Open Access Article
Rohit
Dahule
a,
Chetan C.
Singh
a,
Kenta
Hongo
bc,
Ryo
Maezono
b and
Emila
Panda
*a
aMaterials Engineering Discipline, Indian Institute of Technology, Gandhinagar, Palaj-382355, Gujarat, India. E-mail: emila@iitgn.ac.in
bSchool of Information Science, Japan Advanced Institute of Science and Technology (JAIST), 1-1 Asahidai, Nomi, Ishikawa 923-1292, Japan
cResearch Centre for Advanced Computing Infrastructure, JAIST, 1-1 Asahidai, Nomi, Ishikawa 923-1292, Japan
First published on 28th February 2022
SnS is a promising photovoltaic absorber material because of its low cost and lower toxicity and is usually present in a heterostructure. Understanding the bulk and the surface electrical properties would help in the understanding of the transport behavior and hence would be extremely useful in fabricating high performance devices. In this regard, here a combinatorial approach of experiment and theory was used to understand the anomalies in the bulk and the surface electrical properties of SnS. Experimentally, single phase polycrystalline SnS films are fabricated by RF magnetron sputtering and characterized for their detailed microstructure, and optical, bulk and surface electrical properties. The observed anomalies in their bulk and surface electrical properties are then interpreted through first-principles density functional theory (DFT) calculations of the bulk and the surface electronic structures. DFT calculations on various native surface defects provided further insights into the experimentally observed semi-metallic behavior using scanning tunnelling spectroscopy.
To this end, here, a combinatorial approach of experiment and theory was used to understand the anomalies in the (defect-induced) surface and bulk electrical properties of SnS. Experimentally, SnS films are fabricated by varying the substrate temperature (Ts) from 303 to 623 K on the soda lime glass (SLG) substrate, for 60 min using radio frequency (RF) magnetron sputtering. RF magnetron sputtering is used here as this is a well-known technique for depositing high-quality thin films and is commercially adopted for fabricating thin film solar cells. Following fabrication, a detailed investigation on the microstructure, and optical, bulk and surface electrical properties of these films is carried out using a combination of experimental techniques. First-principles density functional theory (DFT) calculations are then performed on the SnS orthorhombic crystal structure to compute its electronic structure, as this is (experimentally) observed to be the most preferred crystallographic orientation. Moreover, DFT calculations are applied to the SnS (111) surface structures with the inclusion and exclusion of various native defects of vacancies, interstitials and antisites in order to investigate differences in electronic structures between the defective and defect-free surfaces by comparing with the bulk. In the end, a correlation between the theoretical data and the experimental observations is established.
Though extremely relevant, studies addressing both the theory and experiment are found to be scarce in the literature. Whereas, the experimental approach in the literature mostly focused on the fabrication and bulk optoelectronic properties of SnS thin films, theoretical studies mostly aimed at understanding the electronic structure, the band gap and the thermodynamics of the native defect states in its bulk.3,4,9 To the best of our knowledge, we have, for the first time, made detailed experimental and theoretical studies addressing both inclusive and exclusive of native defects on the SnS (111) surface.
Detailed structural, chemical, optical, and electrical characterization studies of these films were carried out using experimental techniques as elaborated in the following. The phase and the crystallite size of these deposited films were determined using Grazing Incidence X-ray diffractometry (GIXRD; model: D8 Discover, supplier: Bruker Corporation) at a glancing angle of 6° in the 2θ range of 20–80° (step size of 0.02°) using Cu Kα (λ = 0.15418 nm) radiation. The surface topography of these films was characterized using Atomic Force Microscopy (AFM; model: Nanoscope Multimode 8.0, supplier: Bruker Corporation). Surface chemical states were analyzed using X-ray photoelectron spectroscopy (XPS; model: PHI 5000 Versa Prob II, supplier: ULVAC-PHI, Inc.). The elemental chemical compositions of these deposited films were investigated using energy dispersive X-ray spectroscopy along with FESEM (EDS, model: AZtec, supplier: Oxford Instrument). The optical properties (i.e., transmittance and the optical band gap) of these SnS films were investigated using a UV-vis-NIR spectrophotometer (model: Cary 5000, supplier: Agilent Technologies) in the wavelength range of 300–2500 nm. A Hall effect measurement system (HEMS; model: 8404 AC/DC HMS, Lakeshore) in the van der Pauw configuration was used for measuring the bulk electrical resistivity, carrier concentration and carrier mobility of these films.
