DOI:
10.1039/D5QM00399G
(Research Article)
Mater. Chem. Front., 2025,
9, 3197-3207
Stabilization of the cubic π-phase of SnS by calcium substitution
Received
28th May 2025
, Accepted 1st September 2025
First published on 25th September 2025
Abstract
The cubic phase of tin monosulphide, π-SnS, is of significant interest due to its attractive properties, such as a wider band gap suitable for solar photovoltaic application and being easier to epitaxially deposit onto technologically relevant semiconductors compared to the thermodynamically stable orthorhombic phase of α-SnS. Recently, we reported cation-assisted phase control for obtaining π-SnS rather than α-SnS using Pb2+ cations with a concentration of ∼20 cation percent (cat%). However, replacing Pb2+ with alternative non-toxic, environmentally friendly cations for cubic phase stabilization would be clearly advantageous. We have computationally investigated the energetics and electronic properties of calcium ion impurities in both SnS polymorphs. We found that addition of Ca2+ cations enables phase control of SnS grown from solution from α-SnS to π-SnS. Experimentally, we observed compact films of π-SnS after incorporating Ca2+ cations. Computational results indicated that ∼11 cat% of Ca2+ ions are required for preferred growth of π-SnS over α-SnS. Furthermore, the presence of an intermediate layer of CaS is computationally predicted to significantly contribute to the stabilization of the π-SnS phase, thereby reducing the Ca concentration required, which aligns well with experimental observations. Subsequently, we find that CaS is a promising substrate for epitaxial growth of π-SnS in the (111) orientation. Moreover, the bandgap of π-SnS decreased slightly with increasing concentration of Ca cations in the material. These results can facilitate the bulk scale synthesis of π-SnS material, bringing it closer to practical utility for a range of applications.
Introduction
Tin monosulfide (SnS) holds promising characteristics of low toxicity, cost-effectiveness, and earth abundance.1–4 SnS, a member of the IV–VI semiconductor family, features a narrow band gap and demonstrates optical activity in the near-infrared region.5 SnS possesses high absorption properties and high hole mobilities, which provides its potential applications in areas of solar cells,1,3,4,6 field effect transistors,7 photodetectors,2,8 gas sensors,9 and Li-ion and Na-ion batteries.10
At room temperature, SnS manifests two primary polymorphs: the ground state orthorhombic phase (α-SnS) and the metastable cubic phase (π-SnS), first reported in 2015.11,12 The lattice parameters for orthorhombic α-SnS are a = 4.329 Å, b = 11.192
Å, and c = 3.984 Å.13–15 For π-SnS, the lattice parameter has been reported in the range of 11.595–11.7 Å.11,16
Nair et al. investigated solar cells featuring absorber layers made from the α-SnS phase, the π-SnS phase, and a mixture of these two phases.17 Their research revealed that solar cells with absorber layers incorporating both the α- and π-SnS phases exhibited significantly enhanced open-circuit voltage (Voc), conversion efficiency (η), and fill factor (FF) compared to those with layers made from a single-phase.17
A variety of techniques have been employed to produce SnS thin films, including vacuum evaporation,18–20 spray pyrolysis,21–23 electrodeposition,24,25 radio frequency (RF) sputtering,26–28 atomic layer deposition (ALD),29 and ultrasonication.30 Among these, chemical solution deposition (CD) stands out as a cost-effective and straight forward method for growing thin films, making it ideal for large-scale industrial applications.31,32 This technique has been extensively used for different thin film materials,32–37 including SnS.38–41 SnS films have been successfully deposited from solution onto multiple substrates, such as glass, silicon (Si), fluorine-doped tin oxide (FTO), and GaAs.42–45
Foreign impurities in SnS thin films and nanoparticles have been shown to alter their properties, including electrical, optical, structural, and magnetic characteristics. In particular, various elements, such as Se,46 In,47–49 Ni,50 Cu,51 Zn,52,53 Al,54 Bi,55,56 Sb,57 Ag,58 Cd,59 and Fe,60–62 were used to manipulate the properties of SnS thin films and nanoparticles.
Studies on Pb incorporation in SnS thin films have highlighted their potential for optoelectronic applications due to enhanced photocurrent, carrier concentration, and mobility. Niknia et al. found that Pb2+ doping in α-SnS increased the grain size and reduced the bandgap, enhancing the photocurrent.63 Sebastian et al. observed that Pb2+ doping improved the grain size, carrier concentration, and Hall mobility in SnS thin films.64
Furthermore, Chalapathi et al. reported that Pb2+ doping in the cubic π-SnS phase altered structural, optical, and electrical properties, reducing the bandgap and increasing hole mobility.65 Additionally, PbS intermediate layers have been shown to be significantly beneficial for the epitaxial growth of π-SnS films on GaAs.66 Interestingly, a similar trend was observed in MBE-deposited SnSe thin films, where small amounts of Pb were shown to be instrumental for obtaining high-quality growth on GaAs substrates.67,68 Our recent work further demonstrated that Pb2+ promotes the compactness, phase purity, and growth rate of solution-deposited π-SnS thin films. Increased π-SnS content was obtained with increasing Pb2+ concentration in solution under conditions shown to otherwise result in pure α-SnS, the thermodynamically stable phase of tin monosulfide.15 We modelled the Pb impurities in both polymorphs using density functional theory (DFT). The calculated critical concentration of Pb cations to stabilize the π-SnS phase over the stable orthorhombic polymorph was 20 at%, in notable agreement with the experimental results.15
The search for alternative cations to stabilize the cubic π-phase of SnS over its orthorhombic polymorph remains important due to the toxic nature of previously used Pb cations. Identifying safer and equally effective cationic alternatives could support broader applications and meet environmental safety standards. Due to the similar oxidation state of Ca2+ in CaS and Pb2+ in PbS, along with their comparable rocksalt crystal structure and closely matched lattice constants, we hypothesized that Ca2+ can serve as an equally suitable cation for stabilizing the π-SnS phase. This hypothesis was proved to be correct both experimentally and computationally. In this work, we have computationally and experimentally explored the effect of calcium cation impurities on the stabilization and physical properties of the cubic π-SnS phase.
Methodology
Computational methods
Modelling of π-SnS and α-SnS was performed to calculate Ca defects using density functional theory (DFT) in the plane-wave pseudopotential approach. All calculations were performed using the first-principles DFT package Quantum ESPRESSO69 with pseudopotentials obtained from the Garrity, Bennett, Rabe, and Vanderbilt database.70 The Perdew–Burke–Ernzerhof (PBE) generalized gradient approximation71 was employed for exchange and correlation energy terms without relativistic corrections. Furthermore, norm-conserving pseudopotentials72 within GGA-PBE approximation were implemented to investigate the band structure and optical properties. Note that the PBE band gaps are underestimated due to the well-known band-gap error.73 However, the PBE functional captures the bandgap of SnS in reasonably good agreement with the HSE functional.74,75 A cutoff energy of 60 Ry in the plane-wave expansion of the wave functions was applied. Ca defects were calculated in 2 × 2 × 2 supercells of α-SnS with 64 atoms; however, unit cells of π-SnS with 64 atoms were considered for defect modelling. The k-point sampling employed in calculations for π-SnS and α-SnS was 4 × 4 × 4 and 8 × 8 × 4, respectively, and 16 × 16 × 16 for Sn and Ca metals.
