 2017, 121, 6446–6454. [Google Scholar] [CrossRef] [PubMed]
  61. Lindwall, G.; Shang, S.; Kelly, N.R.; Anderson, T.; Liu, Z.K. Thermodynamics of the S-Sn system: Implication for synthesis of earth abundant photovoltaic absorber materials. Sol. Energy 2016, 125, 314–323. [Google Scholar] [CrossRef] [Green Version]
  62. Wang, W.; Winkler, M.T.; Gunawan, O.; Gokmen, T.; Todorov, T.K.; Zhu, Y.; Mitzi, D.B. Device characteristics of CZTSSe thin-film solar cells with 12.6% efficiency. Adv. Energy Mater. 2014, 4, 1–5. [Google Scholar] [CrossRef]
  63. Loferski, J.J. Theoretical considerations governing the choice of the optimum semiconductor for photovoltaic solar energy conversion. J. Appl. Phys. 1956, 27, 777–784. [Google Scholar] [CrossRef]
  64. Skelton, J.M.; Burton, L.A.; Oba, F.; Walsh, A. Metastable cubic tin sulfide: A novel phonon-stable chiral semiconductor. APL Mater. 2017, 5, 036101. [Google Scholar] [CrossRef]
  65. Herzenbergite. Available online: https://www.mindat.org/min-1880.html (accessed on 13 February 2018).
  66. Ottemannite: Ottemannite Mineral Information and Data. Available online: https://www.mindat.org/min-3042.html (accessed on 13 February 2018).
  67. Berndtite: Berndtite Mineral Information and Data. Available online: https://www.mindat.org/min-637.html (accessed on 13 February 2018).
  68. Anthony, J.W.; Bideaux, R.A.; Bladh, K.W.; Nichols, M.C. Handbook of Mineralogy-Borates, Carbonates, Sulfates; Mineral Data Publishing: Tucson, AZ, USA, 2004; Volume V, p. 221. [Google Scholar]
  69. Eastaugh, N.; Walsh, V.; Chaplin, T.; Ruth, S. Pigment Compendium: A Dictionary and Optical Microscopy of Historical Pigments; Butterworth-Heinemann: Oxfordshire, UK, 2008; ISBN 0750689803. [Google Scholar]
  70. Moh, G.H.; Berndt, F. Two new natural tin sulfides, Sn2S3 and SnS2. New Yearb. Mineral. Monatshefte 1964, 4, 94–95. [Google Scholar]
  71. Yao, K.; Li, J.; Shan, S.; Jia, Q. One-step synthesis of urchinlike SnS/SnS2 heterostructures with superior visible-light photocatalytic performance. Catal. Commun. 2017, 101, 51–56. [Google Scholar] [CrossRef]
  72. Jing, J.; Cao, M.; Wu, C.; Huang, J.; Lai, J.; Sun, Y.; Wang, L.; Shen, Y. Chemical bath deposition of SnS nanosheet thin films for FTO/SnS/CdS/Pt photocathode. J. Alloys Compd. 2017, 726, 720–728. [Google Scholar] [CrossRef]
  73. Zhang, F.; Zhang, Y.; Zhang, G.; Yang, Z.; Dionysiou, D.D.; Zhu, A. Exceptional synergistic enhancement of the photocatalytic activity of SnS2 by coupling with polyaniline and N-doped reduced graphene oxide. Appl. Catal. B Environ. 2018, 236, 53–63. [Google Scholar] [CrossRef]
  74. Sun, Y.; Cheng, H.; Gao, S.; Sun, Z.; Liu, Q.; Leu, Q.; Lei, F.; Yao, T.; He, J.; Wei, S.; et al. Freestanding tin disulfide single-layers realizing efficient visible-light water splitting. Angew. Chem. Int. Ed. 2012, 51, 8727–8731. [Google Scholar] [CrossRef]
  75. Chauhan, H.; Singh, M.K.; Kumar, P.; Hashmi, S.A.; Deka, S. Development of SnS2/RGO nanosheet composite for cost-effective aqueous hybrid supercapacitors. Nanotechnology 2017, 28, 025401. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, Y.; Huang, L.; Wei, Z. Photoresponsive field-effect transistors based on multilayer SnS2 nanosheets. J. Semicond. 2017, 38, 034001. [Google Scholar] [CrossRef]
  77. Wu, Q.; Jiao, L.; Du, J.; Yang, J.; Guo, L.; Liu, Y.; Wang, Y. One-pot synthesis of three-dimensional SnS2 hierarchitectures as anode material for lithium-ion batteries. J. Power Source 2013, 239, 89–93. [Google Scholar] [CrossRef]
  78. Li, H.; Zhou, M.; Li, W.; Wang, K.; Cheng, S.; Jiang, K. Layered SnS2 cross-linked by carbon nanotubes as a high performance anode for sodium ion batteries. RSC Adv. 2016, 6, 35197–35202. [Google Scholar] [CrossRef]
  79. Ou, J.Z.; Ge, W.; Carey, B.; Daeneke, T.; Rotbart, A.; Shan, W.; Wang, Y.; Fu, Z.; Chrimes, A.F.; Wlodarski, W.; et al. Physisorption-based charge transfer in two-dimensional SnS2 for selective and reversible NO2 gas sensing. ACS Nano 2015, 9, 10313–10323. [Google Scholar] [CrossRef] [PubMed]
  80. Sanchez-Juarez, A.; Tiburcio-Silver, A.; Ortiz, A. Fabrication of SnS2/SnS heterojunction thin film diodes by plasma-enhanced chemical vapor deposition. Thin Solid Films 2005, 480, 452–456. [Google Scholar] [CrossRef]
  81. Fan, C.; Li, Y.; Lu, F.Y.; Deng, H.X.; Wei, Z.M.; Li, J.B. Wavelength dependent UV-Vis photodetectors from SnS2 flakes. RSC Adv. 2016, 6, 422–427. [Google Scholar] [CrossRef]
  82. Singh, D.J. Optical and electronic properties of semiconducting Sn2S3. Appl. Phys. Lett. 2016, 109, 1–5. [Google Scholar] [CrossRef]
  83. Li, Y.; Xie, H.; Tu, J. Nanostructured SnS/carbon composite for supercapacitor. Mater. Lett. 2009, 63, 1785–1787. [Google Scholar] [CrossRef]
  84. Saito, W.; Hayashi, K.; Nagai, H.; Miyazaki, Y. Preparation and thermoelectric properties of mixed valence compound Sn2S3. Jpn. J. Appl. Phys. 2017, 56, 061201. [Google Scholar] [CrossRef]
  85. Devika, M.; Koteeswara Reddy, N.; Sreekantha Reddy, D.; Ramesh, K.; Gunasekhar, K.; Raja Gopal, E.; Sung Ha, P. Metal–insulator–semiconductor field-effect transistors (MISFETs) using p-type SnS and nanometer-thick Al2S3 layers. RSC Adv. 2017, 7, 11111–11117. [Google Scholar] [CrossRef] [Green Version]
  86. Wang, W.; Shi, L.; Lan, D.; Li, Q. Improving cycle stability of SnS anode for sodium-ion batteries by limiting Sn agglomeration. J. Power Source 2018, 377, 1–6. [Google Scholar] [CrossRef]
  87. Lian, Q.; Zhou, G.; Liu, J.; Wu, C.; Wei, W.; Chen, L.; Li, C. Extrinsic pseudocapacitve Li-ion storage of SnS anode via lithiation-induced structural optimization on cycling. J. Power Source 2017, 366, 1–8. [Google Scholar] [CrossRef]
  88. Afsar, M.F.; Rafiq, M.A.; Tok, A.I.Y. Two-dimensional SnS nanoflakes: Synthesis and application to acetone and alcohol sensors. RSC Adv. 2017, 7, 21556–21566. [Google Scholar] [CrossRef] [Green Version]
  89. Chung, R.J.; Wang, A.N.; Peng, S.Y. An enzymatic glucose sensor composed of carbon-coated nano tin sulfide. Nanomaterials 2017, 7, 39. [Google Scholar] [CrossRef] [Green Version]
  90. Wang, C.; Chen, Y.; Jiang, J.; Zhang, R.; Niu, Y.; Zhou, T.; Xia, J.; Tian, H.; Hu, J.; Yang, P. Improved thermoelectric properties of SnS synthesized by chemical precipitation. RSC Adv. 2017, 7, 16795–16800. [Google Scholar] [CrossRef] [Green Version]
  91. Jayalakshmi, M.; Mohan Rao, M.; Choudary, B.M. Identifying nano SnS as a new electrode material for electrochemical capacitors in aqueous solutions. Electrochem. Commun. 2004, 6, 1119–1122. [Google Scholar] [CrossRef]
  92. Nozaki, H.; Fukano, T.; Ohta, S.; Seno, Y.; Katagiri, H.; Jimbo, K. Crystal structure determination of solar cell materials: Cu2ZnSnS4 thin films using X-ray anomalous dispersion. J. Alloys Compd. 2012, 524, 22–25. [Google Scholar] [CrossRef]
  93. Burton, L.A.; Walsh, A. Phase stability of the earth-abundant tin sulfides SnS, SnS2, and Sn2S3. J. Phys. Chem. C 2012, 116, 24262–24267. [Google Scholar] [CrossRef] [Green Version]
  94. Ahmet, I.Y.; Hill, M.S.; Johnson, A.L.; Peter, L.M. polymorph-selective deposition of high purity SnS thin films from a single source precursor. Chem. Mater. 2015, 27, 7680–7688. [Google Scholar] [CrossRef] [Green Version]
  95. Ehm, L.; Knorr, K.; Dera, P.; Krimmel, A.; Bouvier, P.; Mezouar, M. Pressure-induced structural phase transition in the IV–VI semiconductor SnS. J. Phys. Condens. Matter 2004, 16, 3545–3554. [Google Scholar] [CrossRef]
  96. Brownson, J.R.S.; Georges, C.; Larramona, G.; Jacob, A.; Delatouche, B.; Levy-Clement, C. Chemistry of tin monosulfide (δ-SnS) electrodeposition. J. Electrochem. Soc. 2008, 155, D40. [Google Scholar] [CrossRef]
  97. Brownson, J.R.S.; Georges, C.; Levy-Clement, C. Synthesis of a δ-SnS polymorph by electrodeposition. Chem. Mater. 2006, 18, 6397–6402. [Google Scholar] [CrossRef]
  98. Mariano, A.N.; Chopra, K.L. Polymorphism in some IV-VI compounds induced by high pressure and thin-film epitaxial growth. Appl. Phys. Lett. 1967, 10, 282–284. [Google Scholar] [CrossRef]
  99. Rabkin, A.; Samuha, S.; Abutbul, R.E.; Ezersky, V.; Meshi, L.; Golan, Y. New nanocrystalline materials: A previously unknown simple cubic phase in the SnS binary system. Nano Lett. 2015, 15, 2174–2179. [Google Scholar] [CrossRef]
  100. Biacchi, A.J.; Vaughn, D.D.; Schaak, R.E. Synthesis and crystallographic analysis of shape-controlled SnS nanocrystal photocatalysts: Evidence for a pseudotetragonal structural modification. J. Am. Chem. Soc. 2013, 135, 11634–11644. [Google Scholar] [CrossRef] [PubMed]
  101. Wang, X.; Liu, Z.; Zhao, X.-G.; Lv, J.; Biswas, K.; Zhang, L. Computational design of mixed-valence tin sulfides as solar absorbers. ACS Appl. Mater. Interfaces 2019, 11, 24867–24875. [Google Scholar] [CrossRef] [PubMed]
  102. Sugaki, A.; Kitakaze, A.; Kitazawa, H. Synthesized tin and tin-silver sulfide minerals: Synthetic sulfide minerals (XIII). Sci. Rep. 1985, 3, 199–211. [Google Scholar]
  103. Guenter, J.R.; Oswald, H. New polytype form of tin (IV) sulfide. Nat. Sci. 1968, 55, 171–177. [Google Scholar] [CrossRef]
  104. Mosburg, S.; Ross, D.R.; Bethke, P.M.; Toulmin, P. X-ray powder data for herzenbergite, teallite and tin trisulfide. US Geol. Surv. Prof. Pap. C 1961, 424, 347. [Google Scholar]
  105. Babu, P.; Vasudeva Reddy, M.; Sreedevi, G.; Chinho, P. Review on earth-abundant and environmentally benign Cu–Sn–X(X = S, Se) nanoparticles by chemical synthesis for sustainable solar energy conversion. J. Ind. Eng. Chem. 2018, 60, 19–52. [Google Scholar] [CrossRef]
  106. Nikolic, P.M.; Mihajlovic, P.; Lavrencic, B. Splitting and coupling of lattice modes in the layer compound SnS. J. Phys. C Solid State Phys. 1977, 10, L289–L292. [Google Scholar] [CrossRef]
  107. Wiley, J.D.; Buckel, W.J.; Schmidt, R.L. Infrared reflectivity and Raman scattering in GeS. Phys. Rev. B 1976, 13, 2489–2496. [Google Scholar] [CrossRef]
  108. Smith, A.J.; Meek, P.E.; Liang, W.Y. Raman scattering studies of SnS2 and SnSe2. J. Phys. C Solid State Phys. 1977, 10, 1321–1323. [Google Scholar] [CrossRef]
  109. Chandrasekhar, H.R.; Mead, D.G. Long-wavelength phonons in mixed-valence semiconductor SnIISnIVS3. Phys. Rev. B 1979, 19, 932–937. [Google Scholar] [CrossRef]