Additionally, the local (nanoscale) surface electrical properties of these SnS films were measured using scanning tunnelling microscopy/spectroscopy (STM/STS; the module of a scanning probe microscope (SPM)) under ambient conditions. STM images were acquired using a Pt/Ir tip (PT-10; supplier: Bruker Corporation) of 0.25 mm diameter and 8 mm length, following which local STS I–V characteristics were recorded at 21 different locations from each of these captured 1 μm × 1 μm STM images by applying a bias voltage of ±0.8 V. To maintain the electrical continuity while performing STM/STS measurement, silver paint was used for grounding the bottom metal plate and the top SnS thin film.
Our DFT simulations were performed using the Vienna Ab initio Simulation Package (VASP)13,14 and were used to first compute the bulk and the surface electronic structures of defect-free SnS; the Perdew–Burke–Ernzerhof (PBE)15 functional assuming the generalized gradient approximation was adopted as our exchange–correlation functional; the core–valence electron interaction was treated here using the projector augmented wave (PAW)16 method, where 4d, 5s, and 5p electrons of Sn and 3s and 3p electrons of S were treated as valence electron configurations. In addition, defective surface electronic structures were computed in a similar way to the pristine case, thereby investigating the influence of various thermodynamically stable native defect states on the surface electronic structure of SnS. Computational details of DFT settings are as follows: converged 3 × 7 × 7, 7 × 9 × 9 and 7 × 9 × 1 Monkhorst–Pack K-point grids17 were used for the computations of the bulk, (111) bulk, and (111) surface of SnS, respectively; a cut off energy of 450 eV was used for the plane wave expansion. In the case of surfaces, a dipole correction along the z-axis was employed to remove the artificial electric field in the vacuum region.18 The geometry optimizations were performed with thresholds for the energy (1.0 × 10−8 eV per cell) and force (0.01 eV Å−1) convergences.
In order to compare and understand the relative stabilities of various surface defects, their formation energy (Ef[X]) was calculated by using the following relation:19
![]() | (1) |
To calculate the chemical potentials of Sn and S under Sn- and/or S-rich conditions, the following boundary criteria were considered:
| μSn + μS ≤ μSnS,bulk, | (2) |
| μSn ≤ μSnS,bulk, | (3) |
| μS ≤ μSnS,bulk, | (4) |
The formation energy of these native defects at each possible defect sites was calculated to obtain the thermodynamically preferred site. For example, to obtain the preferred VSn surface site, various possible surface configurations were used by creating Sn vacancies at each possible location of the top SnS (111) surface (i.e., Sn1, Sn2, Sn3 and Sn4; see Fig. 1). Once the thermodynamically preferred defect site for a particular native defect is ascertained, further calculations concerning to the surface electronic structures such as the band structure and density of states were carried out only for the preferred defect sites. This procedure was followed for all the native defects considered in this study; the details can be found in Section S2 of the SI (ESI†). N.B. in order to avoid the defect-defect interaction, a bigger SnS (111) surface supercell with 32 atoms per slab (i.e., Sn16S16) was constructed for calculating the surface electronic structure and the density of states.20 These preferred sites were found out using the formation energy calculation as for the single native defect.
The surface chemical states of SnS films were investigated using XPS to understand the effect of Ts. Fig. 2(c) and 2(d) represent Sn 3d and S 2p core-level XPS spectra, respectively. The binding energy positions of Sn 3d5/2 and Sn 3d3/2 were found to be at 485.79 ± 0.02 eV and 494.27 ± 0.01 eV, respectively, with a doublet separation of 8.4 ± 0.01 eV, suggesting Sn in the +2 state.24 The peaks located at 161.45 ± 0.01 eV and 162.58 ± 0.02 eV with a separation of 1.1 eV are attributed to S 2p3/2 and S 2p1/2, confirming the presence of S in the −2 form only.25 All the SnS films were found to be Sn-rich, with the concentration of Sn (and hence Sn-to-S-ratio) increasing with increasing Ts (see Table 1 shown and discussed later). Note that the Sn-to-S-ratio was found to increase from 1.08 to 1.23 (as measured with XPS) and 1.19 to 1.26 (as measured with EDS) with increasing Ts.