The thermodynamic stability of Ca defects in both phases of SnS can be determined by the calculation of their formation energies. The formation energy of the defects can be calculated as73
| | | Uf[CaqSn] = Etot[CaqSn] − Etot[bulk] − niμCa + niμSn + q(EVBM + ΔEF + ΔV) | (1) |
where
Uf[Ca
qSn] is the formation energy of the defect Ca
Sn with charge
q,
Etot[Ca
qSn] is the total energy of a supercell containing a defect after the relaxation of the ion positions,
Etot[bulk] is the total energy of a bulk (perfect) supercell with the same number of atoms,
ni is the number of atoms added and removed from the supercell,
μCa and
μSn are the chemical potentials of Ca and Sn atoms, respectively, and Δ
EF is the Fermi level referenced to the valence band maximum (VBM)
EVBM. The VBM is referenced to the pure cell, and electrostatic potentials between the pristine cell and the defect supercell are aligned by adding a term Δ
V. This term, Δ
V, is taken as the difference in electrostatic potentials between the atomic site farthest from the defect in the defect supercell and its equivalent site in the defect-free cell.
76 For a neutral defect,
q equals zero in
eqn (1).
For a material with a diatomic basis, the chemical potential involved in defect formation energy calculations generally has an arbitrary degree of freedom, controlled by experimental boundary conditions. Formation energies per atom of π-SnS and α-SnS were calculated and found to be nearly the same (the difference is 5 × 10−3 eV) and in accord with the previous studies.15,16,77,78 Therefore, the chemical potential range of Sn and S should be the same in both phases. To maintain stable π-SnS and α-SnS compounds, the chemical potentials of Sn and S must satisfy equation:
| | | μSn + μS = ΔHf(SnSπ/α) + μbulkSn + μbulkS | (2) |
or equivalently,
| | | ΔμSn + ΔμS = ΔHf(SnSπ/α) | (3) |
where Δ
Hf(SnS
π/α) is the heat of formation of π-SnS and α-SnS.
However, to put reasonable limits on the chemical potentials of the Ca substitution defects, we consider restrictions that prevent not only the precipitation of the bulk phases of the dopant but also prevent the formation of phases of the dopant and the host atoms.
Moreover, the surface and interface energies were obtained using
eqn (5) and (6). The surface energy π-SnS is obtained from:
| |  | (5) |
and the interfacial energy:
| |  | (6) |
where
Etotslab(π-SnS) is the total energy of slab π-SnS,
Ebulk(π-SnS) is the total energy of the bulk of the π-SnS system having the same number of atoms,
Etotslab(π-SnS/CaS) is the total energy of the heterostructure system, and
A is the surface area of the slab.
Experimental methods
Materials.
CaCl2 (97%, Aldrich), SnCl2·2H2O (≥99.99%, Sigma-Aldrich), Ca(CH3COO)2·H2O (≥99.0%, Sigma-Aldrich), triethanolamine (TEA) (≥99.0%, Sigma-Aldrich), and thioacetamide (≥99.0%, Sigma-Aldrich) were used without further purification. Ammonium hydroxide (25% w/w), acetic acid (Bio-Lab, analytical grade), acetone (Bio-Lab, technical grade), and 2-propanol (Bio-Lab, 99.8%) were used without further purification. Distilled water (DIW) was obtained using a Millipore Direct Q3 water purification system. Monocrystalline GaAs (100) wafers (undoped, epi-polished with _0.11 miscut) were manufactured by AXT Inc and purchased from Geo Semiconductor (UK) Ltd. Fluorine-doped tin oxide (FTO) substrates were purchased from XOP GLASS (TEC 15, sheet resistance 12–14 Ohm cm2).
Substrate preparation.
Double-side polished GaAs(100) wafers were cleaved into 1.5 × 2 cm2 pieces and sonicated in aqueous detergent solution for 10 min at room temperature. The wafers were then washed with DIW, acetone, and 2-propanol in that order. Finally, the substrates were dried under a N2 flow. The FTO substrates were treated similarly.
Deposition procedure.
(a) Surface pretreatment: surface pretreatment was carried out based on our previous experience with Pb2+ cations where treatment in Pb2+ solution was shown to activate the substrate surface.15,44 In the present study, 0.35 M calcium chloride solution was prepared in a 50 ml beaker and degassed under N2 purging for one hour. The beaker was then placed in a thermostatic bath at 50 °C and the substrates were immersed in the solution for one day, then the substrates were rinsed with water and immediately placed into the deposition bath. This pretreatment of the substrates in the calcium chloride solution resulted in adsorption of Ca2+ cations on the GaAs substrates, as shown in Fig. S1, SI.
(b) Thin film preparation: for the deposition of the SnS thin film, the aqueous solution was prepared containing 0.04 M SnCl2·2H2O, 0.5 M acetic acid, 1.1 M TEA, and 0.4 M ammonium hydroxide with a final volume of 50 ml. The TEA and ammonium hydroxide serve as complexing agents, binding to the metal cations and preventing them from quickly reacting or precipitating out of the solution. For introducing Ca2+, different amounts of Ca(CH3COO)2·H2O were added in the solution. The substrates were immersed in the deposition solution immediately after adding thioacetamide to reach a final concentration of 0.08 M. Depositions were carried out at 40 °C for 3 hours.
Characterization.
Grazing incidence X-ray diffraction (GIXD) measurements were performed on thin films of the samples using a Malvern Panalytical Empyrean III diffractometer equipped with a position sensitive PIXcel3D detector and iCore/dCore automated optics combination. Cu Kα radiation (λ = 1.5405 Å) was used as the incident radiation, with a fixed incident beam angle of 0.5°. Two GIXD scans were performed: a standard scan with a step size of 2θ = 0.10° was collected for all samples over a 2θ range of 20–50° with three sequential measurements collected and summed to produce the final GIXD diffractogram. A high-resolution GIXD follow-up scan was also performed on select samples in the 2θ range of 20–37.5°, with a step size of 0.08° and six sequential measurements collected and summed. High resolution scanning electron microscopy (HRSEM) was carried out using an FEI Verios 460L HRSEM. Topography images were taken using 3 keV to 5 keV acceleration voltages and beam currents of 25 pA to 50 pA. The measurements were carried out without coating the samples. Transmission electron microscopy (TEM) was performed using a JEOL JEM-2100F operating at an accelerating voltage of 200 kV. EDS analysis was carried out in the TEM using an Oxford Instruments X-Max 65T SDD detector. AZtec software (v. 3.3) was used for EDS data analysis. X-ray photoelectron spectroscopy (XPS) measurements were carried out using an ESCALAB 250 spectrometer with a monochromatic Al X-ray source (excitation energy 1486.6 eV) at a base pressure of 1 × 10−9 mbar. Raman spectroscopy was carried out on a LabRAM HR Evolution (HORIBA, France) micro-Raman system equipped with cold stage and a rotating polarizer. The excitation source was a 532 nm HeNe laser with a power of 3 mW on the sample. Optical absorption measurements were performed in the range 400–1000 nm using a Cary 5000 UV-vis-NIR spectrophotometer equipped with a PbSmart detector. Samples for inductively coupled plasma optical emission spectrometry (ICP-OES) were prepared by initially cleaning the backside of the FTO substrates with dilute acid, and then dissolving the SnS films in an aqueous solution of 70% HNO3. The resulting solutions were analyzed using a SPECTRO ARCOS ICP-OES spectrometer manufactured by Spectro Ametek.