  110. Gedi, S.; Reddy, V.R.M.; Kang, J.; Jeon, C.-W. Impact of high temperature and short period annealing on SnS films deposited by E-beam evaporation. Appl. Surf. Sci. 2017, 402, 463–468. [Google Scholar] [CrossRef]
  111. Abutbul, R.E.; Segev, E.; Zeiri, L.; Ezersky, V.; Makov, G.; Golan, Y. Synthesis and properties of nanocrystalline π-SnS—A new cubic phase of tin sulphide. RSC Adv. 2016, 6, 5848–5855. [Google Scholar] [CrossRef]
  112. Bharatula, L.D.; Erande, M.B.; Mulla, I.S.; Rout, C.S.; Late, D.J. SnS2 nanoflakes for efficient humidity and alcohol sensing at room temperature. RSC Adv. 2016, 6, 105421–105427. [Google Scholar] [CrossRef]
  113. Reddy, T.S.; Kumar, M.C.S. Effect of substrate temperature on the physical properties of co-evaporated Sn2S3 thin films. Ceram. Int. 2016, 42, 12262–12269. [Google Scholar] [CrossRef]
  114. Kumagai, Y.; Burton, L.A.; Walsh, A.; Oba, F. Electronic structure and defect physics of tin sulfides: SnS, Sn2S3, and SnS2. Phys. Rev. Appl. 2016, 6, 1–14. [Google Scholar] [CrossRef] [Green Version]
  115. Burton, L.A.; Colombara, D.; Abellon, R.D.; Grozema, F.C.; Peter, L.M.; Savenije, T.J.; Dennler, G.; Walsh, A. Synthesis, characterization, and electronic structure of single-crystal SnS, Sn2S3, and SnS2. Chem. Mater. 2013, 25, 4908–4916. [Google Scholar] [CrossRef]
  116. Vasudeva Reddy, M.R.; Mohan Reddy, P.; Phaneendra Reddy, G.; Sreedevi, G.; Kishore Kumar, Y.B.R.; Babu, P.; Woo Kyoung, K.; Ramakrishna Reddy, K.T.; Chinho, P. Review on Cu2SnS3, Cu3SnS4, and Cu4SnS4 thin films and their photovoltaic performance. J. Ind. Eng. Chem. 2019, 76, 39–74. [Google Scholar] [CrossRef]
  117. Khoa, D.Q.; Nguyen, C.V.; Phuc, H.V.; Ilyasov, V.V.; Vu, T.V.; Cuong, N.Q.; Hoi, B.D.; Lu, D.V.; Feddi, E.; El-Yadri, M.; et al. Effect of strains on electronic and optical properties of monolayer SnS: Ab-initio study. Phys. B Condens. Matter 2018, 545, 255–261. [Google Scholar] [CrossRef]
  118. Segev, E.; Abutbul, R.E.; Argaman, U.; Golan, Y.; Makov, G. Surface energies and nanocrystal stability in the orthorhombic and π-phases of tin and germanium monochalcogenides. CrystEngComm 2018, 20, 4237–4248. [Google Scholar] [CrossRef]
  119. Qiu, G.; Zhang, H.; Liu, Y.; Xia, C. Strain effect on the electronic properties of Ce-doped SnS2 monolayer. Phys. B Condens. Matter 2018, 547, 1–5. [Google Scholar] [CrossRef]
  120. Ullah, H.; Noor-A-Alam, M.; Kim, H.J.; Shin, Y.H. Influences of vacancy and doping on electronic and magnetic properties of monolayer SnS. J. Appl. Phys. 2018, 124. [Google Scholar] [CrossRef]
  121. Zandalazini, C.I.; Navarro Sanchez, J.; Albanesi, E.A.; Gupta, Y.; Arun, P. Contribution of lattice parameter and vacancies on anisotropic optical properties of tin sulphide. J. Alloys Compd. 2018, 746, 9–18. [Google Scholar] [CrossRef]
  122. Akhoundi, E.; Faghihnasiri, M.; Memarzadeh, S.; Firouzian, A.H. Mechanical and strain-tunable electronic properties of the SnS monolayer. J. Phys. Chem. Solids 2019, 126, 43–54. [Google Scholar] [CrossRef]
  123. Sreedevi, G.; Vasudeva Reddy, M.; Babu, P.; Chan-Wook, J.; Chinho, P.; Ramakrishna Reddy, K.T. A facile inexpensive route for SnS thin film solar cells with SnS2 buffer. Appl. Surf. Sci. 2016, 372, 116–124. [Google Scholar] [CrossRef]
  124. Srivind, J.; Nagarethinam, V.S.; Balu, A.R. Optimization of S:Sn precursor molar concentration on the physical properties of spray deposited single phase Sn2S3 thin films. Mater. Sci. Pol. 2016, 34, 393–398. [Google Scholar] [CrossRef] [Green Version]
  125. Von Schnering, H.G.; Wiedemeier, H. The high temperature structure of ß-SnS and ß-SnSe and the B16-to-B33 type λ-transition path. J. Crystallogr. Cryst. Mater. 1981, 156, 143–150. [Google Scholar] [CrossRef]
  126. Hazen RM and Finger LW The crystal structres and compressibilities of layer minerals at high pressure. I SnS2, berndite. Am. Mineral. 1978, 63, 289–292. [Google Scholar]
  127. Park, H.K.; Jo, J.; Hong, H.K.; Song, G.Y.; Heo, J. Structural, optical, and electrical properties of tin sulfide thin films grown with electron-beam evaporation. Curr. Appl. Phys. 2015, 15, 964–969. [Google Scholar] [CrossRef]
  128. Nair, M.T.S.; Nair, P.K. Simplified chemical deposition technique for good quality SnS thin films. Semicond. Sci. Technol 1991, 6, 132–134. [Google Scholar] [CrossRef]
  129. Price, L.S.; Parkin, I.P.; Hardy, A.M.E.; Clark, R.J.H.; Hibbert, T.G.; Molloy, K.C. Atmospheric pressure chemical vapor deposition of tin sulfides (SnS, Sn2S3, and SnS2) on glass. Chem. Mater. 1999, 11, 1792–1799. [Google Scholar] [CrossRef]
  130. Lokhande, C.D. A chemical method for tin disulphide thin film deposition. J. Phys. D Appl. Phys. 1990, 23, 1703–1705. [Google Scholar] [CrossRef]
  131. Li, J.; Zhang, Y.C.; Zhang, M. Preparation of SnS2 thin films by chemical bath deposition. Mater. Sci. Forum 2010, 663, 104–107. [Google Scholar] [CrossRef]
  132. Thangaraju, B.; Kaliannan, P. Spray pyrolytic deposition and characterization of SnS and SnS2 thin films. J. Phys. D Appl. Phys. 2000, 33, 1054–1059. [Google Scholar] [CrossRef]
  133. Gopalakrishnan, P.; Anbazhagan, G.; Vijayarajasekaran, J.; Vijayakumar, K. Effect of substrate temperature on tin disulphide thin films. Int. J. Thin Film. Sci. Technol. 2017, 6, 73–75. [Google Scholar] [CrossRef]
  134. Tomas, R.; Lukas, S.; Libor, D.; Marek, B.; Milan, V.; Ludvik, B.; Tomas, W.; Roman, J. SnS and SnS2 thin films deposited using a spin-coating technique from intramolecularly coordinated organotin sulfides. Appl. Organomet. Chem. 2015, 29, 176–180. [Google Scholar] [CrossRef]
  135. Anitha, N.; Anitha, M.; Amalraj, L. Influence of precursor solution volume on the properties of tin disulphide (SnS2) thin films prepared by nebulized spray pyrolysis technique. Opt. Int. J. Light Electron Opt. 2017, 148, 28–38. [Google Scholar] [CrossRef]
  136. Vijayakumar, K.; Sanjeeviraja, C.; Jayachandran, M.; Amalraj, L. Characterization of tin disulphide thin films prepared at different substrate temperature using spray pyrolysis technique. J. Mater. Sci. Mater. Electron. 2011, 22, 929–935. [Google Scholar] [CrossRef]
  137. Sanchez-Juarez, A.; Ortz, A. Effects of precursor concentration on the optical and electrical properties of SnxSy thin films prepared by plasma-enhanced chemical vapour deposition. Semicond. Sci. Technol. 2002, 17, 931–937. [Google Scholar] [CrossRef]
  138. Amalraj, L.; Sanjeeviraja, C.; Jayachandran, M. Spray pyrolysised tin disulphide thin film and characterisation. J. Cryst. Growth 2002, 234, 683–689. [Google Scholar] [CrossRef]
  139. Ragina, A.J.; Murali, K.V.; Preetha, K.C.; Deepa, K.; Remadevi, T.L. UV irradiated wet chemical deposition and characterization of nanostructured tin sulfide thin films. J. Mater. Sci. Mater. Electron. 2012, 23, 2264–2271. [Google Scholar] [CrossRef]
  140. Joshua Gnanamuthu, S.; Johnson Jeyakumar, S.; Kartharinal Punithavathy, I.; Jobe Prabhakar, P.C.; Suganya, M.; Usharani, K.; Balu, A.R. Properties of spray deposited nano needle structured Cu-doped Sn2S3 thin films towards photovoltaic applications. Optik 2016, 127, 3999–4003. [Google Scholar] [CrossRef]
  141. Khadraoui, M.; Benramdane, N.; Mathieu, C.; Bouzidi, A.; Miloua, R.; Kebbab, Z.; Sahraoui, K.; Desfeux, R.; Wang, M. Optical and electrical properties of Sn2S3 thin films grown by spray pyrolysis. Solid State Commun. 2010, 150, 297–300. [Google Scholar] [CrossRef]
  142. Mnari, M.; Kamoun, N.; Bonnet, J.; Dachraoui, M. Chemical bath deposition of tin sulphide thin films in acid solution. Comptes Rendus Chim. 2009, 12, 824–827. [Google Scholar] [CrossRef]
  143. David Smyth-Boyle’s Homepage. Available online: http://www.ch.ic.ac.uk/obrien/barton/dsb/dsbcbd.html (accessed on 14 February 2018).
  144. Sreedevi, G.; Vasudeva Reddy, M.; Ramakrishna Reddy, K.T.; Soo Hyun, K.; Chan Wook, J. Chemically synthesized Ag-doped SnS films for PV applications. Ceram. Int. 2016, 42, 19027–19035. [Google Scholar] [CrossRef]
  145. Pramanik, P.; Basu, P.K.; Biswas, S. Preparation and characterization of chemically deposited tin(II) sulphide thin films. Thin Solid Films 1987, 150, 269–276. [Google Scholar] [CrossRef]
  146. Garcia-Angelmo, A.R.; Nair, M.T.S.; Nair, P.K. Evolution of crystalline structure in SnS thin films prepared by chemical deposition. Solid State Sci. 2014, 30, 26–35. [Google Scholar] [CrossRef]
  147. Lokhande, C.D.; Bhad, V.V.; Dhumure, S.S. Conversion of tin disulphide into silver sulphide by a simple chemical method. J. Phys. D Appl. Phys. 1992, 25, 315–318. [Google Scholar] [CrossRef]
  148. Opasanont, B.; Baxter, J.B. Dynamic speciation modeling to guide selection of complexing agents for chemical bath deposition: Case study for ZnS thin films. Cryst. Growth Des. 2015, 15, 4893–4900. [Google Scholar] [CrossRef]
  149. Lokhande, C.D. Chemical deposition of metal chalcogenide thin films. Mater. Chem. Phys. 1991, 27, 1–43. [Google Scholar] [CrossRef]
  150. Chopra, K.L.; Kainthla, R.C.; Pandya, D.K.; Thakoor, A.P. Physics of Thin Films; Francombe, M.H., Vossen, J.L., Eds.; Academic Press: New York, NY, USA, 1982; Volume 12, p. 201. [Google Scholar]
  151. Chopra, K.L.; Kainthla, R.C.; Pandya, D.K.; Thakoor, A.P. Chemical solution deposition of inorganic films. In Physics of Thin Films; Elsevier: Amsterdam, The Netherlands, 1982; Volume 12, pp. 167–235. ISBN 0079-1970. [Google Scholar]
  152. Savadogo, O.; Mandal, K.C. Studies on new chemically deposited photoconducting antimony trisulphide thin films. Sol. Energy Mater. Sol. Cells 1992, 26, 117–136. [Google Scholar] [CrossRef]
  153. Licht, S.; Longo, K.; Peramunage, D.; Forouzan, F. Conductometric analysis of the second acid dissociation constant of H2S in highly concentrated aqueous media. J. Electroanal. Chem. Interfacial Electrochem. 1991, 318, 111–129. [Google Scholar] [CrossRef]
  154. Pentia, E.; Draghici, V.; Sarau, G.; Mereu, B.; Pintilie, L.; Sava, F.; Popescu, M. Structural, electrical, and photoelectrical properties of CdxPb1-xS thin films prepared by chemical bath deposition. J. Electrochem. Soc. 2004, 151, G729. [Google Scholar] [CrossRef]
  155. Wired Chemist, Solubility Product Constants, Ksp. Available online: http://www.wiredchemist.com/chemistry/data/solubility-product-constants (accessed on 30 April 2018).
  156. Vaxasoftware, Solubility Product Constants. Available online: http://www.vaxasoftware.com/doc_eduen/qui/ks.pdf (accessed on 30 April 2018).