| XPS(EDS)-measured composition | Electronic properties | |||||
|---|---|---|---|---|---|---|
| Sn (at%) | S (at%) | Sn–S ratio | ρ (Ω cm) | p h (× 1015 cm−3) | μ h (cm2 V s−1) | |
| T s = 303 K | 52.0(54.3) | 48.0(45.7) | 1.1(1.2) | 3420 | 6.13 | 0.29 |
| T s = 573 K | 54.9(55.6) | 45.1(44.4) | 1.2(1.3) | 592 | 3.17 | 3.32 |
| T s = 603 K | 55.1(55.9) | 44.9(44.1) | 1.2(1.3) | 632 | 3.4 | 2.9 |
Fig. 3(a) plots the optical transmittance of all the deposited SnS films (in the wavelength range of 300–2500 nm) as a function of Ts. All the films displayed sharp fundamental absorption edges with absorption edge shifting towards a longer wavelength with increasing Ts. Moreover, interference fringes caused by interference between these films and the respective SLG substrates were seen in all the samples, indicating a nearly uniform film with a smooth surface.26 Their indirect optical band gaps were determined from the intercept of the linear portion of
vs. (hν) on the x-axis using the Tauc equation:
| (αhν) = C(hν − Eg)n, | (5) |
![]() | ||
| Fig. 3 (a) Optical transmission spectra in the range of 400–2500 nm. (b) Optical band gap of SnS thin films deposited on soda-lime glass substrates at Ts of 303 K, 573 K and 623 K, respectively. | ||
Table 1 shows the variation in the electrical properties (i.e., electrical resistivity (ρ), carrier concentration (ph), and carrier mobility (μh)) of these SnS films with respect to Ts. All these films were found to be of p-type, with holes being the majority charge carriers. ph was found to decrease from 6.13 × 1015 cm−3 to 3.17 × 1015 cm−3 with increasing Ts from 300 to 573 K, beyond which it increased to 3.40 × 1015 cm−3 at 623 K. Similarly, μh was found to increase from 0.29 cm2 V−1 s−1 to 3.32 cm2 V−1 s−1 with increasing Ts from 300 to 573 K as a result of the reduction in grain boundary area because of an increase in crystallite size in these films, beyond which it decreased to 2.90 cm2 V−1 s−1 at 623 K, which can be attributed to an overall increase in the grain boundary area due to the slight decrease in the crystallite size (increasing the grain boundary scattering). Subsequently, ρ decreased from 3.42 × 103 Ω cm at Ts = 303 K to 0.632 × 103 Ω cm at Ts = 623 K. It has been stated in numerous SnS thin film studies that the carrier mobility of films increases with increasing crystallite size.31–33 Issei Suzuki et al. studied the temperature dependent carrier mobility of SnS films and found grain boundary scattering dominating the carrier transport in these films.34 Binqiang Zhou also investigated the transport properties of SnS, in which grain boundary scattering was found to strongly influence these transport properties at or below room temperature, whereas acoustic phonons35 were found to be the dominant scattering of holes at higher temperature (i.e., 400 K).
The varying electrical properties of these films can be attributed to the chemical composition (Sn/S ratio) that in turn induces Sn vacancies in these films. Note that Sn vacancies within these films induce the p-type characteristic.4 The formation of an increased Sn-rich (or S-deficient) film with increasing Ts can be attributed to the high volatility of sulfur. First-principles calculations on point defects in SnS4 predicted the spontaneous formation of Sn2+ vacancies in Sn-rich SnS due to their lower formation energies. In the calculations, these Sn vacancies form at the shallow acceptor levels, thereby giving rise to p-type conductivity in these films. The observed decrease in hole concentration in the SnS film with increasing Ts could be associated with a reduction in Sn vacancies because of the increased Sn atomic ratio (or Sn-to-S-atomic ratio). Formation of Sn vacancies in Sn-rich SnS implied the presence of these excess Sn atoms (apart from those in Sn lattice sites in SnS) either in interstitials (i.e., Sn interstitials (Sni)) or in S lattice positions (i.e., Sn antisites; SnS). Moreover, the overall S-deficient SnS films and the comparatively lower formation energies of sulfur vacancies (Vs) and SnS over Sni suggest a higher possibility of the formation of Vs and SnS in these SnS films. Nevertheless, the presence of Sni in these SnS films cannot be ignored entirely. As shown in the first-principles calculations by Brad. D. Malone et al.,4 the electronic defect states caused by the presence of Vs, SnS and/or Sni in Sn-rich SnS are formed at the donor levels.