Results and discussion
The optimized lattice constants and bandgap for cubic π-SnS and orthorhombic α-SnS were calculated (see Table 1) and found to be a = 11.885 Å and a = 4.023 Å, b = 4.441 Å, c = 11.429 Å, in accord with reported experimental values a = 11.594 Å, and a = 3.987 Å, b = 4.334 Å, and c = 11.20 Å,79 and other theoretical studies reported a = 11.885 Å, and a = 4.03 Å, b = 4.42 Å, and c = 11.41 Å.80 Moreover, an indirect bandgap was obtained for cubic π-SnS and orthorhombic α-SnS, i.e., 1.31 eV and 1.04 eV, respectively, in good agreement with a previous theoretical result of 1.34 eV81 and 1.06 eV,81 respectively, slightly underestimating the reported experimental results of 1.53 eV82 and 1.25 eV,82 as might be expected.
Table 1 Lattice constants (Å) and band gap (eV) calculated for π-SnS and α-SnS
| Lattice constant (Å) |
| SnS phase |
Our calculation |
Previous calculation8 |
Experimental value7 |
| π-SnS |
a = 11.885 |
a = 11.885 |
a = 11.594 |
| α-SnS |
a = 4.023 |
a = 4.03 |
a = 3.987 |
|
b = 4.441 |
b = 4.42 |
b = 4.334 |
|
c = 11.429 |
c = 11.41 |
c = 11.20 |
| Band gap (eV) |
| π-SnS |
1.34 |
1.349 |
1.5310 |
| α-SnS |
1.04 |
1.069 |
1.2510 |
The crystal structures of π-SnS and α-SnS are illustrated in Fig. 1. The low-symmetry π-SnS unit cell with 64 atoms is depicted in Fig. 1 and compared with a 2 × 2 × 2 supercell of α-SnS, which contains 8 atoms in each unit cell, altogether 64 atoms in the supercell. The substitutional sites in π-SnS and α-SnS are highlighted in red in Fig. 1.
 |
| | Fig. 1 Schematic representation of π-SnS and α-SnS with four Ca substitutions at the Sn sites in both supercell models. | |
The thermodynamic stability of Ca incorporation at selected concentrations was examined in both phase systems by calculating the formation energy using eqn (1). From energetic calculations, the difference in the formation enthalpy of α-SnS and π-SnS is very small, 5 × 10−3 eV atom−1. Therefore, the chemical potential of the components (Sn and S) in both systems can be considered equal. Hence, doping with suitable elements at sufficient concentrations may cause π-SnS to be preferred over the ground state α-SnS phase. Given the structural similarity between PbS and CaS, we hypothesized that the Ca2+ cation may serve a role similar to that of Pb2+ cations in SnS systems. Incorporating Ca2+ cations could thermodynamically stabilize the π-SnS phase in comparison to the ground state α-SnS phase. To test this hypothesis, we recalculated the formation enthalpy of both systems for selected concentrations of calcium cations.
The defect formation energies of Ca incorporated at Sn substitutional sites (using the conventional chemical potential of Ca from the FCC elemental phase) were calculated for α-SnS and π-SnS (Table 2). We found negative defect formation energies of Ca substitution in both systems, with a difference of 0.09 eV in favour of the π-SnS phase. Comparing this result with the previous study for Pb, the energy difference obtained was thrice as high for Ca incorporation (in favour of π-SnS),15 which results in Ca being a more suitable cation for stabilizing π-SnS, requiring relatively lower substitutional concentration.
Table 2 The formation energies (U) of Ca substitution in π-SnS and α-SnS, the difference of the formation energy (ΔU) between π-SnS and α-SnS, the difference of formation energy per atom (ΔE), difference in total enthalpy per atom (ΔH) and equilibrium concentrations (c) of Ca for selected fixed concentrations
|
c
fix
|
U
f[CaSn] (eV) |
ΔUf[CaSn] (meV) |
ΔEf[CaSn] (meV atom−1) |
ΔHfπ-SnS/α-SnS[CaSn] (meV atom−1) |
c (cat%) |
| π-SnS |
α-SnS |
| 0 |
N/A |
N/A |
N/A |
N/A |
5 |
N/A |
| 3.13 |
−1.86 |
−1.77 |
−90 |
−1.4 |
3.6 |
11.2 ± 1 |
| 12.5 |
−7.87 |
−7.52 |
−350 |
−5.5 |
−0.5 |
11.4 ± 3.5 |
For CaSn substitutions to facilitate the transformation of the stable SnS phase from α-SnS to π-SnS, the formation enthalpy of π-SnS (ΔHfπ-SnS[CaSn]) should be lower than that of α-SnS (ΔHfα-SnS[CaSn]), i.e., ΔHfπ-SnS[CaSn] < ΔHfα-SnS[CaSn].
We calculated the defect formation energies of Ca for selected concentrations. Ca substitution in π-SnS was found to be more energetically favorable than in α-SnS, with energy gains of 0.09 eV and 0.35 eV for 3.13 cat% and 12.5 cat% substituted Ca atoms in the simulated supercell, respectively.
We determined the equilibrium Ca defect concentration, c, at which the π-SnS phase becomes more stable than the α-SnS phase. Assuming a linear relationship between defect concentration and formation energy, the relative formation energy of the SnS phases with Ca substitutional defects is expressed by eqn (7).
| | | ΔHfπ-SnS/α-SnS[CaSn] = ΔHfπ-SnS/α-SnS + c′ΔEf[CaSn] | (7) |
where Δ
Hfπ-SnS/α-SnS[Ca
Sn] represents the difference in lattice formation energies per atom between π-SnS and α-SnS containing Ca defects at concentration
c. Δ
Hfπ-SnS/α-SnS denotes the formation energy difference per atom between pure π-SnS and α-SnS, while Δ
Ef[Ca
Sn] corresponds to the difference in formation energies per atom of π-SnS and α-SnS with concentration
cfix of Ca.
Setting the left-hand side of eqn (7) to zero allows us to estimate the desired concentration, c. The calculated enthalpy differences and c′ (calculated using eqn (8)) for selected values of cfix are reported in Table 2. We calculated critical concentration, c, for cfix of 3.13 cation percent (cat%, namely the percentage of Ca cations out of the total number of cations) Ca substitution to be 11.2 at% (complete calculations can be found in the SI). The difference in the lattice formation energy (ΔHfπ-SnS/α-SnS[CaSn]) of both phases was found to decrease to 4.65 meV atom−1, favoring α-SnS. For a fixed Ca concentration of 12.5 cat%, the difference in lattice formation energy nearly vanished, with a value of 0.4 meV atom−1—below the computational error limit. We used these data to calculate the critical equilibrium concentration value c using eqn (7) and found it to be 11.4 cat%.