  157. Engelken, R.D.; Ali, S.; Chang, L.N.; Brinkley, C.; Turner, K.; Hester, C. Study and development of a generic electrochemical ion-exchange process to form MxS optoelectronic materials from ZnS precursor films formed by chemical-precipitation solution deposition. Mater. Lett. 1990, 10, 264–274. [Google Scholar] [CrossRef]
  158. Nair, M.T.S.; Nair, P.K. SnS-CuxS thin film combination: A desirable solar control coating for architectural and automobile glazings. J. Phys. D Appl. Phys. 1991, 24, 450–453. [Google Scholar] [CrossRef]
  159. Nair, P.K.; Nair, M.T.S. Chemically deposited SnS-CuxS thin films with high solar absorptance: New approach to all-glass tubular solar collectors. J. Phys. D. Appl. Phys. 1991, 24, 83–87. [Google Scholar] [CrossRef]
  160. Nair, P.K.; Nair, M.T.S.; Campos, J.; Sanchez, A. SnS-SnO2 conversion of chemically deposited SnS thin films. Adv. Mater. Opt. Electron. 1992, 1, 117–121. [Google Scholar] [CrossRef]
  161. Nair, P.K. Photoconductive SnO2 thin films from thermal decomposition of chemically deposited SnS thin films. J. Electrochem. Soc. 1993, 140, 539. [Google Scholar] [CrossRef]
  162. Nair, P.K.; Nair, M.T.S.; Zingaro, R.A.; Meyers, E.A. XRD, XPS, optical and electrical studies on the conversion of SnS thin films to SnO2. Thin Solid Films 1994, 239, 85–92. [Google Scholar] [CrossRef]
  163. Ray, S.C.; Karanjai, M.K.; DasGupta, D. Structure and photoconductive properties of dip-deposited SnS and SnS2 thin films and their conversion to tin dioxide by annealing in air. Thin Solid Films 1999, 350, 72–78. [Google Scholar] [CrossRef]
  164. Ristov, M.; Sinadinovski, G.; Mitreski, M.; Ristova, M. Photovoltaic cells based on chemically deposited p-type SnS. Sol. Energy Mater. Sol. Cells 2001, 69, 17–24. [Google Scholar] [CrossRef]
  165. Tanusevski, A. Optical and photoelectric properties of SnS thin films prepared by chemical bath deposition. Semicond. Sci. Technol. 2003, 18, 501–505. [Google Scholar] [CrossRef]
  166. Nair, M.T.S.; Lopez-Mata, C.; GomezDaza, O.; Nair, P.K. Copper tin sulfide semiconductor thin films produced by heating SnS–CuS layers deposited from chemical bath. Semicond. Sci. Technol. 2003, 18, 755–759. [Google Scholar] [CrossRef]
  167. Niinobe, D.; Wada, Y. Controlled deposition of SnS into/onto SnO2 nano-particle film and application to photoelectrochemical cells. Bull. Chem. Soc. Jpn. 2006, 79, 495–497. [Google Scholar] [CrossRef]
  168. Avellaneda, D.; Delgado, G.; Nair, M.T.S.; Nair, P.K. Structural and chemical transformations in SnS thin films used in chemically deposited photovoltaic cells. Thin Solid Films 2007, 515, 5771–5776. [Google Scholar] [CrossRef]
  169. Ge, Y.H.; Guo, Y.Y.; Shi, W.M.; Qiu, Y.H.; Wei, G.P. Influence of In-doping on resistivity of chemical bath deposited SnS films. J. Shanghai Univ. 2007, 11, 403–406. [Google Scholar] [CrossRef]
  170. Hankare, P.P.; Jadhav, A.V.; Chate, P.A.; Rathod, K.C.; Chavan, P.A.; Ingole, S.A. Synthesis and characterization of tin sulphide thin films grown by chemical bath deposition technique. J. Alloys Compd. 2008, 463, 581–584. [Google Scholar] [CrossRef]
  171. Avellaneda, D.; Nair, M.T.S.; Nair, P.K. Photovoltaic structures using chemically deposited tin sulfide thin films. Thin Solid Films 2009, 517, 2500–2502. [Google Scholar] [CrossRef]
  172. Turan, E.; Kul, M.; Aybek, A.S.; Zor, M. Structural and optical properties of SnS semiconductor films produced by chemical bath deposition. J. Phys. D Appl. Phys. 2009, 42, 245408. [Google Scholar] [CrossRef]
  173. Mathews, N.R.; Avellaneda, D.; Anaya, H.B.M.; Campos, J.; Nair, M.T.S.; Nair, P.K. Chemically and electrochemically deposited thin films of tin sulfide for photovoltaic structures. Mater. Res. Soc. Symp. Proc. 2009, 1165, 367–373. [Google Scholar] [CrossRef]
  174. Aksay, S.; Ozer, T.; Zor, M. Vibrational and X-ray diffraction spectra of SnS film deposited by chemical bath deposition method. Eur. Phys. J. Appl. Phys. 2009, 47, 30502. [Google Scholar] [CrossRef]
  175. Wang, Y.; Reddy, Y.B.K.; Gong, H. Large-surface-area nanowall SnS films prepared by chemical bath deposition. J. Electrochem. Soc. 2009, 156, H157. [Google Scholar] [CrossRef]
  176. Akkari, A.; Guasch, C.; Kamoun, T.N. Chemically deposited tin sulphide. J. Alloys Compd. 2010, 490, 180–183. [Google Scholar] [CrossRef]
  177. Guneri, E.; Ulutas, C.; Kirmizigul, F.; Altindemir, G.; Gode, F.; Gumus, C. Effect of deposition time on structural, electrical, and optical properties of SnS thin films deposited by chemical bath deposition. Appl. Surf. Sci. 2010, 257, 1189–1195. [Google Scholar] [CrossRef]
  178. Guneri, E.; Gode, F.; Ulutas, C.; Kirmizigul, F.; Altindemir, G.; Gumus, C. Properties of p-type SnS thin films prepared by chemical bath deposition. Chalcogenide Lett. 2010, 7, 685–694. [Google Scholar]
  179. Gaied, I.; Akkari, A.; Yacoubi, N.; Kamoun, N. Influence of the triethanolamine concentration on the optical properties of tin sulphide thin films by the photothermal deflection spectroscopy. J. Phys. Conf. Ser. 2010, 214, 012128. [Google Scholar] [CrossRef]
  180. Wang, Y.; Gong, H.; Fan, B.; Hu, G. Photovoltaic behavior of nanocrystalline SnS/TiO2. J. Phys. Chem. C 2010, 114, 3256–3259. [Google Scholar] [CrossRef]
  181. Ragina, A.J.; Preetha, K.C.; Murali, K.V.; Deepa, K.; Remadevi, T.L. Wet chemical synthesis and characterization of tin sulphide thin films from different host solutions. Adv. Appl. Sci. Res. 2011, 2, 438–444. [Google Scholar]
  182. Kassim, A.; Min, H.S.; Sharif, A.; Haron, J.; Nagalingam, S. Chemical bath deposition of SnS thin films: AFM, EDAX and UV-Visible characterization. Orient. J. Chem. 2011, 27, 1375–1381. [Google Scholar]
  183. Kassim, A.; Min, H.S.; Shariff, A.; Haron, M.J. The effect of the pH value on the growth and properties of chemical bath deposited SnS thin films. Mater. Chem. Phys. 2011, 15, 45–48. [Google Scholar] [CrossRef]
  184. Gao, C.; Shen, H.; Sun, L.; Shen, Z. Chemical bath deposition of SnS films with different crystal structures. Mater. Lett. 2011, 65, 1413–1415. [Google Scholar] [CrossRef]
  185. Gao, C.; Shen, H.; Sun, L. Preparation and properties of zinc blende and orthorhombic SnS films by chemical bath deposition. Appl. Surf. Sci. 2011, 257, 6750–6755. [Google Scholar] [CrossRef]
  186. Herron, S.M.; Wangperawong, A.; Bent, S.F. Chemical bath deposition and microstructuring of tin (II) sulfide films for photovoltaics. In Proceedings of the 37th IEEE Photovoltaic Specialists Conference, Seattle, WA, USA, 19–24 June 2011; pp. 368–371. [Google Scholar]
  187. Gao, C.; Shen, H. Influence of the deposition parameters on the properties of orthorhombic SnS films by chemical bath deposition. Thin Solid Films 2012, 520, 3523–3527. [Google Scholar] [CrossRef]
  188. Xia, D.L.; Xu, J.; Shi, W.Q.; Lei, P.; Zhao, X.J. Synthesis and properties of SnS thin films by chemical bath deposition. Eng. Mater. 2012, 509, 333–338. [Google Scholar] [CrossRef]
  189. Martinez, H.; Avellaneda, D. Modifications in SnS thin films by plasma treatments. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2012, 272, 351–356. [Google Scholar] [CrossRef]
  190. Patel, T.H. Influence of deposition time on structural and optical properties of chemically deposited SnS thin films. Open Surf. Sci. J. 2012, 4, 6–13. [Google Scholar] [CrossRef] [Green Version]
  191. Jayasree, Y.; Chalapathi, U.; Uday Bhaskar, P.; Raja, V.S. Effect of precursor concentration and bath temperature on the growth of chemical bath deposited tin sulphide thin films. Appl. Surf. Sci. 2012, 258, 2732–2740. [Google Scholar] [CrossRef]
  192. Avellaneda, D.; Krishnan, B.; Das Roy, T.K.; Castillo, G.A.; Shaji, S. Modification of structure, morphology and physical properties of tin sulfide thin films by pulsed laser irradiation. Appl. Phys. A 2013, 110, 667–672. [Google Scholar] [CrossRef]
  193. Sreedevi, G.; Ramakrishna Reddy, K.T. Properties of tin monosulphide films grown by chemical bath deposition. Conf. Pap. Energy 2013, 2013, 528724. [Google Scholar] [CrossRef]
  194. Soonmin, H.; Kassim, A.; Weetee, T. Thickness dependent characteristics of chemically deposited tin sulfide films. Univ. J. Chem. 2013, 1, 170–174. [Google Scholar] [CrossRef]
  195. Osuwa, J.C.; Ugochukwu, J. Effects of aluminum and manganese impurity concentrations on optoelectronic properties of thin films of tin sulfide (SnS) using CBD method. IOSR J. Environ. Sci. Toxicol. Food Technol. 2013, 5, 32–36. [Google Scholar] [CrossRef]
  196. Onwuemeka, J.I.; Ezike, F.M.; Nwulu, N.C. The effect of annealing temperature and time on the optical properties of SnS thin films prepared by chemical bath deposition. Int. J. Innov. Educ. Res. 2013, 1, 121–130. [Google Scholar] [CrossRef]
  197. Reghima, M.; Akkari, A.; Guasch, C.; Castagné, M.; Kamoun-Turki, N. Synthesis and characterization of Fe-doped SnS thin films by chemical bath deposition technique for solar cells applications. J. Renew. Sustain. Energy 2013, 5, 063109. [Google Scholar] [CrossRef]
  198. Jayasree, Y.; Chalapathi, U.; Raja, V.S. Growth and characterization of tin sulphide thin films by chemical bath deposition using ethylene diamine tetra-acetic acid as the complexing agent. Thin Solid Films 2013, 537, 149–155. [Google Scholar] [CrossRef]
  199. Dhanya, A.C.; Deepa, K.; Geetanjali, P.M.; Anupama, M.; Remadevi, T.L. Effect of post deposition by UV irradiation on chemical bath deposited tin sulfide thin films. Appl. Phys. A 2014, 116, 1467–1472. [Google Scholar] [CrossRef]
  200. He, H.Y.; Fei, J.; Lu, J. Optical and electrical properties of pure and Sn4+-doped n-SnS films deposited by chemical bath deposition. Mater. Sci. Semicond. Process. 2014, 24, 90–95. [Google Scholar] [CrossRef]
  201. Safonova, M.; Nair, P.K.; Mellikov, E.; Garcia, A.R.; Kerm, K.; Revathi, N.; Romann, T.; Mikli, V.; Volobujeva, O. Chemical bath deposition of SnS thin films on ZnS and CdS substrates. J. Mater. Sci. Mater. Electron. 2014, 25, 3160–3165. [Google Scholar] [CrossRef]
  202. Avellaneda, D.; Krishnan, B.; Rodriguez, A.C.; Das Roy, T.K.; Shaji, S. Heat treatments in chemically deposited SnS thin films and their influence in CdS/SnS photovoltaic structures. J. Mater. Sci. Mater. Electron. 2015, 26, 5585–5592. [Google Scholar] [CrossRef]
  203. Joshi, L.P.; Risal, L.; Shrestha, S.P. Effects of concentration of triethanolamine and annealing temperature on band gap of thin film of tin sulphide prepared by chemical bath deposition method. J. Nepal Phys. Soc. 2016, 3, 1–5. [Google Scholar] [CrossRef] [Green Version]
  204. Safonova, M.; Mellikov, E.; Mikli, V.; Kerm, K.; Revathi, N.; Volobujeva, O. Chemical bath deposition of SnS thin films from the solutions with different concentrations of tin and sulphur. Adv. Mater. Res. 2015, 1117, 183–186. [Google Scholar] [CrossRef]