Burton et al. determined the defect energy and defect concentrations of Sn and S vacancies in tin monosulphide. According to their calculations, the concentration of S vacancies is 107 times lower than that of Sn vacancies with the defect energy of Sn and S vacancies being 0.78 and 2.17 eV. This low defect energy of Sn vacancies leads to the formation of higher concentration of Sn vacancies at the shallow acceptor level, resulting in p-type conductivity in SnS, irrespective of Sn rich or S rich conditions.36 Vidal et al. projected Sn rich conditions resulting in the accumulation of Sn into interstitials, antisites and/or grain boundaries. Additionally, they also predicted that these excess Sn atoms (with the +2 oxidation state) undergo a relaxation phenomenon and get converted to Sn donor states with an oxidation state of +4. The experimental identification of these point defects in SnS based materials was done by photoluminescence measurements in the literature.3 Arepalli et al. observed red and orange emission peaks in the photoluminescence spectra of SnS particles, which they assigned to tin/sulfur vacancies, tin interstitials and/or other crystallographic defects.37 Ghosh et al. also observed emission peaks at about red and near the infrared region for SnS films deposited on glass. They have also ascribed these peaks to tin vacancies, interstitials and stacking faults which are generally formed during the growth process.38 The presence of dislocations also affects the electronic properties of SnS. The dislocation density (δ) can be evaluated using the Williamson and Smallman formula, δ = 1/D2. The dislocation densities of the SnS films deposited at Ts of 303 K, 573 K, and 623 K in this study corresponded to 2.1 × 10−3, 0.9 × 10−3 and 1.0 × 10−3 nm−2, respectively. These values are comparatively smaller, indicating that deposited layers contained densely packed grains with minimum line defects. P. A. Nwofe et al. also calculated a δ value of 2.2 × 10−2 nm−2 for SnS thin films deposited by a thermal evaporation method at 623 K on the SLG substrate.39
SnS thin films were also characterized in the scanning tunnelling spectroscopy mode by recording the tunnelling current with respect to the sample voltage from 21 different points and their corresponding tunnelling conductance (dI/dV), which represent the local density of states of the sample (see Fig. 4(i–vi)). The tunnelling current (I) is related to the charge carrier density and transmission coefficient via
![]() | (6) |
![]() | ||
| Fig. 4 (i–iii) Local STS I–V curves and (iv–vi) dI/dV curves of SnS thin films deposited on soda lime glass substrates at Ts of 303 K, 573 K, and 623 K, respectively. | ||
Note that all the films showed a non-zero conductance region around the Fermi level, indicating their semi-metallic nature. To quantify, the average tunnelling current was determined by averaging the local tunnelling current values at 0.5 V. This average tunnelling current was found to increase with increasing Ts from 8 to 17 nA. This increasing local tunnelling current in these semi-metallic films with increasing Ts may be associated with the change in the (local) surface electronic structure of these enhanced Sn-rich SnS thin films due to the formation of a higher number of point defects (i.e., VSn, VS, SnS and/or Sni; see Section 3.2).
Bulk SnS was found to be an indirect band gap semiconductor with a band gap of 0.9235 eV (Fig. 5(a)); the valence band was found to be mainly composed of the S 3p orbital along with a minor contribution from Sn 5s; the conduction band was found to primarily consist of Sn 5p and a minor contribution of S 3p orbitals (see Fig. 5(b)). The electronic structure and the PDOS of the bulk SnS slab with the (111) surface orientation showed similar features and orbital compositions to those of the bulk SnS, but with a reduced band gap of 0.75 eV (see Fig. 6(a and b)).