For 25 cat% substitution of Ca, the total energy of π-SnS was found to be lower than that of α-SnS and became the ground state phase of SnS as expected (Table 2). Furthermore, we determined the optimized lattice constants for 25 cat% Ca (equivalent to eight Ca substitutions) in both π-SnS and α-SnS. Our results show that the lattice constants remain almost unchanged compared to the pristine material, with a difference of approximately 1% for π-SnS.
Finally, we calculated the bandgap of the bulk and Ca-substituted systems, and their values are reported in Table 3. The bandgap of π-SnS was found to decrease slightly with increasing Ca substitution concentrations in the system (Table 3), as confirmed by the experimental measurements reported below. The calculated bandgap values underestimate the experimental results (Fig. 4), which is a known limitation of the PBE-GGA approximation (Table 1).
Table 3 Bandgap (eV) of π-SnS with Ca substitution for selected fixed concentrations (cat%)
| Ca substitution (cat%) |
Bandgap (eV) |
| 0 |
1.34 |
| 3.13 |
1.33 |
| 12.5 |
1.22 |
| 25 |
1.05 |
Based on these computational findings, we moved to experimentally study the effect of Ca2+ incorporation on the phase stability of SnS thin films synthesized by CD.44 The amount of Ca(CH3COO)2·H2O (Ca/Sn molar ratio ranges from 0.02 to 0.14) was changed to vary the concentrations of Ca2+ in the reaction solution. HRSEM of the pristine SnS thin film exhibits vertically stacked platelet morphology (Fig. 2a). Fig. 2b–d represent the gradual changes in morphology of the deposited SnS films upon addition of Ca2+ cations to the deposition bath. The morphology remains unchanged up to 6% Ca2+ (mole percentage; Fig. 2b), however at higher concentrations of Ca2+, a clear change in film morphology is observed. At 10% Ca2+ cation concentrations, rounded grains along with the vertical platelets start to appear (Fig. 2c). At 12% Ca2+ concentrations, the film morphology is exclusively composed of rounded grains (Fig. 2d). At 14% Ca2+ cation concentrations, HRSEM indicated the formation of large box shaped particles underneath the film (Fig. S2a, SI). Cross-sectional TEM analysis was carried out to understand the nature of the large box shape particles. TEM images indicate the formation of a distinct layer inner to the SnS layer (Fig. S2b, SI). Analyzing the composition of the layer using EDS in TEM indicated that the atomic concentration ratio in the additional layer is Ca
:
Sn
:
O = 20
:
15
:
65. However, we could not obtain any SAED pattern from the layer, which pointed out to the amorphous/very weak crystalline nature of the layer (Fig. S2c, SI). The vertically stacked platelets and the rounded grains are typical, characteristic morphologies of the α-SnS and π-SnS thin film, respectively.15,44 Hence, HRSEM suggested the formation of a different phase obtained upon incorporation of increasing concentrations of Ca2+cations.
 |
| | Fig. 2 HRSEM images of SnS thin films deposited onto GaAs(100) substrates from solutions containing (a) 0% Ca2+, (b) 6% Ca2+, (c) 10% Ca2+, and (d) 12% Ca2+. All scale bars correspond to 3 μm. | |
To verify the phase composition upon Ca2+ addition, GIXD measurements were performed (Fig. 3). The GIXD pattern of the film with 0% Ca2+ (black curve, Fig. 3a) exhibits a single peak corresponding to the (111) plane of the orthorhombic crystallographic phase (α-phase) (COD file #1011253; space group: Pmcn (62)). The GIXD pattern of the film with 6% Ca2+ exhibits somewhat improved crystallinity, however it still exhibits the pure α-phase (red curve, Fig. 3a), which is consistent with the HRSEM images. Upon increasing the Ca2+ concentration to 10%, a new peak appears at 2θ = 26.7°, corresponding to the (222) plane of the cubic π-phase of SnS (ICDD # 04-023-2702; space group: P213 (198)). The GIXD pattern of the film with 10% Ca2+ also exhibits a broad peak in the 2θ range of 30.7–31.9° (blue curve, Fig. 3a). From the high-resolution measurement in the small 2θ range (20°–36°) (Fig. 3b), we can distinguish the peak corresponding to the (400) plane of the π-phase, which nearly overlaps with that corresponding to the (111) plane of the α-phase, contributing together to the originally observed broad peak. At 12% Ca2+ concentration, only peaks corresponding to the π-phase exist (green curve, Fig. 3a).44 The observed changes in the XRD patterns of the films deposited in the presence of different concentrations of Ca2+ cations confirm the stabilization of the π-phase over the α-phase and are consistent with the HRSEM results.
 |
| | Fig. 3 (a) GIXD pattern of SnS thin films deposited onto GaAs (100) substrates in the presence of 0% Ca2+ (black curve), 6% Ca2+ (red curve), 10% Ca2+(blue curve), and 12% Ca2+ (green curve). (b) GIXD pattern of the SnS thin film deposited with 10% Ca2+ over the 2θ range of 20–36°. | |
Raman spectroscopy was employed to further investigate the SnS films deposited with various concentrations of Ca2+ cations. The Raman spectra (Fig. 4a) of the SnS film with 0% Ca2+ exhibit peaks at 93 cm−1, 160 cm−1, and 227.6 cm−1 corresponding to the Ag and B3g modes of the α-SnS.44 In the case of 6% Ca2+, the Raman spectrum (Fig. 4a) exhibits similar peaks corresponding to the Ag and B3g modes of α-SnS. At 10% Ca2+ concentrations, new peaks at 193 cm−1 and 224 cm−1 corresponding to the Ag mode of π-SnS11 appeared along with the peak at 93 cm−1 of α-SnS (Fig. 4a). The Raman spectrum of the film with 12% Ca2+ concentration exhibits peaks at 87 cm−1, 176.9 cm−1, and 195 cm−1, corresponding to the B1g, B3g, and Ag modes of π-SnS (Fig. 4a).11 The peak positions obtained are in good agreement with previous Raman measurements for α-SnS and π-SnS, respectively.11 The Raman spectroscopy results reaffirmed the HRSEM and GIXD results in Fig. 2 and 3, respectively.
 |
| | Fig. 4 (a) Raman spectra of SnS films deposited onto GaAs(100) substrates in the presence of 0% Ca2+ (cyan curve), 6% Ca2+ (violet curve), 10% Ca2+ (maroon curve), and 12% Ca2+ (blue curve). (b) UV-vis absorption spectra of the SnS film with 12% Ca2+ (π-SnS) on FTO coated glass substrates. Inset: Tauc plot indicates an indirect band gap of 1.57 eV for the SnS film deposited with 12% Ca2+ which matches with the reported value of π-SnS. (c) UV-vis absorption spectra of the SnS film deposited from the solution following the protocol which generally results in the formation of π-SnS. Inset: Tauc plot indicates an indirect band gap of 1.61 eV demonstrating a slight decrease in the bandgap as a result of Ca-incorporation in π-SnS. | |
Subsequently, we investigated the optical properties of the films deposited with different Ca2+ concentrations in solution using UV-vis absorption spectroscopy (Fig. S3, SI). Fig. 4b shows the UV-vis absorption spectra of π-SnS films deposited from a solution containing 12% Ca2+ onto FTO coated glass substrates. Tauc plot analysis shows an indirect band gap of 1.57 eV for the film deposited in the presence of 12% Ca2+ (inset of Fig. 4b), which is in good agreement with the literature values previously reported for π-SnS.11,15,83 For comparison, we have also measured the absorbance spectra of the π-SnS film deposited following a deposition protocol that generally produces a π-phase SnS film (Fig. 4c). Corresponding Tauc plot analysis shows an indirect band gap of 1.61 eV for the π-SnS film. In summary, this value points out to a slight decrease in the bandgap of π-SnS upon Ca2+ incorporation.