  205. Mahdi, M.S.; Ibrahim, K.; Hmood, A.; Ahmed, N.M.; Azzez, S.A.; Mustafa, F.I. A highly sensitive flexible SnS thin film photodetector in the ultraviolet to near infrared prepared by chemical bath deposition. RSC Adv. 2016, 6, 114980–114988. [Google Scholar] [CrossRef]
  206. Chaki, S.H.; Chaudhary, M.D.; Deshpande, M.P. SnS thin films deposited by chemical bath deposition, dip coating and SILAR techniques. J. Semicond. 2016, 37, 1–9. [Google Scholar] [CrossRef] [Green Version]
  207. Mahdi, M.S.; Ibrahim, K.; Hmood, A.; Ahmed, N.M.; Mustafa, F.I. Control of phase, structural and optical properties of tin sulfide nanostructured thin films grown via chemical bath deposition. J. Electron. Mater. 2017, 46, 4227–4235. [Google Scholar] [CrossRef]
  208. Avellaneda, D.; Sánchez-Orozco, I.; Martínez, J.A.A.; Shaji, S.; Krishnan, B. Thin films of tin sulfides: Structure, composition and optoelectronic properties. Mater. Res. Express 2018, 6, 016409. [Google Scholar] [CrossRef]
  209. Cao, M.; Wu, C.; Yao, K.; Jing, J.; Huang, J.; Cao, M.; Zhang, J.; Lai, J.; Ali, O.; Wang, L.; et al. Chemical bath deposition of single crystal SnS nanobelts on glass substrates. Mater. Res. Bull. 2018, 104, 244–249. [Google Scholar] [CrossRef]
  210. Patil, A.M.; Lokhande, V.C.; Patil, U.M.; Shinde, P.A.; Lokhande, C.D. High performance all-solid-state asymmetric supercapacitor device based on 3D nanospheres of β-MnO2 and Nanoflowers of O-SnS. ACS Sustain. Chem. Eng. 2018, 6, 787–802. [Google Scholar] [CrossRef]
  211. Gedi, S.; Minnam Reddy, V.R.; Kotte, T.R.R.; Park, Y.; Kim, W.K. Effect of C4H6O6 concentration on the properties of SnS thin films for solar cell applications. Appl. Surf. Sci. 2019, 465, 802–815. [Google Scholar] [CrossRef]
  212. Gedi, S.; Minnam Reddy, V.R.; Alhammadi, S.; Reddy Guddeti, P.; Kotte, T.R.R.; Park, C.; Kim, W.K. Influence of deposition temperature on the efficiency of SnS solar cells. Sol. Energy 2019, 184, 305–314. [Google Scholar] [CrossRef]
  213. Cabrera-German, D.; García-Valenzuela, J.A.; Cota-Leal, M.; Martínez-Gil, M.; Aceves, R.; Sotelo-Lerma, M. Detailed characterization of good-quality SnS thin films obtained by chemical solution deposition at different reaction temperatures. Mater. Sci. Semicond. Process. 2019, 89, 131–142. [Google Scholar] [CrossRef]
  214. Mallika Bramaramba Devi, P.; Phaneendra Reddy, G.; Ramakrishna Reddy, K.T. Structural and optical studies on PVA capped SnS films grown by chemical bath deposition for solar cell application. J. Semicond. 2019, 40, 052101. [Google Scholar] [CrossRef]
  215. Mahdi, M.S.; Hmood, A.; Ibrahim, K.; Ahmed, N.M.; Bououdina, M. Dependence of pH on phase stability, optical and photoelectrical properties of SnS thin films. Superlattices Microstruct. 2019, 128, 170–176. [Google Scholar] [CrossRef]
  216. Huang, J.; Ma, Y.; Yao, K.; Wu, C.; Cao, M.; Lai, J.; Zhang, J.; Sun, Y.; Wang, L.; Shen, Y. Chemical bath deposition of SnS:In thin films for Pt/CdS/SnS:In/Mo photocathode. Surf. Coat. Technol. 2019, 358, 84–90. [Google Scholar] [CrossRef]
  217. Gonzaez-Flores, V.E.; Mohan, R.N.; Ballinas-Morales, R.; Nair, M.T.S.; Nair, P.K. Thin film solar cells of chemically deposited SnS of cubic and orthorhombic structures. Thin Solid Films 2019, 672, 62–65. [Google Scholar] [CrossRef]
  218. Mallika Bramaramba Devi, P.; Reddy, G.P.; Reddy, K.T.R. Optical investigations on PVA capped SnS nanocrystalline films deposited by CBD process. Mater. Res. Express 2019, 6, 115523. [Google Scholar] [CrossRef]
  219. John, S.; Geetha, V.; Francis, M. Effect of pH on the optical and structural properties of SnS prepared by chemical bath deposition method. IOP Conf. Ser. Mater. Sci. Eng. 2020, 872, 012139. [Google Scholar] [CrossRef]
  220. Muthalif, M.P.A.; Choe, Y. Control of the interfacial charge transfer resistance to improve the performance of quantum dot sensitized solar cells with highly electrocatalytic Cu-doped SnS counter electrodes. Appl. Surf. Sci. 2020, 508, 145297. [Google Scholar] [CrossRef]
  221. Ben Mbarek, M.; Reghima, M.; Yacoubi, N.; Barradas, N.; Alves, E.; Bundaleski, N.; Teodoro, O.; Kunst, M.; Schwarz, R. Microwave transient reflection in annealed SnS thin films. Mater. Sci. Semicond. Process. 2021, 121, 105302. [Google Scholar] [CrossRef]
  222. Marquez, I.G.; Romano-Trujillo, R.; Gracia-Jimenez, J.M.; Galeazzi, R.; Silva-González, N.R.; García, G.; Coyopol, A.; Nieto-Caballero, F.G.; Rosendo, E.; Morales, C. Cubic, orthorhombic and amorphous SnS thin films on flexible plastic substrates by CBD. J. Mater. Sci. Mater. Electron. 2021, 1–9. [Google Scholar] [CrossRef]
  223. Cao, M.; Zhang, X.; Ren, J.; Sun, Y.; Cui, Y.; Zhang, J.; Ling, J.; Huang, J.; Shen, Y.; Wang, L.; et al. Chemical bath deposition of SnS:Cu/ZnS for solar hydrogen production and solar cells. J. Alloys Compd. 2021, 863, 158727. [Google Scholar] [CrossRef]
  224. Higareda-Sanchez, A.; Mis-Fernandez, R.; Rimmaudo, I.; Camacho-Espinosa, E.; Pena, 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]
  225. Akkari, A.; Regima, M.; Guasch, C.; Kamoun Turki, N. Effect of deposition time on physical properties of nanocrystallized SnS zinc blend thin films grown by chemical bath deposition. Adv. Mater. Res. 2011, 324, 101–104. [Google Scholar] [CrossRef]
  226. Akkari, A.; Guasch, C.; Castagne, M.; Kamoun-Turki, N. Optical study of zinc blend SnS and cubic In2S3:Al thin films prepared by chemical bath deposition. J. Mater. Sci. 2011, 46, 6285–6292. [Google Scholar] [CrossRef]
  227. Akkari, A.; Reghima, M.; Guasch, C.; Kamoun-Turki, N. Effect of copper doping on physical properties of nanocrystallized SnS zinc blend thin films grown by chemical bath deposition. J. Mater. Sci. 2012, 47, 1365–1371. [Google Scholar] [CrossRef]
  228. Reghima, M.; Akkari, A.; Guasch, C.; Kamoun-Turki, N. Effect of indium doping on physical properties of nanocrystallized SnS zinc blend thin films grown by chemical bath deposition. J. Renew. Sustain. Energy 2012, 4, 011602. [Google Scholar] [CrossRef]
  229. Reghima, M.; Akkari, A.; Guasch, C.; Kamoun, N. Structural, optical, and electrical properties of SnS:Ag thin films. J. Electron. Mater. 2015, 44, 4392–4399. [Google Scholar] [CrossRef]
  230. Barrios-Salgado, E.; Rodríguez-Guadarrama, L.A.; Garcia-Angelmo, A.R.; Campos Álvarez, J.; Nair, M.T.S.; Nair, P.K. Large cubic tin sulfide–tin selenide thin film stacks for energy conversion. Thin Solid Films 2016, 615, 415–422. [Google Scholar] [CrossRef]
  231. Nair, P.K.; Barrios-Salgado, E.; Nair, M.T.S. Cubic-structured tin selenide thin film as a novel solar cell absorber. Phys. Status Solidi Appl. Mater. Sci. 2016, 213, 2229–2236. [Google Scholar] [CrossRef]
  232. Abutbul, R.E.; Garcia-Angelmo, A.R.; Burshtein, Z.; Nair, M.T.S.; Nair, P.K.; Golan, Y. Crystal structure of a large cubic tin monosulfide polymorph: An unraveled puzzle. CrystEngComm 2016, 18, 5188–5194. [Google Scholar] [CrossRef] [Green Version]
  233. Sanal, K.C.; Nair, P.K.; Nair, M.T.S. Band offset in zinc oxy-sulfide/cubic-tin sulfide interface from X-ray photoelectron spectroscopy. Appl. Surf. Sci. 2017, 396, 1092–1097. [Google Scholar] [CrossRef]
  234. Mahdi, M.S.; Ibrahim, K.; Ahmed, N.M.; Hmood, A.; Mustafa, F.I.; Azzez, S.A.; Bououdina, M. High performance and low-cost UV–Visible–NIR photodetector based on tin sulphide nanostructures. J. Alloys Compd. 2018, 735, 2256–2262. [Google Scholar] [CrossRef]
  235. Javed, A.; Bashir, M. Controlled growth, structure and optical properties of Fe-doped cubic π-SnS thin films. J. Alloys Compd. 2018, 759, 14–21. [Google Scholar] [CrossRef]
  236. Flores, V.E.G.; Nair, M.T.S.; Nair, P.K. Thermal stability of “metastable” cubic tin sulfide and its relevance to applications. Semicond. Sci. Technol. 2018, 33. [Google Scholar] [CrossRef]
  237. Abutbul, R.E.; Golan, Y. Chemical epitaxy of π-phase cubic tin monosulphide. CrystEngComm 2020, 22, 6170–6181. [Google Scholar] [CrossRef]
  238. Mahdi, M.S.; Al-Arab, H.S.; Al-Salman, H.S.; Ibrahim, K.; Ahmed, N.M.; Hmood, A.; Bououdina, M. A high-performance near-infrared photodetector based on π-SnS phase. Mater. Lett. 2020, 273, 127910. [Google Scholar] [CrossRef]
  239. Javed, A.; Khan, N.; Bashir, S.; Ahmad, M.; Bashir, M. Thickness dependent structural, electrical and optical properties of cubic SnS thin films. Mater. Chem. Phys. 2020, 246, 122831. [Google Scholar] [CrossRef]
  240. Barrios Salgado, E.; Lara Llanderal, D.E.; Nair, M.T.S.; Nair, P.K. Thin film thermoelectric elements of p–n tin chalcogenides from chemically deposited SnS–SnSe stacks of cubic crystalline structure. Semicond. Sci. Technol. 2020, 35, 045006. [Google Scholar] [CrossRef]
  241. Chalapathi, U.; Poornaprakash, B.; Choi, W.J.; Park, S.-H. Ammonia(aq)-enhanced growth of cubic SnS thin films by chemical bath deposition for solar cell applications. Appl. Phys. A 2020, 126, 583. [Google Scholar] [CrossRef]
  242. Rodriguez-Guadarrama, L.A.; Escorcia-Garcia, J.; Alonso-Lemus, I.L.; Campos-Álvarez, J. Synthesis of π-SnS thin films through chemical bath deposition: Effects of pH, deposition time, and annealing temperature. J. Mater. Sci. Mater. Electron. 2021, 32, 7464–7480. [Google Scholar] [CrossRef]
  243. Gaitan-Arevalo, J.R.; Gonzalez, L.A.; Escorcia-García, J. Cubic tin sulfide thin films by a Sn-NTA system based chemical bath process. Mater. Lett. 2021, 286, 129222. [Google Scholar] [CrossRef]
  244. Eswar Neerugatti, K.; Shivaji Pawar, P.; Heo, J. Differential growth and evaluation of band structure of π-SnS for thin-film solar cell applications. Mater. Lett. 2021, 284, 129026. [Google Scholar] [CrossRef]
  245. Ramakrishna Reddy, K.T.; Sreedevi, G.; Miles, R.W. Thickness effect on the structural and optical properties of SnS2 films grown by CBD process. J. Mater. Sci. Eng. A 2013, 3, 182–186. [Google Scholar] [CrossRef]
  246. Chalapathi, U.; Poornaprakash, B.; Purushotham Reddy, B.; Si Hyun, P. Preparation of SnS2 thin films by conversion of chemically deposited cubic SnS films into SnS2. Thin Solid Films 2017, 640, 81–87. [Google Scholar] [CrossRef]
  247. Zhou, P.; Huang, Z.; Sun, X.; Ran, G.; Shen, R.; Ouyang, Q. Saturable absorption properties of the SnS2/FTO thin film. Optik 2018, 171, 839–844. [Google Scholar] [CrossRef]
  248. Joshi, M.P.; Khot, K.V.; Patil, S.S.; Mali, S.S.; Hong, C.K.; Bhosale, P.N. Investigating the light harvesting capacity of sulfur ion concentration dependent SnS2 thin films synthesized by self-assembled arrested precipitation technique. Mater. Res. Express 2019, 6, 086467. [Google Scholar] [CrossRef]
  249. Noppakuadrittidej, P.; Vailikhit, V.; Teesetsopon, P.; Choopun, S.; Tubtimtae, A. Copper incorporation in Mn2+ doped Sn2S3 nanocrystals and the resultant structural, optical, and electrochemical characteristics. Ceram. Int. 2018, 44, 13973–13985. [Google Scholar] [CrossRef]
