You are on page 1of 13

Solar Energy 174 (2018) 373–385

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Phase optimization study of orthorhombic structured SnS nanorods from T


CTAB assisted polyol synthesis for higher efficiency thin film solar cells
Benjamin Hudson Baby, D. Bharathi Mohan*
Department of Physics, School of Physical, Chemical and Applied Sciences, Pondicherry University, R.V. Nagar, Kalapet, Puducherry 605 014, India

A R T I C LE I N FO A B S T R A C T

Keywords: Orthorhombic SnS nanorods were prepared from CTAB assisted polyol synthesis method at reaction temperature
Tin sulphide of 150 °C. In this method, the source (Sn, S) concentration ratio, CTAB concentration and reaction time were
Polyol synthesis optimized to obtain single phase SnS as evident from Raman and XPS analyses. Crystal structure from XRD along
Structural properties with SAED pattern confirmed the formation of orthorhombic SnS structure. Morphological analysis from HR-
Optical
TEM revealed the formation of SnS nanorods with improved homogeneity in size and shape by increasing the
Electrical properties
reaction time and CTAB concentration. The observed shift in the absorption band edge and absorption maxima
by varying the CTAB concentration and reaction time is very well correlated with the morphological and che-
mical structural variations. Tauc plot shows the successful tuning of band gap value to 1.5 eV which is the
optimum value of thin film solar cells, by varying the reaction time of sample with 0.025 M of CTAB. Electrical
properties were studied from the Hall measurement which showed p-type conductivity for single phase SnS
while n-type conductivity observed for mixed phases.

1. Introduction atom with three short Sn-S bonds of ca 2.7 A° and three long Sn-S bonds
of almost 3.4 A° (Lewis et al., 2014; Pejjai et al., 2017). Each unit cell
In the field of thin film solar cells (TFSCs), Tin (II) Sulfide (SnS) is a consists of two SnS layers and one of the long distance sulfur resides on
promising alternative p-type material as an absorber layer because of its the neighboring SnS layer. This weak Sn-S interaction binds two SnS
excellent physical, electrical and optical properties compared to the layers together to form a double layer structure which is perpendicular
existing absorber layer materials such as CZTS, CIGS, GaAs and CdTe to the b axis with tin and sulfur atoms covalently bonded within the
(Wang et al., 2014; Hirai et al., 2013; Green et al., 2015). Although layers and weak van der Waals bonds between the layers (Yang et al.,
these materials exhibit very high efficiency, the toxicity, scarcity, and 2015). This van der Waals interaction between the double layers in SnS
cost associated with In, Ga, As, Te and Cd present in CIGS, GaAs and gives a chemically inert surface with few surface states and these defect
CdTe and the complexities associated with manufacturing quaternary tolerant surface reduce carrier recombination loss due to the defects at
compounds such as CIGS and CZTS limit their sustainability (Banu p-n junction and at grain boundaries (Sinsermsuksakul et al., 2011;
et al., 2016; Park et al., 2015; Baby and Mohan, 2017a). While SnS is Albers et al., 1961; Avellaneda et al., 2008). Also SnS has majority
less toxic, its constituent elements are abundant in nature and it is a carrier Hall mobility 90 cm2/Vs or higher and tunable majority carrier
binary compound which involves simple growth chemistry compared to density in the range of 1015–1018 cm−3(Sucharitakul et al., 2016; Banai
quaternary compounds. It has direct a band gap value of 1.38 eV close et al., 2016). It exhibit p-type conductivity due to the availability of tin
to the optimum value of 1.5 eV (Cheng and Conibeer, 2011). It has high vacancy (VSn) which act as a shallow acceptor (Vidal et al., 2012). It
optical absorption coefficient of α > 105 cm−1 for photons with en- undergoes a phase transition from low symmetric α phase to high
ergy higher than the band gap value and this minimizes the thickness of symmetric β phase at 600 °C which is a second order phase transition
an absorber layer in TFSCs and hence relaxing the need of long minority produced by the continuous shift of Sn and S atom along (1 0 0) axis
carrier diffusion length (Devika et al., 2007, 2010; Klochko et al., 2016; (Reddy et al., 2015b; Reddy, 2013).
Banai et al., 2016). In SnS, each Sn atom is coordinated with six sulfur Although the theoretical efficiency for SnS based single junction