XPS measurements were carried out to quantitatively evaluate the amount of Ca2+ incorporated into the SnS films. For Ca2+ concentrations lower than 10% Ca2+ in the bath, we are unable to quantify the presence of Ca2+ in the film. However, atomic % calculation from the XPS measurements of the SnS film, deposited with 12% of Ca2+, clearly points to the presence of Ca2+ in the SnS film with a Ca to Sn ratio of ∼0.02
:
1 (2 cat%). The result is in qualitative agreement with the theoretical calculation, although they differ quantitatively as the data in Table 2 predict a higher critical concentration of Ca2+ ions of 11 cat%. The pre-treatment of the substrate with Ca2+ cations, followed by subsequent reaction with sulfide in the reaction bath is likely to result in the formation of an ultrathin cubic (rocksalt) CaS intermediate layer which is lattice-matched with the GaAs substrate (less than 1% mismatch) and is likely to strongly stabilize the surface of the incipient π-SnS, yet this could not be verified experimentally, probably due to the ultrathin nature and/or poor crystallinity of the intermediate layer. Computational studies clearly support the above assumption. We modelled the epitaxial growth of a π-SnS(111) overlayer on rocksalt CaS(111) substrate (Fig. 6). The interfacial energy of the π-SnS(111)/CaS(111) heterostructure and the surface energies of π-SnS(111) and CaS(111) were calculated using eqn (5) and (6). The results are given in Table 4. Our calculation reveals that the interfacial energy of the π-SnS(111)/CaS(111) interface is lower than the surface energies of both π-SnS(111) and CaS(111), indicating thermodynamic preference for interface formation over surface exposure. The gain in energy from the difference in interfacial and surface energies is approximately 6.4 meV Å−2, which corresponds to 104 meV per surface atom. This energy gain is sufficient to stabilize the π-SnS phase without incorporation of significant Ca in the bulk. However, experimentally we found that surface treatment is not solely sufficient for stabilizing the π-phase. Nevertheless, Ca substitution is energetically favorable, which suggests that limited incorporation of Ca atoms can occur within the bulk of π-SnS. The survey spectrum of the SnS film deposited with 12 cat% of Ca2+ is shown in Fig. S4a, SI. The high-resolution Sn 3d XPS spectrum reveals peaks at 486.4 eV and 494.8 eV, which suggest that there is surface Sn bonding with O (Fig. S4b, SI).15 The mild surface oxidation of the SnS film is typical for solution-deposited thin films.15 We have therefore removed the surface oxide layer from the SnS sample by mild etching using Ar ions. After etching, the Sn 3d core-level spectra display peaks at 493.9 eV and 485.5 eV corresponding to the Sn 3d3/2 and Sn 3d5/2, respectively (Fig. 5a). The doublet nature of both components is the result of Ar ion bombardment which leads to the formation of reduced states due to partial destruction of bonds.15,84 The peaks corresponding to the Sn dangling bonds appear at 492.8 eV and 484.5 eV (Fig. 5a). S 2p core-level spectra reflect peaks at 161.8 eV and 160.7 eV corresponding to S 2p1/2 and S 2p3/2, respectively (Fig. 5b). Ca 2p core-level spectra show peaks at 349.9 eV and 346.3 eV corresponding to the Ca 2p1/2 and Ca 2p3/2, respectively (Fig. 5c). O 1s core-level spectra show a peak at 529.5 eV (Fig. 5d).
 |
| | Fig. 5 XPS core-level spectra obtained from an SnS film deposited from a solution with 12 cat% Ca2+, showing (a) Sn 3d (b) S 2p (c) Ca 2p and (d) O 1s. | |
 |
| | Fig. 6 Structural models used to calculate surface energies: (a) π-SnS(100) slab and (b) interfacial energy CaS(111)/π-SnS(111). The blue, red and yellow atoms represent tin, calcium and sulfur atoms, respectively. Heterostructure CaS(111)/π-SnS(111) is constructed with both anion (sulfur) and cation (Ca or Sn) termination surfaces of CaS(111) and π-SnS(111). | |
Table 4 Surface and interfacial energies (meV Å−2) for π-SnS and CaS systems
| Surface/interface |
Surface/interface energy (meV Å−2) |
Previous results (meV Å−2) |
| π-SnS(111) (S-termination) |
32.5 |
32.5,66 32.478 |
| CaS(111) (Ca-termination) |
49.5 |
— |
| π-SnS(111)/CaS(111) |
26.1 |
— |
To confirm the amount of Ca incorporated in the film, we performed ICP-OES after dissolving the films in 70% nitric acid. The relative concentrations of the Ca incorporated in the films deposited using 0–12% Ca2+ in the solutions are provided in Table S1, SI. The ICP-OES results indicated that the Ca2+ concentration in the film deposited with 12% Ca2+ in the solution was 3.92% (slightly higher than the concentration of one substitutional atom which is 3.125% or 1/32), in reasonable agreement with the concentration obtained from XPS.
Stabilization of the π-SnS phase with the assistance of Ca2+ offers notable advantages over the previously studied Pb2+ cations.15 Pb substitution, while effective, is environmentally undesirable and requires a relatively high concentration (ca. ∼20 cat%) to stabilize the π-SnS phase. In comparison, Ca2+ offers environmental sustainability and is more effective, theoretically requiring ∼11 cat% to stabilize the π-SnS phase. Notably, similar to PbS substrates, CaS can also serve as a potential substrate for the epitaxial growth of π-SnS overlayers.66 This is likely to explain the relatively low experimentally measured concentration of approximately 3.92 cat% Ca in the π-SnS phase because surface stabilization is likely to be taking place in addition to bulk stabilization.
Conclusions
The cubic polymorph π-SnS was stabilized with the assistance of non-toxic Ca2+ cations. The free enthalpy of Ca incorporation in both SnS phases was calculated using DFT. The defect formation energies for Ca substitution indicate that its incorporation is energetically stable and relatively more favorable in π-SnS. We calculated the critical concentration of Ca substitution, which can lead to the phase stabilization of SnS from α-SnS to π-SnS and found it to be ∼11 cat%. We demonstrated the theoretical prediction of phase stabilization using CD methods and found that Ca2+-cation incorporation can lead to stabilization of π-SnS over α-SnS. Experimental characterization using HRSEM, GIXD, and Raman spectroscopy confirmed the computational results. Interestingly, concentrations of 3.92 cat% Ca2+ appear to be sufficient for stabilizing the π-SnS phase and favoring it over the thermodynamically preferred α-SnS. It is likely that the effect of adsorbed Ca onto the substrate can also contribute to the stabilization of π-SnS, decreasing the need for Ca in the bulk by means of substitutional incorporation. The small concentrations of Ca2+ in the films did not significantly alter the optical properties of π-SnS.