  250. Mohan, R.N.; Nair, M.T.S.; Nair, P.K. Thin film Sn2S3 via chemical deposition and controlled heating—Its prospects as a solar cell absorber. Appl. Surf. Sci. 2020, 504, 144162. [Google Scholar] [CrossRef]
  251. Liu, T.; Ke, H.; Zhang, H.; Duo, S.; Sun, Q.; Fei, X.; Zhou, G.; Liu, H.; Fan, L. Effect of four different zinc salts and annealing treatment on growth, structural, mechanical and optical properties of nanocrystalline ZnS thin films by chemical bath deposition. Mater. Sci. Semicond. Process. 2014, 26, 301–311. [Google Scholar] [CrossRef]
  252. Salh, A.; Kyeongchan, M.; Hyeonwook, P.; Woo Kyoung, K. Effect of different cadmium salts on the properties of chemical-bath-deposited CdS thin films and Cu(InGa)Se2 solar cells. Thin Solid Films 2017, 625, 56–61. [Google Scholar] [CrossRef]
  253. Hodes, G. Chemical Solution Deposition of Semiconductor Films, 1st ed.; Marcel Dekker: New York, NY, USA, 2002; ISBN 0824708512. [Google Scholar]
  254. Jeffery, G.H.; Bassett, J.; Mendham, J.; Denney, R.C. Vogel’s Textbook of Quantitative Inorganic Analysis; ELBS Publ.: London, UK, 1989. [Google Scholar]
  255. Flaschka, H.A. EDTA Titrations. An Introduction to Theory and Practice; Macmillan: New York, NY, USA, 1959. [Google Scholar]
  256. Kaur, I. Growth kinetics and polymorphism of chemically deposited CdS films. J. Electrochem. Soc. 1980, 127, 943. [Google Scholar] [CrossRef]
  257. Donna, J.M. Chemical bath deposition of CdS thin films: Electrochemical in situ kinetic studies. J. Electrochem. Soc. 1992, 139, 2810. [Google Scholar] [CrossRef]
  258. Gonzalez Panzo, I.J.; Martin Varguez, P.E.; Oliva, A.I. Role of thiourea in the kinetic of growth of the chemical bath deposited ZnS films. J. Electrochem. Soc. 2014, 161, D761–D767. [Google Scholar] [CrossRef] [Green Version]
  259. Kumarage, W.G.C.; Wijesundara, L.B.D.R.P.; Seneviratne, V.A.; Jayalath, C.P.; Dassanayake, B.S. Influence of bath temperature on CBD-CdS thin films. Procedia Eng. 2016, 139, 64–68. [Google Scholar] [CrossRef] [Green Version]
  260. Bayon, R.; Hernandez-Mayoral, M.; Herrero, J. Growth mechanism of CBD-In(OH)xSy thin films. J. Electrochem. Soc. 2002, 149, C59–C67. [Google Scholar] [CrossRef]
  261. Cortes, A.; Gómez, H.; Marotti, R.E.; Riveros, G.; Dalchiele, E.A. Grain size dependence of the bandgap in chemical bath deposited CdS thin films. Sol. Energy Mater. Sol. Cells 2004, 82, 21–34. [Google Scholar] [CrossRef]
  262. Balasubramanian, V.; Suriyanarayanan, N.; Prabahar, S. Thickness-dependent structural properties of chemically deposited Bi2S3 thin films. Adv. Apllied Sci. Res. 2012, 3, 2369–2373. [Google Scholar]
  263. Patil, S.A.; Mengal, N.; Memon, A.A.; Jeong, S.H.; Kim, H.S. CuS thin film grown using the one pot, solution-process method for dye-sensitized solar cell applications. J. Alloys Compd. 2017, 708, 568–574. [Google Scholar] [CrossRef]
  264. Gupta, V.; Mansingh, A. Influence of postdeposition annealing on the structural and optical properties of sputtered zinc oxide film. J. Appl. Phys. 1996, 80, 1063–1073. [Google Scholar] [CrossRef]
  265. Jin Hong, L.; Byung Ok, P. Transparent conducting In2O3 thin films prepared by ultrasonic spray pyrolysis. Surf. Coat. Technol. 2004, 184, 102–107. [Google Scholar] [CrossRef]
  266. Brien, P.O.; Mcaleese, J. bath deposition of ZnS and CdS. Technology 1998, 8, 2309–2314. [Google Scholar] [CrossRef]
  267. Avrami, M. Granulation, phase change, and microstructure kinetics of phase change. III. J. Chem. Phys. 1941, 9, 177–184. [Google Scholar] [CrossRef]
  268. Phillips, J.M. Substrate selection for thin-film growth. MRS Bull. 1995, 20, 35–39. [Google Scholar] [CrossRef]
  269. Vasudeva Reddy, M.; Sreedevi, G.; Chinho, P.; Miles, R.W.; Ramakrishna Reddy, K.T. Development of sulphurized SnS thin film solar cells. Curr. Appl. Phys. 2015, 15, 588–598. [Google Scholar] [CrossRef]
  270. Devika, M.; Koteeswara Reddy, N.; Ramesh, K.; Ganesan, V.; Gopal, E.S.R.; Ramakrishna Reddy, K.T. Influence of substrate temperature on surface structure and electrical resistivity of the evaporated tin sulphide films. Appl. Surf. Sci. 2006, 253, 1673–1676. [Google Scholar] [CrossRef]
  271. Vasudeva Reddy, M.; Sreedevi, G.; Babu, P.; Ramakrishna Reddy, K.T.; Guillaume, Z.; Chinho, P. Influence of different substrates on the properties of sulfurized SnS films. Sci. Adv. Mater. 2016, 8, 247–251. [Google Scholar] [CrossRef]
  272. Camacho Espinosa, E.; Oliva Aviles, A.I.; Oliva, A.I. Effect of the substrate cleaning process on pinhole formation in sputtered CdTe films. J. Mater. Eng. Perform. 2017, 26, 4020–4028. [Google Scholar] [CrossRef]
  273. Tarkeshwar, S.; Devjyoti, L.; Ayush, K. Effects of various parameters on structural and optical properties of CBD-grown ZnS thin films: A review. J. Electron. Mater. 2018, 47, 1730–1751. [Google Scholar] [CrossRef]
  274. Hubert, C.; Naghavi, N.; Roussel, O.; Etcheberry, A.; Hariskos, D.; Menner, R.; Powalla, M.; Kerrec, O.; Lincot, D. The Zn(S,O,OH)/ZnMgO buffer in thin film Cu(In,Ga)(S,Se)2-based solar cells part I: Fast chemical bath deposition of Zn(S,O,OH) buffer layers for industrial application on co-evaporated Cu(In,Ga)Se2 and electrodeposited CuIn(S,Se)2 solar cells. Prog. Photovolt. Res. Appl. 2009, 17, 470–478. [Google Scholar] [CrossRef]
  275. Li, Z.Q.; Shi, J.H.; Liu, Q.Q.; Wang, Z.A.; Sun, Z.; Huang, S.M. Effect of [Zn]/[S] ratios on the properties of chemical bath deposited zinc sulfide thin films. Appl. Surf. Sci. 2010, 257, 122–126. [Google Scholar] [CrossRef]
Figure 1. Applications of SnxSy [4,5,6,7,8,9,10,11,12]. Solar cell (o-SnS, c-SnS, Sn2S3 as light absorber and SnS2 as buffer); photodetector (o-SnS, c-SnS, Sn2S3 as light absorber and SnS2 as buffer); Li- and Na-ion batteries (o-SnS, c-SnS, and SnS2 as anode materials); gas- and bio sensors (o-SnS, c-SnS, and SnS2 as sensing materials); tunnel field-effect transistors (TFET) (o-SnS, c-SnS, and SnS2 as top or back gates); electrochemical and super capacitors (o-SnS, c-SnS, and SnS2 as electrode materials); capacitor; thermoelectrics (o-SnS, c-SnS, and Sn2S3 as grids); and water-splitting (o-SnS, c-SnS, and SnS2 as photocathodes).
Figure 1. Applications of SnxSy [4,5,6,7,8,9,10,11,12]. Solar cell (o-SnS, c-SnS, Sn2S3 as light absorber and SnS2 as buffer); photodetector (o-SnS, c-SnS, Sn2S3 as light absorber and SnS2 as buffer); Li- and Na-ion batteries (o-SnS, c-SnS, and SnS2 as anode materials); gas- and bio sensors (o-SnS, c-SnS, and SnS2 as sensing materials); tunnel field-effect transistors (TFET) (o-SnS, c-SnS, and SnS2 as top or back gates); electrochemical and super capacitors (o-SnS, c-SnS, and SnS2 as electrode materials); capacitor; thermoelectrics (o-SnS, c-SnS, and Sn2S3 as grids); and water-splitting (o-SnS, c-SnS, and SnS2 as photocathodes).
Nanomaterials 11 01955 g001
Figure 2. (a). Crystal structures of ground state SnxSy forms (reprinted with permission [51] © 2017, Royal Society of Chemistry) and (b) standard powder diffraction patterns for SnxSy.
Figure 2. (a). Crystal structures of ground state SnxSy forms (reprinted with permission [51] © 2017, Royal Society of Chemistry) and (b) standard powder diffraction patterns for SnxSy.
Nanomaterials 11 01955 g002
Figure 3. (a) XRD profiles of o-SnS and c-SnS (reprinted with permission [26]. © 2016, Elsevier), hexagonal SnS2 (reprinted with permission [123]. © 2016, Elsevier), and orthorhombic Sn2S3 (reprinted with permission [113]. © 2016, Elsevier). (b) Simulated Raman spectra for SnxSy at different temperatures of 10 K, 150 K, and 300 K (reprinted with permission [51]. © 2017, Royal Society of Chemistry). (c) Experimental Raman spectra of o-SnS (reprinted with permission [110]. © 2017, Elsevier), c-SnS (reprinted with permission [26]. © 2016, Elsevier), hexagonal SnS2 (reprinted with permission [123]. © 2016, Elsevier), and orthorhombic Sn2S3 (reprinted with permission [113]. © 2016, Elsevier). (d) Band structures of SnxSy (reprinted with permission [114]. © 2016, American Physical Society), and (e) bandgap estimation of o-SnS (reprinted with permission [123]. © 2016, Elsevier), c-SnS (reprinted with permission [26]. © 2016, Elsevier), hexagonal SnS2 (reprinted with permission [123]. © 2016, Elsevier), and orthorhombic Sn2S3 (reprinted with permission [124]. © 2016, Sciendo).
Figure 3. (a) XRD profiles of o-SnS and c-SnS (reprinted with permission [26]. © 2016, Elsevier), hexagonal SnS2 (reprinted with permission [123]. © 2016, Elsevier), and orthorhombic Sn2S3 (reprinted with permission [113]. © 2016, Elsevier). (b) Simulated Raman spectra for SnxSy at different temperatures of 10 K, 150 K, and 300 K (reprinted with permission [51]. © 2017, Royal Society of Chemistry). (c) Experimental Raman spectra of o-SnS (reprinted with permission [110]. © 2017, Elsevier), c-SnS (reprinted with permission [26]. © 2016, Elsevier), hexagonal SnS2 (reprinted with permission [123]. © 2016, Elsevier), and orthorhombic Sn2S3 (reprinted with permission [113]. © 2016, Elsevier). (d) Band structures of SnxSy (reprinted with permission [114]. © 2016, American Physical Society), and (e) bandgap estimation of o-SnS (reprinted with permission [123]. © 2016, Elsevier), c-SnS (reprinted with permission [26]. © 2016, Elsevier), hexagonal SnS2 (reprinted with permission [123]. © 2016, Elsevier), and orthorhombic Sn2S3 (reprinted with permission [124]. © 2016, Sciendo).
Nanomaterials 11 01955 g003
Figure 4. (a) Possibility of different defect formation in SnxSy, and (b) defect formation energies in SnxSy as a function of Fermi energy under Sn-rich (S-poor) and Sn-poor (S-rich) conditions (reprinted with permission [114]. © 2016, American Physical Society).
Figure 4. (a) Possibility of different defect formation in SnxSy, and (b) defect formation energies in SnxSy as a function of Fermi energy under Sn-rich (S-poor) and Sn-poor (S-rich) conditions (reprinted with permission [114]. © 2016, American Physical Society).
Nanomaterials 11 01955 g004
Figure 5. Schematic representation of (a) CBD, (b) main chemical reagents used for the preparation of SnxSy thin films from 1987 to the present, and (c) importance of Ksp and Qip relation on the films by CBD.
Figure 5. Schematic representation of (a) CBD, (b) main chemical reagents used for the preparation of SnxSy thin films from 1987 to the present, and (c) importance of Ksp and Qip relation on the films by CBD.
Nanomaterials 11 01955 g005
Figure 6. Schematic representation of (a) ion-by-ion mechanism and (b) cluster-by-cluster (hydroxide) mechanism.
Figure 6. Schematic representation of (a) ion-by-ion mechanism and (b) cluster-by-cluster (hydroxide) mechanism.