For surface calculation, all the atoms of the top two slabs of surface supercells, which were obtained by terminating the (111) surface of the bulk SnS, were relaxed by keeping their lattice parameters fixed. These optimized surface structures (after relaxation) and the position of defects are presented in Fig. 7(j). The geometry optimization of these atoms caused surface reconstruction for all these configurations; the atomic displacements are reported in Tables S2–S11 in the ESI.†
The optimized surface structure of the SnS (111) surface (see Fig. 7(a)) showed an outward displacement of the surface atoms (towards the vacuum), except the Sn4 atom, which was found to have penetrated inside the slab to minimize the energy of the system.46 Moreover, the SnS (111) surface was found to have almost the same orbital contributions to its valence band (VB) and conduction band (CB) as that of the bulk SnS (see Fig. 8(a)). The band structure revealed the surface to be a direct band gap semiconductor with a band gap of 0.2344 eV (Fig. 9(a)), which is contrary to that obtained for its bulk (Subsection 2.1). The surface associated band line appeared just below and above the bandgap ends in the band structure. This is consistent with the reported study for the low index surface in SnS.47 More precisely, it was found that the lowest VB (i.e., in the energy range of −2.5 to −2 eV) mainly consisted of S 3p and Sn 5p occupied states, whereas the highest VB (i.e., in the energy range of −2 to 0 eV) consisted of S 3p, Sn 5s, and Sn 5p occupied states. The CB (i.e., in the energy range of 0 to 2.5 eV) of the SnS (111) surface were mainly composed of Sn 5p and S 3p unoccupied states. The SnS (111) surface was found to have almost the same orbital contributions to both the valence band maximum (VBM) and conduction band minimum (CBM) as that of its bulk. In contrast, the DOS peaks of the SnS (111) surface were found to show sharp peaks (unlike bulk), which could be attributed to localized electronic states indicating a mixed ionic–covalent type of bonding.48
Table 3 lists the formation energies of all the above-mentioned native defect states at their preferred sites (see Section 2) evaluated from eqn (1) under both Sn-rich and S-rich conditions. Although all these native defects were found to be favorable, sulfur vacancies (VS) and sulfur antisites (SSn) were found to be the most favorable under Sn-rich and S-rich conditions respectively.
| Defect | Formation energy (eV) | |
|---|---|---|
| Sn-rich | S-rich | |
| VSn | 1.4123 | 0.5054 |
| VS | 0.5456 | 1.4525 |
| Sni | 0.8847 | 1.7916 |
| Si | 1.4273 | 0.5204 |
| SnS | 0.9520 | 2.7659 |
| SSn | 2.0551 | 0.2412 |
| VSn + 2 × VS | 7.9627 | 8.8696 |
| VSn + 2 × Sni | 12.9152 | 13.8221 |
| VSn + SnS | 7.5144 | 8.4214 |
In contrast to the PDOS of the pristine SnS (111) surface, DOS peaks corresponding to S 3p, Sn 5s, and Sn 5p unoccupied states above the Fermi level were observed in PDOS of the defective SnS (111) surface with VSn (see Fig. 8(b)). The band structure showed its band gap to be 0.48 eV and the above-mentioned unoccupied states were found to cross the Fermi level in the k-point path of H–Y–Γ as shown in Fig. 9(b). Moreover, the band line that was found to cross the Fermi line was attributed to the surface dangling bond because of Sn vacancies (Fig. 7(b)).49 Additionally, the states formed above the Fermi level (i.e., acceptor states) could cause p-type conductivity in the material system.