Author contributions
Dr Neeraj Mishra (computational studies) and Dr Susmita Paul (experimental studies) contributed equally. The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the SI. See DOI: https://doi.org/10.1039/d5qm00399g.
Raw data are available from the repository of the Ilse Katz Institute for Nanoscale Science and Technology at BGU.
Acknowledgements
This work was funded by the Israel Science Foundation, Grant no. 512/23. We thank Dr Lee Shelly (XPS), Dr Mariela Pavan (Raman spectroscopy), Dr Nitzan Maman (FIB-HRSEM), Dr Bar Elisha (ICP-OES) and Dr Vladimir Ezersky (TEM) for expert assistance. S. P. extends sincere appreciation to Dr Noy Zakay for her invaluable assistance.
References
- R. Banai, M. Horn and J. Brownson, A review of tin(II) monosulfide and its potential as a photovoltaic absorber, Sol. Energy Mater. Sol. Cells, 2016, 150, 112–129 CrossRef CAS.
- B. Zhou, S. Li, W. Li, J. Li, X. Zhang, S. Lin, Z. Chen and Y. Pei, Thermoelectric properties of SnS with Na-doping, ACS Appl. Mater. Interfaces, 2017, 9, 34033–34041 CrossRef CAS.
- J. A. Andrade-Arvizu, M. Courel-Piedrahita and O. Vigil-Galán, SnS-based thin film solar cells: perspectives over the last 25 years, J. Mater. Sci.: Mater. Electron., 2015, 26, 4541–4556 CrossRef CAS.
- K. R. Reddy, N. K. Reddy and R. Miles, Photovoltaic properties of SnS based solar cells, Sol. Energy Mater. Sol. Cells, 2006, 90, 3041–3046 CrossRef.
- G. A. Tritsaris, B. D. Malone and E. Kaxiras, Optoelectronic properties of single-layer, double-layer, and bulk tin sulfide: A theoretical study, J. Appl. Phys., 2013, 113, 233507 CrossRef.
- R. W. Miles, O. E. Ogah, G. Zoppi and I. Forbes, Thermally evaporated thin films of SnS for application in solar cell devices, Thin Solid Films, 2009, 517, 4702–4705 CrossRef CAS.
- D. Mudusu, K. R. Nandanapalli, S. R. Dugasani, R. Karuppannan, G. K. R. Reddy, R. G. E. Subramanian and S. H. Park, Metal–insulator–semiconductor field-effect transistors (MISFETs) using p-type SnS and nanometer-thick Al2S3 layers, RSC Adv., 2017, 7, 11111–11117 RSC.
- T. S. Reddy and M. C. S. Kumar, Co-evaporated SnS thin films for visible light photodetector applications, RSC Adv., 2016, 6, 95680–95692 Search PubMed.
- Y. Qin, Y. Liang, C. Zhou and Y. Bai, Homogeneous heterophase junction based on cubic/orthomorphic SnS for chemiresistive room temperature trace-ethanol sensor, Sens. Actuators, B, 2024, 404, 135285 CrossRef CAS.
- Z. Wei, L. Wang, M. Zhuo, W. Ni, H. Wang and J. Ma, Layered tin sulfide and selenide anode materials for Li-and Na-ion batteries, J. Mater. Chem. A, 2018, 6, 12185–12214 RSC.
- R. E. Abutbul, E. Segev, L. Zeiri, V. Ezersky, G. Makov and Y. Golan, Synthesis and properties of nanocrystalline π-SnS–a new cubic phase of tin sulphide, RSC Adv., 2016, 6, 5848–5855 RSC.
- A. Rabkin, S. Samuha, R. E. Abutbul, V. Ezersky, L. Meshi and Y. Golan, New nanocrystalline materials: a previously unknown simple cubic phase in the SnS binary system, Nano Lett., 2015, 15, 2174–2179 CrossRef CAS PubMed.
- I. Pallikara and J. M. Skelton, Phase stability of the tin monochalcogenides SnS and SnSe: a quasi-harmonic latticedynamics study, Phys. Chem. Chem. Phys., 2021, 23, 19219–19236 RSC.
- R. E. Abutbul, E. Segev, S. Samuha, L. Zeiri, V. Ezersky, G. Makov and Y. Golan, A new nanocrystalline binary phase: synthesis and properties of cubic Tin monoselenide, CrystEngComm, 2016, 18, 1918–1923 RSC.
- N. Zakay, N. Mishra, N. Maman, A. Rashkovskiy, A. Schlesinger, R. E. Abutbul, V. Ezersky, D. Azulay, G. Makov and Y. Golan, Phase control in solution deposited tin monosulfide thin films: The role of Pb2+ cations, Mater. Chem. Front., 2023, 7, 3714–3727 RSC.
- E. Segev, U. Argaman, R. E. Abutbul, Y. Golan and G. Makov, A new cubic prototype structure in the IV–VI monochalcogenide system: A DFT Study, CrystEngComm, 2017, 19, 1751–1761 RSC.
- V. E. González-Flores, R. N. Mohan, R. Ballinas-Morales, M. Nair and P. Nair, Thin film solar cells of chemically deposited SnS of cubic and orthorhombic structures, Thin Solid Films, 2019, 672, 62–65 CrossRef.
- P. Jain and P. Arun, Influence of grain size on the band-gap of annealed SnS thin films, Thin Solid Films, 2013, 548, 241–246 CrossRef CAS.
- B. Ghosh, R. Bhattacharjee, P. Banerjee and S. Das, Structural and optoelectronic properties of vacuum evaporated SnS thin films annealed in argon ambient, Appl. Surf. Sci., 2011, 257, 3670–3676 Search PubMed.
- F. De. B. Sánchez, M. T. S. Nair and P. K. Nair, Insights to the production of SnS-cubic thin films by vacuum thermal evaporation for photovoltaics, Semicond. Sci. Technol., 2024, 39, 015003 CrossRef.
- M. Calixto-Rodriguez, H. Martinez, A. Sanchez-Juarez, J. Campos-Alvarez, A. Tiburcio-Silver and M. Calixto, Structural, optical, and electrical properties of tin sulfide thin films grown by spray pyrolysis, Thin Solid Films, 2009, 517, 2497–2499 CrossRef CAS.
- E. Sarica, Investigation of spray pyrolyzed cubic structured Cu doped SnS films, Phosphorus, Sulfur Silicon Relat. Elem., 2021, 196, 1103–1108 Search PubMed.
- T. Sall, M. Mollar and B. Marí, Substrate influences on the properties of SnS thin films deposited by chemical spray pyrolysis technique for photovoltaic applications, J. Mater. Sci., 2016, 51, 7607–7613 CrossRef CAS.
- N. Sato, M. Ichimura, E. Arai and Y. Yamazaki, Characterization of electrical properties and photosensitivity of SnS thin films prepared by the electrochemical deposition method, Sol. Energy Mater. Sol. Cells, 2005, 85, 153–165 CrossRef CAS.