Nanomaterials 11 01955 g006
Figure 7. (ac) XRD patterns, SEM images, and (αhυ)2 versus (hυ) graph of o-SnS films grown at various SnCl2 concentrations (reprinted with permission [191]. © 2012, Elsevier), and variation in o-SnS film (d) thickness with different TA concentrations (reprinted with permission [145]. © 1987, Elsevier). (e,f) Morphology and crystallinity changes of o-SnS films with ST concentrations (reprinted with permission [187]. © 2012, Elsevier). (g) Variation in SnS film thickness with different TEA concentrations (reprinted with permission [145]. © 1987, Elsevier). (h,i) XRD patterns of the o-SnS films and c-SnS films deposited at different TTA (reprinted with permission [211]. © 2019, Elsevier) and EDTA concentrations, respectively (reprinted with permission [26]. © 2016, Elsevier). (j) XRD patterns of SnS2 films deposited at various volumes of ammonia solution (reprinted with permission [41]. © 2012, Elsevier). (k) (αhυ)2 versus (hυ) for c-SnS films prepared using various EDTA amounts reprinted with permission [26]. © 2016, Elsevier). (l) SEM images of o-SnS films deposited with various TTA concentrations (reprinted with permission [211]. © 2019, Elsevier). (m,n) Scheme of the formation (reprinted with permission [185]. © 2011, Elsevier) and XRD patterns of o-SnS and c-SnS films (reprinted with permission [27]. © 2018, Elsevier). (o) Morphologies of o-SnS and c-SnS films (reprinted with permission [184]. © 2011, Elsevier), and (p) variation in the band gap o-SnS films at different pH values (reprinted with permission [27]. © 2018, Elsevier).
Figure 7. (ac) XRD patterns, SEM images, and (αhυ)2 versus (hυ) graph of o-SnS films grown at various SnCl2 concentrations (reprinted with permission [191]. © 2012, Elsevier), and variation in o-SnS film (d) thickness with different TA concentrations (reprinted with permission [145]. © 1987, Elsevier). (e,f) Morphology and crystallinity changes of o-SnS films with ST concentrations (reprinted with permission [187]. © 2012, Elsevier). (g) Variation in SnS film thickness with different TEA concentrations (reprinted with permission [145]. © 1987, Elsevier). (h,i) XRD patterns of the o-SnS films and c-SnS films deposited at different TTA (reprinted with permission [211]. © 2019, Elsevier) and EDTA concentrations, respectively (reprinted with permission [26]. © 2016, Elsevier). (j) XRD patterns of SnS2 films deposited at various volumes of ammonia solution (reprinted with permission [41]. © 2012, Elsevier). (k) (αhυ)2 versus (hυ) for c-SnS films prepared using various EDTA amounts reprinted with permission [26]. © 2016, Elsevier). (l) SEM images of o-SnS films deposited with various TTA concentrations (reprinted with permission [211]. © 2019, Elsevier). (m,n) Scheme of the formation (reprinted with permission [185]. © 2011, Elsevier) and XRD patterns of o-SnS and c-SnS films (reprinted with permission [27]. © 2018, Elsevier). (o) Morphologies of o-SnS and c-SnS films (reprinted with permission [184]. © 2011, Elsevier), and (p) variation in the band gap o-SnS films at different pH values (reprinted with permission [27]. © 2018, Elsevier).
Nanomaterials 11 01955 g007
Figure 8. (a) Variation in o-SnS film thickness with bath temperature (reprinted with permission [213]. © 2019, Elsevier). (b,f) Morphological and electrical properties of o-SnS at different bath temperatures (reprinted with permission [212]. © 2019, Elsevier). (c) Change in crystallinity of c-SnS films with bath temperature (reprinted with permission [50]. © 2016, Elsevier). (d,e) Plots of absorption coefficient refractive index and extinction coefficient of o-SnS films (reprinted with permission [17]. © 2015, Elsevier). (g,i) Variation in thickness and crystallinity of SnS2 films with deposition time (reprinted with permission [56]. © 2017, Elsevier). (h) Morphology of ORT- SnS thin films deposited at various times (reprinted with permission [177]. © 2010, Elsevier). (j) Variation in refractive index and (k) variation in extinction coefficient of Sn2S3 films with deposition time (reprinted with permission [34]. © 2012, Elsevier).
Figure 8. (a) Variation in o-SnS film thickness with bath temperature (reprinted with permission [213]. © 2019, Elsevier). (b,f) Morphological and electrical properties of o-SnS at different bath temperatures (reprinted with permission [212]. © 2019, Elsevier). (c) Change in crystallinity of c-SnS films with bath temperature (reprinted with permission [50]. © 2016, Elsevier). (d,e) Plots of absorption coefficient refractive index and extinction coefficient of o-SnS films (reprinted with permission [17]. © 2015, Elsevier). (g,i) Variation in thickness and crystallinity of SnS2 films with deposition time (reprinted with permission [56]. © 2017, Elsevier). (h) Morphology of ORT- SnS thin films deposited at various times (reprinted with permission [177]. © 2010, Elsevier). (j) Variation in refractive index and (k) variation in extinction coefficient of Sn2S3 films with deposition time (reprinted with permission [34]. © 2012, Elsevier).
Nanomaterials 11 01955 g008
Table 1. The advantages of SnxSy compared to the conventional absorbers (CdTe, CIGS, and CZTS) and the buffer (CdS).
Table 1. The advantages of SnxSy compared to the conventional absorbers (CdTe, CIGS, and CZTS) and the buffer (CdS).
CharacteristicsPV AbsorbersPV Buffers
CdTeCIGSCZTSo-SnSc-SnSSn2S3CdSSnS2
Earth abundanceNoNoYesYesYesYesYesYes
Eco-friendlyNoNoYesYesYesYesNoYes
Band gap
(eV)
1.45–1.5 eV
[13]
1.1–1.5
[14]
1.0–1.5
[15]
1.16–1.79
[16,17,18,19,20,21,22,23]
1.64–1.75
[24,25,26,27,28,29]
0.95–2.03
[30,31,32,33,34]
2.35–2.50
[14,35]
2.04–3.30
[36,37,38,39,40,41]
Absorption coefficient>104105>104105105104
Conductivity typep-typep-typep-typep-/n-typep-typep-/n-typen-typen-type
Carrier density
(cm−3)
1014–1017
[35]
1012–1018
[14]
1016–1018
[42]
1011–1018
[43,44,45,46,47,48,49]
1011–1018
[29,50,51]
1014–1016
[45,52,53]
1012–1018
[14,35]
1013–1017
[54,55,56]
StructureZinc blend
[13]
Chalcopyrite
[14]
Kesterite
[57]
Orthorhombic
[58,59]
Cubic
[60]
Orthorhombic
[51]
Hexagonal
[35]
Hexagonal
[61]
Maximum theoretical efficiency
(%)
~29~2931 [62]31 [63]>25 [64]
Table 2. The historical information and the applications of SnxSy.
Table 2. The historical information and the applications of SnxSy.
Tin
Sulfides
Mineral Form
[65,66,67]
Appearance
[68]
Other NamesDiscovered/Reported
[69,70]
Applications
[24,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91]
o-SnSHerzenbergite
Nanomaterials 11 01955 i001
Black color with dark red–brown internal reflections.KolbeckineReported by Ramdohr from the Maria-Teresa mine (Oruro, Bolivia) in 1934.PV, photodetectors [24], photocatalysts [71], water splitting [72], supercapacitors [83], field-effect transistors [85], sodium-ion and lithium-ion batteries [86,87], gas sensors [88], biosensors [89] thermoelectric [90], and electro chemical capacitors [91].
SnS2Berndtite
Nanomaterials 11 01955 i002
Pale yellow with intense brownish to yellow–orange internal reflections.Mosaic goldDiscovered at the Stiepelmann mine in Arandis, Namibia, as described by Ramdohr in 1935.PV, photocatalysts [73], water splitting [74], supercapacitors [75], field-effect transistors [76], lithium-ion and sodium-ion batteries [77,78], gas sensors [79], thin film diodes [80], and high-speed photodetectors [81].
Sn2S3Ottemanite
Nanomaterials 11 01955 i003
Gray with orange–brown internal reflections.-Reported by Moh from the Cerro de Potosi mine (Bolivia) in 1964.PV, optoelectronic [82], thermoelectric and IR detectors [84].
Table 3. Optimized theoretical and experimental lattice parameters, reported optical bandgaps, and electrical parameters of SnxSy (Th: theoretical, Exp: experimental).
Table 3. Optimized theoretical and experimental lattice parameters, reported optical bandgaps, and electrical parameters of SnxSy (Th: theoretical, Exp: experimental).
SnxSy PhaseStructural PropertiesOptical PropertiesElectrical Properties
Structure (Space Group)Oxidation State of SnParameters of unit cellOptical Band Gap
(eV)
Carrier Concentration
(cm−3)
Mobility
(cm2V−1s−1)
Resistivity
(Ω-cm)
Angles and RuleIntercepts a (Å), b (Å), c (Å)
Theoretical [51]Experimental [111,125,126]
o-SnSOrthorhombic (Pnma)2+α = β = γ = 90°
a ≠ b ≠ c
4.251,
11.082,
3.978
4.33,
11.18,
3.98
1.16 [16], 1.30 [17], 1.32 [18], 1.35 [19], 1.42 [20], 1.43 [20], 1.48 [21], 1.70 [22], 1.79 [23].1 × 1011 [43],
3.6 × 1012 [44],
(1–1.2) × 1015[21,45],
1.5 × 1016 [46],
(1–1.16) ×1017 [47,48],
(1–3) ×1018 [49].
3.7 [20],
15.3 [46],
90 [43,49],
228 [44],
385 [19],
400–500 [17].
12.98 [17],
14.49 [20],
30 [16],
33.33 [18],
0.63 × 103 [43],
2.1 × 104 [44],
(0.16–0.25) × 105 [127,128].
c-SnSCubic
(P213)
2+α = β = γ = 90°
a = b = c
11.50611.6031.64 [24], 1.66 [25], 1.67 [26], 1.73 [27], 1.74 [28], 1.75 [29].5.87 × 1011 [29],
7.93 × 1012 [50],
6 × 1018 [51].
1.47 × 10−2 [51],
75 [50],
77.7 [29].
70 [51],
1 × 104 [50],
1.37 × 105 [29],
1 × 106 [25],
1 × 107 [28].
SnS2Hexagonal
(P 3 ¯ ml)
4+α = β = 90°; γ = 120˚
a = b ≠ c
3.651,
3.651,
6.015
3.638,
3.638,
5.880
2.04 [36], 2.12 [39], 2.14[129], 2.18 [37], 2.30[38], 2.35 [130], 2.40[131], 2.41 [39], 2.44[132],2.45[133] 2.50[134],2.67[135] 2.75[136], 2.80 [56], 3.08 [40], 3.30 [41].1 × 1013 [54],
2 × 1017 [55],
6.8 × 1017 [56].
15 [54],
48 [56],
51.5 [55].
1.11 [55],
11.2 [56],
0.77 × 102[137],
0.42 × 105 [54],
0.26 × 107 [138].
Sn2S3Orthorhombic (Pnma)2+ and 4+α = β = γ = 90°a ≠ b ≠ c8.11,
3.76,
13.83
8.878,
3.751,
14.020
0.95 [30], 1.16 [31] 1.2 [139], 1.65 [32], 1.9 [33], 1.96 [140], 2.0 [141], 2.03 [34].9.4 × 1014 [52]
1 × 1015 [45],
4.0 × 1016 [53].
20.5 [53]0.359 [124],
7.57 [53],
0.66 × 102 [45],
(0.22–0.36) × 103 [52,141],
(0.4–2.5) × 105 [137].
Table 4. Deposition conditions and the physical properties of ORT- and c-SnS films grown by CBD.
Table 4. Deposition conditions and the physical properties of ORT- and c-SnS films grown by CBD.