In the case of the SnS (111) surface with VS, the atoms (i.e., Sn2, Sn3, and Sn4 atoms) adjacent to the vacancy defect were found to move away due to the repulsion of these cations (Fig. 7(c)).49 The contribution of DOS obtained from this defective surface with Vs was found to be similar to that of the pristine surface, but with a missing sharp Sn 5p peak at the CBE and a direct band gap of 0.213 eV (see Fig. 8(c) and 9(c)). An additional band line was obtained at the VBE of the band structure in Fig. 9(c) because of S vacancies. This band line can act as an electron trap. A similar observation has also been reported in the literature, which carried out first principles DFT calculations on bulk SnS.3,4 Moreover, the defective surface with Vs was found to have the lowest formation energy under Sn-rich conditions (see Table 3), which is comparable to the study conducted by Vidal et al. on the bulk SnS system.3
In the case of interstitials, the adatom (i.e., Sni) in the reconstructed SnS surface with Sni showed a large displacement of 1.39 Å towards the vacuum, which could be attributed to the repulsion of Sni from the neighboring Sn atoms (see Fig. 7(d)). Moreover, the change in the surface structure due to the reconstruction resulted in an increase in the band gap to 0.381 eV as compared to that of the pure SnS (111) surface. Additionally, both the highest occupied and lowest unoccupied DOS states were found to be similar to that of the pure SnS surface, with the exception of the Sn 5p states, which was observed in the energy range of −0.1 to −0.3 eV. Furthermore, the band structure showed a peak near to the Fermi line at a high symmetry point X because of the Sn 5p state. Similarly, the adatom (i.e., Si) in the SnS (111) surface structure with Si as shown in Fig. 7(e) was found to displace upward by 0.947 Å and was attributed to the strong repulsion from the neighboring S atoms. Moreover, the DOS structure was found to be similar to that of the pure SnS (111) surface, except for the sharp Sn 5p peak at the CBE. Furthermore, the defective surface with Si showed a direct band gap of 0.201 eV at the C point (see Fig. 9(e)).
In the case of the SnS defect, the band gap disappeared because of the overlapping of VB and CB (see Fig. 8(f) and 9(f)). As a result, this surface configuration would exhibit a metallic character. Its DOS below and above the Fermi level in the energy range of −0.2 to 3 eV was found to be composed of mainly Sn 5p along with minor contribution from S 3p states. Note that its lowest VB (i.e., in the energy range of −3 to −0.2 eV) was found to be similar to that of the pristine SnS (111) surface. As evident from Fig. 8 (g) and 9(g), the orbital contribution of the SnS (111) surface with SSn at the VBE and CBE as well as its band structure was found to be similar to that of the pristine surface. Moreover, a direct band gap of 0.33 eV is found at the C point (see Fig. 9(g)).
As shown in Subsection 3.1, the experimentally fabricated SnS thin films were found to be Sn-rich and p-type (because of the presence of Sn vacancies as VSn has been found to form acceptor states). Additionally, as observed using the DFT calculation in this study, defects like VS, Sni, and SnS could easily form owing to their low formation energy under Sn-rich conditions (see Table 3). These findings motivated us to further calculate multiple surface native defects for conditions favoring Sn-rich along with Sn vacancies: (i) VSn + 2 × VS, (ii) VSn + 2 × Sni and (iii) VSn + SnS.
(i) In the case of VSn + 2 × VS defects, the unoccupied states of Sn 5p (major) and S 3p (minor) were found to form above the Fermi level (see Fig. 8(h)). These unoccupied states were found to have crossed the Fermi level in the k-point path of X–C–Γ (see Fig. 9(h)), thereby reducing the band gap (0.21 eV) compared to VSn (0.48 eV).
(ii) The VSn + 2 × Sni defect configuration had an indirect band gap of 0.2216 eV (see Fig. 9(i)), whereas the single native Sni defect had a direct band gap of 0.381 eV, though their relative contributions were found to be similar (see Fig. 8(i)).
(iii) In the case of VSn + SnS defects, the VB and CB were found to be overlapping (see Fig. 8(j) and 9(j)), indicating their metallic characters. The orbital contributions to the VB in the surface with VSn + SnS were found to be similar to those in the pristine surface; the DOS above the Fermi level (i.e., CB) was found to be similar to that of the defective surface with SnS. Additionally, the smaller Sn 5p DOS peaks in the energy range of 0 to 0.2 eV were found for the SnS surface with VSn + SnS (unlike with the PDOS of the SnS defect configuration).
Footnote |
| † Electronic supplementary information (ESI) available: Schematics of bulk SnS and SnS (111) surface supercells. Selection of suitable surface defect sites of vacancies, interstitials and antisites. The ionic displacements of the atoms in the top slab for the SnS (111) surface supercell and the SnS (111) surface supercell with various defects. See DOI: 10.1039/d1tc04738h |
| This journal is © The Royal Society of Chemistry 2022 |