- Z. Zainal, M. Z. Hussein and A. Ghazali, Cathodic electrodeposition of SnS thin films from aqueous solution, Sol. Energy Mater. Sol. Cells, 1996, 40, 347–357 Search PubMed.
- R. Banai, J. Cordell, G. Lindwall, N. Tanen, S.-L. Shang, J. Nasr, Z.-K. Liu, J. Brownson and M. Horn, Control of phase in tin sulfide thin films produced via RF-sputtering of SnS2 target with post-deposition annealing, J. Electron. Mater., 2016, 45, 499–508 Search PubMed.
- V. K. Arepalli and J. Kim, Effect of substrate temperature on the structural and optical properties of radio frequency sputtered tin sulfide thin films for solar cell application, Thin Solid Films, 2018, 666, 34–39 CrossRef CAS.
- V. K. Arepalli, Y. Shin and J. Kim, Photovoltaic behavior of the room temperature grown RF-Sputtered SnS thin films, Opt. Mater., 2019, 88, 594–600 CrossRef CAS.
- A. A. Voznyi, O. V. Bilousov, B. R. Landeke-Wilsmark, J. Keller, J. Ren, S.-L. Zhang and C. Hägglund, Ultrathin solar cells based on atomic layer deposition of cubic versus orthorhombic tin monosulfide, ACS Appl. Energy Mater., 2021, 4, 8085–8097 CrossRef CAS.
- H. Kafashan, M. Azizieh and H. N. Vatan, Ultrasound-assisted electrodeposition of SnS: Effect of ultrasound waves on the physical properties of nanostructured SnS thin films, J. Alloys Compd., 2016, 686, 962–968 CrossRef CAS.
- P. O'Brien and J. McAleese, Developing an understanding of the processes controlling the chemical bath deposition of ZnS and CdS, J. Mater. Chem., 1998, 8, 2309–2314 RSC.
-
G. Hodes, Chemical solution deposition of semiconductor films, CRC Press, 2002 Search PubMed.
- J. A. Switzer and G. Hodes, Electrodeposition and chemical bath deposition of functional nanomaterials, MRS Bull., 2010, 35, 743–750 Search PubMed.
- R. Mane and C. Lokhande, Chemical deposition method for metal chalcogenide thin films, Mater. Chem. Phys., 2000, 65, 1–31 CrossRef CAS.
- D. Lincot, Solution chemistry and semiconductors: from heterogeneous precipitation to photovoltaic industry, Actual. Chim., 1999, 5, 23–30 Search PubMed.
- P. Nair, M. Nair, V. Garcıa, O. Arenas, Y. Pena, A. Castillo, I. Ayala, O. Gomezdaza, A. Sanchez and J. Campos, Semiconductor thin films by chemical bath deposition for solar energy related applications, Sol. Energy Mater. Sol. Cells, 1998, 52, 313–344 CrossRef CAS.
- A. Osherov, V. Ezersky and Y. Golan, The role of solution composition in chemical bath deposition of epitaxial thin films of PbS on GaAs(100), J. Cryst. Grow., 2007, 308, 334–339 Search PubMed.
- M. Cao, X. Zhang, J. Ren, Y. Sun, Y. Cui, J. Zhang, J. Ling, J. Huang, Y. Shen and L. Wang, Chemical bath deposition of SnS: Cu/ZnS for solar hydrogen production and solar cells, J. Alloys Compd., 2021, 863, 158727 Search PubMed.
- L. Rodríguez-Guadarrama, J. Escorcia-García, I. Alonso- Lemus and J. Campos-Álvarez, 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 CrossRef.
- E. Guneri, C. Ulutas, F. Kirmizigul, G. Altindemir, F. Gode and C. Gumus, 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 Search PubMed.
- A. Tanusevski, Optical and photoelectric properties of SnS thin films prepared by chemical bath deposition, Semicond. Sci. Technol., 2003, 18, 501–505 CrossRef CAS.
- J. Jing, M. Cao, C. Wu, J. Huang, J. Lai, Y. Sun, L. Wang and Y. Shen, Chemical bath deposition of SnS nanosheet thin films for FTO/SnS/CdS/Pt photocathode, J. Alloys Compd., 2017, 726, 720–728 CrossRef CAS.
- M. S. Mahdi, H. S. Al-Arab, H. S. Al-Salman, K. Ibrahim, N. M. Ahmed, A. Hmood and M. Bououdina, A high performance near-infrared photodetector based on π-SnS phase, Mater. Lett., 2020, 273, 127910 CrossRef CAS.
- R. E. Abutbul and Y. Golan, Chemical epitaxy of π-phase cubic tin monosulphide, CrystEngComm, 2020, 22, 6170–6181 Search PubMed.
- M. S. Mahdi, K. Ibrahim, A. Hmood, N. M. Ahmed, F. I. Mustafa and S. A. Azzez, High performance near infrared photodetector based on cubic crystal structure SnS thin film on a glass substrate, Mater. Lett., 2017, 200, 10–13 Search PubMed.
- V. P. Jethwa, K. Patel, V. Pathak and G. Solanki, Enhanced electrical and optoelectronic performance of SnS crystal by Se doping, J. Alloys Compd., 2021, 883, 160941 Search PubMed.
- H. Kafashan, Optoelectronic properties of In-doped SnS thin films, Ceram. Int., 2019, 45, 334–345 CrossRef CAS.
- F. Jamali-Sheini, M. Cheraghizade and R. Yousefi, Ultrasonic synthesis of In-doped SnS nanoparticles and their physical properties, Solid State Sci., 2018, 79, 30–37 CrossRef CAS.
- M. Asghar, H. Ullah, M. W. Iqbal and Y.-H. Shin, Doping dependent optoelectronic, and magnetic properties of monolayer SnS, Mater. Sci. Semicond. Process., 2022, 152, 107049 CrossRef CAS.
- I. Ammar, A. Gassoumi and N. Turki-Kamoun, The effect of TSC and nickel doping on SnS thin films, Silicon, 2021, 13, 1933–1938 CrossRef CAS.
- S. L. Mousavi, F. Jamali-Sheini, M. Sabaeian and R. Yousefi, Correlation of physical features and the photovoltaic performance of P3HT: PCBM solar cells by Cu-doped SnS nanoparticles, J. Phys. Chem. C, 2021, 125, 15841–15852 CrossRef CAS.
- W. Lei, F. Wang, B. Lu, Z. Ye and X. Pan, Zn-doped SnS with sulfur vacancies for enhanced photocatalytic hydrogen evolution from water, New J. Chem., 2022, 46, 17791–17800 RSC.
- F. Niknia, F. Jamali-Sheini and R. Yousefi, Examining the effect of Zn dopant on physical properties of nanostructured SnS thin film by using electrodeposition, J. Appl. Electrochem., 2016, 46, 323–330 Search PubMed.
- S. Sebastian, I. Kulandaisamy, A. Arulanantham, S. Valanarasu, A. Kathalingam, A. J. Jebathew, M. Shkir and M. Karunakaran, Influence of Al doping concentration on the opto-electronic chattels of SnS thin films readied by NSP, Opt. Quantum Electron., 2019, 51, 1–16 CrossRef CAS.