SnxSy PhasePrecursorsComplexing GentDeposition ParametersStructureBand Gap
(eV)
Electrical ParametersRef
TypeR
(Ωcm)
µ
(cm2V−1S−1)
N
(cm−3)
o-SnS
o-SnST(II)C = 0.1 M
TA = 0.1 M
TEA = 15 mL
NH3 = 8 mL
Tb = 27 °C
td = 20 h
pH = 10.5 ± 1
Amorphous1.51 (i)n1987 [145]
o-SnST(II)C = 0.025 mol
SDS/AS = 0.025 mol
Tb = –
td = –
pH = 3,10,12
ORT(013)1.08p107–1031989 [36]
o-SnST(II)C = –
TU = –
Tb = –
td = –
pH = –
Polycrystalline1.3 (i)1990 [157]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 75 °C, 25 °C
td = 5 h, 40 h
pH = –
Polycrystalline1.3p1991 [128]
o-SnST(II)C = 2 mL
TA = 1 mol L−1
TEA = 0.5 mL
NH3 = –
Tb = 50 °C, 25 °C
td = 2–4 h, 5–10 h
pH = –
Crystalline1991 [158]
o-SnST(II)C = 1 g
TA = 4 mL, 8 mL
TEA = 12 mL
NH3 = 12 mL
Tb = 75 °C, 25 °C
td = 5 h, 40 h
pH = –
1991 [159]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 60 °C
td = 7 h 30 min
pH = –
ORT(111)1992 [160]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 60 °C
td = 7 h 30 min
pH = 9.5
ORT(111)p1993 [161]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 13 mL
Tb = 50–75 °C,
td = 1.5 h, 20 h
pH = –
ORT(111)p1994 [162]
o-SnST(II)C = 15 g
TU = 5 g, 10 g
Tb = –
td = 5 min
pH = 3
Sp = 1.33 mm/s
ORT(040) 1.41999 [163]
o-SnST(II)C = 0.1 M
SDS = 0.05 M
Tb = 80 °C
td = –
pH = 12
p2001 [164]
o-SnST(II)C = 1.125 g
ST = 2 M
AFTb = Tr
td = 18 h
pH = 7
ORT(111)1.38 (d)
0.96–1.14 (i)
2003 [165]
o-SnST(II)C = 1 g
TA = 8 mL
TEA = 12 mL
NH3 = 10 mL
Tb = 35 °C
td = 15 h
pH = 9.5
ORT(111)1.18 (d)p107–1042003 [166]
o-SnST(II)C = 0.56 g
SS = 0.025 M
Tb = 80 °C
td = –
pH = 12
ORT(111)2006 [167]
o-SnST(II)C = 1.13 g
TA = 0.1
TEA = 30 mL
NH3 = 16 mL
Tb = RT 293–298 K
td = 5–6 h
pH = –
ORT(111)1.17 (d)
1.12 (i)
108–1062007 [168]
T(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 308 K
td = 20 h
pH = –
o-SnST(II)C = 1 g
TA = 1 M
TEA = 10 mL
NH3 = 5 mL
Tb = 45 °C
td = –
pH = –
ORT(111)1.33–1.39 (d)2007 [169]
o-SnST(II)C = 1 g
TA = 8 mL
TEA = 12 mL
NH3 = 10 mL
Tb = 55 °C
td = 8 h
pH = –
ORT(111)p1039010112008 [43]
o-SnST(II)C = 1.12 g
ST = 0.5 M
TTA = 10 mLTb = Tr
td = 24 h
pH = 7
ORT(111)1.1 (d)106-2008 [170]
o-SnST(II)C = 1 g
TA = 8 mL
TEA = 12 mL
NH3 = 10 mL
Tb = 313 K
td = 8–22 h
pH = –
ORT(111)1.2–1.7 (d)p2009 [171]
o-SnST(II)C = 0.15 M
ST = 2 M
NH4OH = 6 mLTb = 30 °C
td = 24 h
pH = 7
ORT(040)/(141)1.31 (d)2009 [172]
o-SnST(II)C = 2 × 10−2 M
TA = 1 × 10−2–8 × 10−2 M
Tb = 80 °C
td = 60 min
pH = 1.87
Amorphous2009 [142]
o-SnST(II)C = 1 g
TA = 8 mL
TEA = 12 mL
NH3 = 10 mL
Tb = 55 °C
td = 8 h
pH = –
ORT(111)1.12 (i)2009 [173]
o-SnST(II)C = 0.15 M
ST = 2 M
AH = 6 mLTb = Tr
td = 24 h
pH = 7
ORT(111)2009 [174]
o-SnST(II)C = –
TA = –
TEA = –
NH3 = –
Tb = 75 °C
td = –
pH = –
ORT(111)0.82–1.22 (i)2009 [175]
o-SnST(II)C
TA = 0.1 M
TEA = 30 mL
NH4OH = 16 mL
Tb = –
td = 5 h
pH = –
ORT(111)/(040)1.76 (i)2010 [176]
o-SnST(II)C = 1 M
TA = 1 M
TEA = 10 mL
TSS = 5 mL
NH3/NH4Cl = 5 mL
Tb = 60 °C
td = 2–10 h
pH = 9.31
ORT(111)/(040)1.30–1.97 (d)
0.83–1.36 (i)
p9.9–12.32010 [177]
o-SnST(II)C = 1 M
TA = 1 M
TEA = 10 mL
TSS = 5 mL
NH3/NH4Cl = 5 mL
Tb = 27 °C
td = 24 h
pH = 10.7
ORT(110)1.37 (d)
1.05 (i)
p1059 × 1052010 [178]
o-SnSTEA = 12.5 M, 13 MTb = –
td = –
pH = –
1.93–2.16 (d)2010 [179]
o-SnST(II)C = 0.95 g
TA = 0.1 M
TEA = 8 mL
NH3 = 6 mL
Tb = 75 °C
td = 1 h
pH = –
ORT(111)/(040)1.3 (i)p2010 [180]
o-SnST(II)C = –
TA = –
TEA, NH3
TTA
Tb = Tr, 90 °C
td = 24 h, 3 h
pH = –
ORT(400)1.1–1.9 (d)2011 [181]
o-SnST(II)C = 0.2 M
ST = 0.2 M
Na2EDTA = 25 mL of 0.2 MTb = 40–80 °C
td = 30 h
pH = 1.5
1.2–1.5 (d)2011 [182]
o-SnST(II)C = 0.15 M
ST = 0.15 M
Na2EDTA = 25 mL of 0.2 MTb = 75 °C
td = 150 min
pH = –
1.2–1.6 (d)2011 [183]
o-SnST(II)C = 0.1 M
ST = 0.25 M
AC = 50 mL of 0.2 MTb = 35 °C
td = 10 h
pH = 5, 6
ORT(111)1.75 (d)
1.12 (i)
2011 [184]
o-SnST(II)C = 0.1 M
ST = 0.25 M
AC = 50 mL of 0.2 MTb = 35°C
td = 10 h
pH = 5
ORT(111)1.75 (d)
1.15 (i)
4202011 [185]
o-SnST(II)C
TA
TEA = –
NH3 = –
Tb = 20–50 °C
td = –
pH = –
ORT(111)1.15 (i)
1.35(d)
p6.3 ± 0.111 ± 71016–10172011 [186]
o-SnST(II)C = 0.1 M
ST = 0.25–0.75 M
AC = 50 mL of 0.3 MTb = 60–80 °C
td = 3 h
pH = 5
ORT(111)/(040))1.01–1.26 (i)p1032012 [187]
o-SnST(II)C = –
TA = –
TEA = –
NH4Cl = –
Tb = 45 °C
td = 5 h
pH = –
ORT(111)0.7–1.3 (i)2012 [188]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 60 °C
td = 6 h
pH = 6
ORT(111)/(101)0.9–1.1106–1012012 [189]
o-SnST(II)C = 1 M
TA = 1 M
TEA = 10 mL
NH3 = 2 mL
Tb = RT = 27 °C
td = 24–72 h
pH = 9.7
ORT(111)1.14–1.18 (i)
1.32–1.44 (d)
2012 [190]
o-SnST(II)C = 0.06 M–0.12 M
TA = 0.1 M
TEA = 1.85 M
NH3 = 1.5 M
Tb = 30 °C
td = 90 min
pH = –
ORT(040)/(111)1.5–1.95 (d)2012 [191]
T(II)C = 0.1 M
TA = 0.1 M
TEA = 1.75–1.90 M
NH3 = 1.5 M
Tb = 30 °C
td = 90 min
pH = –
T(II)C = 0.1 M
TA = 0.1 M
TEA = 1.85 M
NH3 = 1.5 M
Tb = 40–60 °C
td = 90 min
pH = –
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH4OH = 10 mL
Tb = 60 °C
td = 6 h
pH = –
ORT(111)1.9 (d)
1.1 (i)
2013 [192]
o-SnST(II)C = 0.1 M
TA = 0.6 M
TTA = 1 MTb = 50–70 °C
td = 50 min
pH = 1.5
ORT(111)1.30–1.35 (d)2013 [193]
o-SnST(II)C = 0.05–0.2 M
TA = 0.4–0.7 M
Na2EDTA = 20 mL of 0.1 MTb = 50–80 °C
td = 0.5–3 h
pH = 9–12
ORT(200)2013 [194]
o-SnST(II)C = 0.1 M
ST = 0.3 M
Na2EDTA = 5 mL of 0.1 M
TSC = 5 mL of 0.66 M
Tb = Tr
td = 24 h
pH = 10
1.50–1.90 (d)2013 [195]
o-SnST(II)C = 0.5 M
TU = 1 M
NH3 = 3 MTb = Tr
td = 60–180 min
pH = –
1.98–2.01(d) 1.82–1.98 (i)p2013 [196]
o-SnST(II)C
TA = 0.1 M
TEA = 30 mL
NH4OH = 16 mL
Tb = –
td = 5 h
pH = –
ORT(111)/(200)1.64–1.7 (f)2013 [197]
o-SnST(II)C = 0.1 M
TA = 0.1 M
EDTA = 0.05 M–0.08 M
NH3 = 1.4 M
Tb = –
td = 3–4 h
pH = –
ORT(111)/(101)1.5–1.60 (d)p4002013 [198]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 6 mL
NH3 = 10 mL
Tb = –
td = –
pH = –
ORT(240)1.78–1.75 (d)109–1082014 [199]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 20–40 °C
td = 24 h
pH = 11
ORT(111)ORT 1.1 (i)p107–1022014 [146]
o-SnST(II)C = 0.5 g
TA = 1 M
TEA = 6 mL
TSC = 0.006–0.008 M
NH3 = 5 mL
Tb = 30 °C
td = 24 h
pH = –
ORT(111)1.17–1.40 (d)104148–22810122014 [44]
o-SnST(II)C = –
TA = –
TSC = –Tb = 50 °C
td = 2.5 h
pH = 5
ORT(111)1.25–1.83 (d)
1.1–1.65 (i)
n1032014 [200]
o-SnST(II)C = 0.03 M
ST = 0.03 M
TTA = 0.44 MTb = Tr
td = 24 h
pH = 7
ORT(400)1.49–1.39 (i)
1.28–1.5 (i)
2014 [201]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 40 °C
td = 17 h
pH = –
ORT(111)1.25–1.1 (i)1032015 [202]
o-SnST(II)C = 0.1 M
TA = 0.1 M
TEA = 15 mL
NH3 = 8 mL
Tb = 26 °C
td = 22 h
pH = –
ORT(021)1.76–3.32 (d)2015 [203]
o-SnST(II)C = –
ST = 0.01–0.09 M
TTATb = 22 °C
td = 24 h
pH = 7
2015 [204]
o-SnST(II)C = 20 mL
TA = 20 mL
TTA = 1 MTb = 40–80 °C
td = 50 min
pH = 1.5
1.33–1.41 (d)2015 [17]
o-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 M
NH3 = –
Tb = 80 °C
td = 4 h
pH = 7
ORT(040)1.65 (d)p2016 [205]
o-SnST(II)C = 1 g
TA = 8 mL
TEA = 12 mL
NH3 = 10 mL
Tb = 40 °C
td = 10 h
pH = 11
ORT(111)ORT = 1.1 (i)p1062016 [25]
o-SnST(II)C = 0.1 M
TA = 20 mL
TTA = 1 MTb = 70 °C
td = –
pH = –
ORT(111)1.31–1.26 (d)p6–381241015–10162016 [144]
o-SnST(II)C = 1 g
TA = 0.3 g
TEA = 5.5 mL
NH3 = 5 mL
Tb = 70 °C
td = –
pH = –
ORT(002)1.14–1.75 (d)2016 [206]
o-SnST(II)C = –
TA = –
TTA = 1 MTb = 70 °C
td = –
pH = –
ORT(111)1.3 (d)p38–14.255–231015–10192016 [123]
o-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.15–0.21 MTb = 80 °C
td = 4 h
pH = 5.8
ORT(111)1.64–1.1 (d)2017 [207]
o-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 MTb = 80 °C
td = 4 h
pH = 6.5–7.5
ORT(111)1.51 (d)2018 [27]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 312 mL
NH3 = 10 mL
Tb = 40 °C
td = 17 h
pH = 1.5
ORT(111)1.1 (i)2018 [208]
o-SnST(II)C = 4 mmol
TA = 4–8 mmol
TSC = 0.15–0.21 MTb = 80 °C
td = 1–2 h
pH = 0.4–1.0
ORT(111)1.39–1.41 (d)2018 [209]
o-SnST(II)C = 0.1 M
TA = 0.15 M
TEA = –Tb = 343 K
td = 120, 240, 369 min
pH = 4
ORT(013)2018 [210]
o-SnST(II)C = 20 mL
TA = 20 mL
TTA = 0.6–1.6 MTb = 70 °C
td = 50 min
pH = –
ORT(111)1.28–1.45 (d)p38–6229–1081.92 × 1015–4.12 × 10152019 [211]
o-SnST(II)C = 0.1 M
TA = 0.6 M
TTA = 1 MTb = 40–80 °C
td = 50 min
pH = 1.5
ORT(111)1.30–1.41 (d)p38551.5 × 1015–3.4 × 10152019 [212]
o-SnST(II)C = 1 g
TA = 1 M
TEA = 18 mL
NH3 = 10 mL
Tb = 40–70 °C
td = 3 H
pH = 10
ORT(040)1.32–2.08 (d)2019 [213]
o-SnST(II)C = 0.2 M
TA = 0.4 M