- X. Guo, Y. Wang, A. Elbourne, A. Mazumder, C. K. Nguyen, V. Krishnamurthi, J. Yu, P. C. Sherrell, T. Daeneke and S. Walia, Doped 2D SnS materials derived from liquid metal solution for tunable optoelectronic devices, Nanoscale, 2022, 14, 6802–6810 Search PubMed.
- J. Guan, Z. Zhang, M. Dou, J. Ji, Y. Song, J. Liu, Z. Li and F. Wang, Thermoelectric properties of Bi-doped SnS: First-principle study, J. Phys. Chem. Solids, 2020, 137, 109182 CrossRef CAS.
- K. S. Kumar, C. Manoharan, S. Dhanapandian and A. G. Manohari, Effect of Sb dopant on the structural, optical and electrical properties of SnS thin films by spray pyrolysis technique, Spectrochim. Acta, Part A, 2013, 115, 840–844 CrossRef PubMed.
- V. K. Arepalli, S. J. Kim and J. Kim, Influence of substrate temperature on the growth properties of Ag-doped SnS films deposited by sputtering method, J. Phys. Chem. Solids, 2021, 155, 110099 CrossRef CAS.
- S. Bano, M. Khan, H. Albalawi, G. U. Islam, M. Siddique, T. Ahmad, H. Alkhaldi, R. Ahson and S. Hussain, Effect of Cd doping on the structural, optical, and photovoltaic properties of SnS films, J. Mater. Res. Technol., 2022, 19, 1982–1992 CrossRef.
- M. A. Dar, D. Govindarajan, K. M. Batoo and C. Siva, Supercapacitor and magnetic properties of Fe doped SnS nanoparticles synthesized through solvothermal method, J. Energy Storage, 2022, 52, 105034 CrossRef.
- A. Javed, Q.-U. Ain and M. Bashir, Controlled growth, structure and optical properties of Fe-doped cubic π-SnS thin films, J. Alloys Compd., 2018, 759, 14–21 CrossRef CAS.
- M. Reghima, A. Akkari, C. Guasch, M. Castagné and N. Kamoun-Turki, Synthesis and characterization of Fe-doped SnS thin films by chemical bath deposition technique for solar cells applications, J. Renewable Sustainable Energy, 2013, 5, 063109 CrossRef.
- F. Niknia, F. Jamali-Sheini and R. Yousefi, Photocurrent properties of undoped and Pb-doped SnS nanostructures grown using electrodeposition method, J. Electron. Mater., 2015, 44, 4734–4739 CrossRef CAS.
- S. Sebastian, I. Kulandaisamy, S. Valanarasu, I. Yahia, H.-S. Kim and D. Vikraman, Microstructural and electrical properties evaluation of lead doped tin sulfide thin films, J. Sol-Gel Sci. Technol., 2020, 93, 52–61 CrossRef CAS.
- U. Chalapathi, Y. Jayasree and S.-H. Park, Lead-doped cubic tin sulfide thin films for solar cell applications, Mater. Sci. Semicond. Process., 2022, 150, 106958 CrossRef CAS.
- N. Mishra, N. Zakay, N. Maman, Y. Golan and G. Makov, Epitaxial growth of π-SnS on PbS substrates, Cryst. Growth Des., 2023, 23, 2301–2305 CrossRef CAS.
- B. B. Haidet, E. Hughes and K. Mukherjee, Epitaxial integration and defect structure of layered SnSe films on PbSe/III–V substrates, Cryst. Growth Des., 2022, 22, 3824–3833 CrossRef CAS.
- E. T. Hughes, B. B. Haidet, B. Bonef, W. Cai and K. Mukherjee, Pipe-diffusion-enriched dislocations and interfaces in SnSe/PbSe heterostructures, Phys. Rev. Mater., 2021, 5, 073402 CrossRef CAS.
- P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. D. Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio Jr, A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçükbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H.-V. Nguyen, A. Otero-de-la-Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu and S. Baroni, QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
- K. F. Garrity, J. W. Bennett, K. M. Rabe and D. Vanderbilt, Pseudopotentials for high-throughput DFT calculations, Comput. Mater. Sci., 2014, 81, 446–452 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
- D. R. Hamann, Optimized norm-conserving Vanderbilt pseudopotentials, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 85117 CrossRef.
- C. Freysoldt, B. Grabowski, T. Hickel, J. Neugebauer, G. Kresse, A. Janotti and C. G. Van de Walle, First-principles calculations for point defects in solids, Rev. Mod. Phys., 2014, 86, 253 CrossRef.
- J. Heyd, G. E. Scuseria and M. Ernzerhof, Hybrid functionals based on a screened Coulomb potential, J. Chem. Phys., 2003, 118, 8207–8215 CrossRef CAS.
- J. Heyd, G. E. Scuseria and M. Ernzerhof, Erratum: Hybrid functionals based on a screened Coulomb potential (J. Chem. Phys. 2003, 118, 8207–8215), J. Chem. Phys., 2006, 124, 219906 CrossRef.
- C. G. Van De Walle and J. Neugebauer, First-principles calculations for defects and impurities: Applications to III-nitrides, J. Appl. Phys., 2004, 95, 3851–3879 CrossRef CAS.
- E. Segev, U. Argaman and G. Makov, Materials Modelling of π-Phase Monochalcogenides, Isr. J. Chem., 2020, 60, 876–887 CrossRef CAS.
- E. Segev, R. E. Abutbul, U. Argaman, Y. Golan and G. Makov, Surface energies and nanocrystal stability in the orthorhombic and π-phases of tin and germanium monochalcogenides, CrystEngComm, 2018, 20, 4237–4248 RSC.
- P. D. Antunez, J. J. Buckley and R. L. Brutchey, Tin and germanium monochalcogenide IV–VI semiconductor nanocrystals for use in solar cells, Nanoscale, 2011, 3, 2399–2411 RSC.
- B. D. Malone, A. Gali and E. Kaxiras, First principles study of point defects in SnS, Phys. Chem. Chem. Phys., 2014, 16, 26176–26183 RSC.
- E. Segev, U. Argaman, R. E. Abutbul, Y. Golan and G. Makov, A new cubic prototype structure in the IV-VI monochalcogenide system: a DFT study, CrystEngComm, 2017, 19, 1751–1761 RSC.
- R. E. Abutbul, E. Segev, U. Argaman, G. Makov and Y. Golan, π-Phase tin and germanium monochalcogenide semiconductors: an emerging materials system, Adv. Mater., 2018, 30, 1–8 Search PubMed.
- R. Miranti, R. D. Septianto, T. Kikitsu, D. Hashizume, N. Matsushita, Y. Iwasa and S. Z. Bisri, π-SnS colloidal nanocrystals with size-dependent band gaps, J. Phys. Chem. C, 2022, 126, 5323–5332 CrossRef CAS.
- Z. Yang, M. Xu, X. Cheng, H. Tong and X. Miao, Manipulation of dangling bonds of interfacial states coupled in GeTe-rich GeTe/Sb2Te3 superlattices, Sci. Rep., 2017, 7, 17353 CrossRef PubMed.
Footnote |
| † These authors contributed equally to this work. |
|
| This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.