TTA = 0.5 MTb = 50–80 °C
td = 90 min
pH = 1.5
ORT(040)1.55–1.92 (d)2019 [214]
o-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 MTb = 80 °C
td = 4 h
pH = 5.0–6.5
ORT(111)1.34–1.51 (d)2019 [215]
o-SnST(II)C = 4 mmol
TA = 6 mmol
Tb = 80 °C
td = 120 min
pH = 0.7
ORT(111)1.41–1.49 (d)2019 [216]
o-SnST(II)C = 2 g
ST = 0.2 M
TEA = 70 mL
CA = 0.4 M
NH3 = 10 mL
Tb = 55 °C
td = 4 h
pH = 11
ORT(111)1.33 (i) 2019 [217]
o-SnST(II)C = 20 mL
TA = 10 mL
PVA = 2 g
TTA = 0.5 MTb = 80 °C
td = 45–90 min
pH = 10
ORT(040)1.55–1.79 (d)2019 [218]
o-SnST(II)C = 0.1 M
TA = 1 M
TEA = 10 mL
TSC = 0.66 M
Tb = –
td = –
pH = 9.2–9.6
ORT(102)1.36–1.99 (d)2020 [219]
o-SnST(II)C = 0.1 mol
TA = 0.4 mol
AA = 0.8 mLTb = 75 °C
td = 70 min
pH = 9.2–9.6
ORT(110)2020 [220]
o-SnST(II)C = –
TA = 0.1 M
TEA = –
NH3 = 15 mL
Tb = 25 °C
td = 4 h
pH = –
200–600 °C
1.5–1.7 (d)2021 [221]
o-SnST(II)C = 1 g
TA = 0.6 g
TEA = 12 mL
NH3 = 15 mL
Tb = 70 °C
td = 2 h
pH = 10.93
ORT(111)1.38 (d)2021 [222]
o-SnST(II)C = 4 m mol
TA = 6 m mol
Tb = 80 °C
td = 120 min
pH = 0.7
ORT(111)0.78–1.13 (d)2021 [223]
o-SnSTb = 65 °C
td = 3 h
pH = 5.5–8.5
ORT(111)1.41–1.75 (d)2021 [224]
c-SnS
c-SnST(II)C = 2.26 g
TA = 10 mL
TEA = 30 mL
NH3 = 16 mL
Tb = 25 °C
td = 6 h
pH = –
CUB (111)/(200)1.64–1.73 (d)p1051041092008 [43]
c-SnST(II)C = 2.26 g
TA = 10 mL
TEA = 30 mL
NH3 = 16 mL
Tb = 25 °C
td = 6 h
pH = –
CUB (111)/(200)1.7 (d) 2009 [171]
c-SnST(II)C = 2.26 g
TA = 10 mL
TEA = 30 mL
NH3 = 16 mL
Tb = 25 °C
td = 6 h
pH = –
CUB (111)/(200)1.7 (d)2009 [173]
c-SnST(II)C = –
TA = 0.1 M
TEA = 8964 g
NH4OH = 15 M
Tb = 25 °C
td = 2–4 h 30 min
pH = –
CUB (111)/(200)1. 7 (d)2011 [225]
c-SnST(II)C = –
TA = 0.1 M
TEA = –
NH4OH = 15 M
Tb = 25 °C
td = –
pH = –
2011 [226]
c-SnST(II)C = –
TA = 0.1 M
TEA = 8964 g
NH4OH = 15 M
Tb = 25 °C
td = –
pH = –
CUB (111)/(200)1.76 (d)
1.44–1.51 (d)
2012 [227]
c-SnST(II)C = –
TA = 0.1 M
TEA = 8964 g
NH4OH = 15 M
Tb = 25 °C
td = –
pH = –
CUB (111)/(200)1.76 (d)2012 [228]
c-SnST(II)C = 1 g
TA = 1 M
TEA = 12 mL
NH3 = 10 mL
Tb = 20–40 °C
td = 24 h
pH = 11
CUB (111)/(200)1.67 (d)p107–1022014 [146]
c-SnST(II)C = 2.26 g
TA = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C
td = 10 h
pH = –
CUB (222)/(400)1.74 (d)2015 [28]
c-SnST(II)C
TA = 0.1 M
TEA = 0.1 M
NH4OH = 15 M
Tb = 25 °C
td = –
pH = –
CUB(111)/(200)1.70 (d)2015 [229]
c-SnST(II)C = 2.26 g
T(II)C = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C
td = 15 h
pH = 11
CUB(222)/(400)1.73 (d)1032016 [230]
c-SnS T(II)C = 2.26 g
TA = 10 mL
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C, 10 °C
td = 4 h, 18 h
pH = 11
CUB(222)/(400)1.66–1.72 (d)p1062016 [25]
c-SnST(II)C = 0.1 M
TA = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 25 °C
td = 6 h
pH = 11
CUB(222)/(400)2016 [231]
c-SnST(II)C = 2.26 g
TA = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C
td = 10 h
pH = 11
CUB(222)/(400)2016 [232]
c-SnST(II)C = 2.26 g
ST = 1 M
EDTA = 20 mL of 0.5 MTb = 25–65 °C
td = 6 h
pH = 10.5
CUB(222)/(400)1.74–1.68 (d)p105–1048.98–28.61012–10132016 [50]
c-SnST(II)C = 2.26 g
ST = 1 M
EDTA = 15–25 mL of 0.5 M
NH3 = 5 mL
Tb = 45 °C
td = 6 h
pH = 10.5
Sp = –
CUB(222)/(400)1.67–1.73 (d)p105–1040.34–28.61014–10122016 [26]
c-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 MTb = 80 °C
td = 4 h
pH = 7
CUB(222)/(400)1.64 (d)2017 [24]
c-SnST(II)C = 0.1 M
TA = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C, 80 °C
td = 3 h, 21 h
pH = –
2017 [233]
c-SnST(II)C = 0.1 M
ST = 0.125 M
EDTA = 0.1 MTb = 45 °C
td = 6 h
pH = –
CUB(222)/(400)1.67–1.75 (d)p105–1045.22–77.71011–10132017 [29]
c-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 MTb = 80 °C
td = 4 h
pH = 6.5–7.5
CUB(222)/(400) 1.64–1.73 (d) 2018 [27]
c-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 MTb = 80 °C
td = 4 h
pH = 7
CUB(222)/(400)1.5 (d)2018 [234]
c-SnST(II)C = 0.5 M
TA = 0.5 M
TEA = 30 mL
NH3 = 16 mL
Tb = 35 °C
td = 4 h
pH = 9.78
CUB(222)/(400)1.74 (d)2018 [235]
c-SnST(II)C = 2.26 g
TA = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C, 80 °C
td = 3 h, 21 h
pH = –
CUB(222)/(400)1.76 (d)2018 [236]
c-SnST(II)C = 2.26 g
TA = 10 mL
TEA = 30 mL
NH3 = 16 mL
Tb = 17–8 °C
td = 3–21 h
pH = –
CUB(222)/(400) 2019 [217]
c-SnST(II)C = 0.04 M
TA = 0.08 M
TEA = 1.1 M
NH3 = 9.5 mL
Tb = 30 °C
td = 4 h
pH = –
CUB(222)/(400)1.74 (d)2020 [237]
c-SnST(II)C = 0.1 M
TA = 0.15 M
TSC = 0.2 M Tb = –
td = -
pH = –
CUB(222)/(400)2020 [238]
c-SnST(II)C = 1 g
TA = 0.3 g
TEA = 5.5 mL
NH3 = 5 mL
Tb = 24 °C
td = 4.25 h
pH = 9.25
CUB(222)/(400)1.70–1.74 (d)103–1042020 [239]
c-SnST(II)C = 0.2 M
TA = 0.1 M
TEA = 5.5 mL
NH3 = 5 mL
Tb = 17–8 °C
td = 3–21 h
pH = 11
CUB(222)/(400)1.76 (d)p1082020 [240]
c-SnST(II)C = 2.25 g
ST = 0.1 M
EDTA = 0.5 M
NH3 = 5–7.5 mL
Tb = 50 °C
td = 6 h
pH = 10.3
CUB(222)/(400)1.75–1.8 (d)p103–10415–751012–10132020 [241]
c-SnST(II)C = 1 g
TA = 0.6 g
TEA = 12 mL
NH3 = 15 mL
Tb = 70 °C
td = 2 h
pH = 8.24
CUB(200)1.72 (d)2021 [222]
c-SnST(II)C = 1.21 g
TA = 0.5 M
TTA = 1 MTb = 80 °C
td = 2–6 h
pH = 5–8
CUB(222)/(400)1.72–1.90 (d)107–1082021 [242]
c-SnST(II)C = 0.5 g
TA = 1 M
NTA = 0.6 MTb = 40 °C
td = 90–182 min
pH = 10
CUB(222)/(400)1.77–1.81 (d)1062021 [243]
c-SnST(II)C = 0.01 mol
TA = 0.1 M
TEA = 0.6 MTb = 17–8 °C
td = 3–21 h
pH = 10
CUB(222)/(400)1.70–1.80 (d)2021 [244]
Table 5. Deposition conditions and the physical properties of SnS2 and Sn2S3 films grown by CBD.
Table 5. Deposition conditions and the physical properties of SnS2 and Sn2S3 films grown by CBD.
SnxSy PhasePrecursorsComplexing AgentDeposition ParametersStructureBand Gap
(eV)
Electrical ParametersRef
TypeR
(Ωcm)
µ
(cm2V−1S−1)
N
(cm−3)
SnS2
SnS2T(II)C = 0.025 mol
SDS/AS = 0.025 mol
Tb = –
td = –
pH = 3, 10, 12
2.04SnS2-n107–1031989 [36]
SnS2Tin-ingots (99.9%)
ST = 10 mL
Tb = Tr
td = 2 h
pH = –
Amorphous2.35 (d)n103–1041990 [130]
SnS2Tin ingots (99.9%)
ST = 10 mL
Tb = 27 °C
td = –
pH = 1.4
Amorphous2.20 (i)n107–1081992 [147]
SnS2T(II)C = 1.13 g
TA = 0.1 M
EDTA = 25 mL
NH3 = 15 mL
Tb = Tr
td = 10–120 min
pH = 10
2.3 (d)n4 × 10−11997 [38]
SnS2T(II)C = 15 g
TU = 5 g, 10 g
Tb = –
td = 5 min
pH = 3
Sp = 1.33 mm/s
HEX(001)2.05 (i)1999 [163]
SnS2TC(IV) = 0.02 mol
TA = 0.5 mol L−1
CA = 0.375, 0.5, 0.625 mol/LTb = 35 °C
td = –
pH = 1.3
2.40 (d)2011 [131]
SnS2T(II)C = 1 g
TA = 0.5 M
TEA = 24 mL
NH3 = 12 mL–20 mL
Tb = 60 °C
td = 2 h
pH = –
HEX(001)3.3–3.7 (d)2012 [41]
SnS2T(II)C = 0.8 M
TA = 0.5 M
TEA = 3.75 M
NH3 = 12 mL
Tb = 60 °C
td = –
pH = –
HEX(001)2.8–3.0 (d)2013 [245]
SnS2T(II)C = 2.26 g
ST = 1 M
EDTA = 20 mL of 0.5 M
NH3 = 5 mL
Tb = 45 °C
td = 6 h
pH = –
HEX(001)2.58 (d)2017 [246]
SnS2T(II)C
TA
TTA = 1 MTb = –
td = 30–120 min
pH = –
HEX(001)2.95–2.80 (d)n11.24810172017 [56]
SnS2T(II)C = 0.84 g
TA = 0.5 M
TEA = 24 mL
NH3 = 16 mL
Tb = 60 °C
td = 2 h
pH = –
2018 [247]
SnS2T(II)C = 0.1 M
TA = 0.1 M
TTA = 0.1 M Tb = 60 °C
td = 6 h
pH = –
HEX(001)2.25–2.53 (d)2019 [248]
Sn2S3
Sn2S3T(II)C = 1 M
TA = 1 M
TEA = 10 mLTb = 30 °C
td = 20–24 h
pH = 10.7
ORT(131)2.03–2.12 (d)2012 [34]
Sn2S3T(II)C = 1.4 g
TA = 1 M
TEA = 30 mL
NH3 = 50 mL
Tb = RT
td = 24 h
pH = –
ORT(211)1.2 (d)2012 [139]
Sn2S3T(II)C = 0.05 M
SDS = 0.05 M
Tb = –
td = –
pH = –
ORT(021)1.3 (d)2018 [249]
Sn2S3T(II)C = 0.1 M
TA = 0.1 M
TEA = 30 mL
NH3 = 16 mL
Tb = 17 °C
td = 15 h
450 °C
(S-powder: 15 mg), 5–75 min
ORT(211)1.75 (d)p1046 × 10−62020 [250]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gedi, S.; Minnam Reddy, V.R.; Kotte, T.R.R.; Park, C.; Kim, W.K. Fundamental Aspects and Comprehensive Review on Physical Properties of Chemically Grown Tin-Based Binary Sulfides. Nanomaterials 2021, 11, 1955. https://doi.org/10.3390/nano11081955

AMA Style

Gedi S, Minnam Reddy VR, Kotte TRR, Park C, Kim WK. Fundamental Aspects and Comprehensive Review on Physical Properties of Chemically Grown Tin-Based Binary Sulfides. Nanomaterials. 2021; 11(8):1955. https://doi.org/10.3390/nano11081955

Chicago/Turabian Style

Gedi, Sreedevi, Vasudeva Reddy Minnam Reddy, Tulasi Ramakrishna Reddy Kotte, Chinho Park, and Woo Kyoung Kim. 2021. "Fundamental Aspects and Comprehensive Review on Physical Properties of Chemically Grown Tin-Based Binary Sulfides" Nanomaterials 11, no. 8: 1955. https://doi.org/10.3390/nano11081955

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