Abbreviations: TFSCs, thin film solar cells; CZTS, copper zinc tin sulphide; CIGS, copper indium gallium sulphide; GaAs, gallium arsenide; CdTe, cadmium telluride;
NPs, nanoparticles; Rt, reaction time; CTAB, cetyltrimethylammonium bromide; CBD, chemical bath deposition; EG, ethylene glycol; UDEDM, uniform deformation
energy density model
*
Corresponding author.
E-mail address: d.bharathimohan@gmail.com (D. Bharathi Mohan).

https://doi.org/10.1016/j.solener.2018.09.019
Received 28 June 2018; Received in revised form 5 September 2018; Accepted 10 September 2018
Available online 20 September 2018
0038-092X/ © 2018 Elsevier Ltd. All rights reserved.
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Fig. 1. Schematic diagram of CTAB assisted polyol synthesis method for the preparation of SnS NPs.

solar cell is 24% predicted by Prince-Loferski, the conversion efficiency UV–Vis-NIR spectrophotometer. 3. To study the electrical property of
of SnS based solar cells increased from 1.3% to 3.88% by using CdS as a the semiconductor SnS NPs with the effect due to the presence of SnS2
buffer layer and 4.36% using Zn(O,S) as a buffer layer by following and Sn2S3 phases from Hall effect measurement.
thermal evaporation and atomic layer deposition techniques respec-
tively (Loferski, 1956; Reddy et al., 2006; Steinmann et al., 2014;
Sinsermsuksakul et al., 2014). This far discrepancy between the theo- 2. Experimental procedure
retical predictions and experimental results may be due to the presence
of intrinsic defects and impurity phases and also due to poor device SnS nanoparticles were synthesized using ethylene glycol (EG)
fabrication (Malone et al., 2014; Sinsermsuksakul et al., 2014; Burton (Qualigens Fine Chemicals) as the reaction medium as well as the re-
et al., 2013; Baby and Mohan, 2017b). ducing agent for salts and CTAB as the cationic surfactant. For the re-
The different shape of SnS nanostructures were prepared by sol- action, high pure (reagent grade) SnCl2 and Na2S salts used as tin and
vothermal, hydrothermal, PVP assisted, CTAB assisted and surfactant sulfur sources, obtained from Merck (Merck Specialties Pvt. Ltd,
free micro emulsion methods and the phase formation was studied Mumbai.) and Fischer Scientific (Thermo Fischer Scientific India Pvt.
through XRD (Ren et al., 2012; Tarkas et al., 2017; Sohila et al., 2011a; Ltd, Mumbai) respectively. Initially, 0.1 M of SnCl2, 0.6 M of Na2S
Biswas et al., 2007; Panda et al., 2006; Yue et al., 2009; Peng et al., prepared in 10 ml of EG separately. Then Na2S solution added drop by
2007; Yang et al., 2015; Lu et al., 2013; Reddy et al., 2013; Chaki et al., drop into SnCl2 solution placed in silicon oil bath at 150 °C and after the
2014). It is always difficult to prove single phase formation by probing addition, the solution color changes to dark brown indicating the for-
through XRD because it cannot give information about any particular mation of Sn-S phase. Then solution was collected and quenched in a
phase if it was present with less than 5% (Baby and Mohan, 2017a, water bath after 10 min. This colloidal solution was washed several
2017b). Many other researchers studied the single phase SnS formation times using centrifuge at 5000 rpm and then Sn-S NPs were collected by
by characterizing with Raman spectrometer exciting at wavelengths removing supernatant. Then, a thick (few microns) film of Sn-S was
514 and 532 nm (Sohila et al., 2011b; Ali et al., 2017; Bhorde et al., prepared by a simple drop casting technique. Following the above
2018). However, Raman active modes of Sn2S3 phase cannot be excited procedure a set of experiment was carried out by varying the ratio
with laser wavelengths 514 and 532 nm Baby and Mohan (2017a). Liu between SnCl2 and Na2S from 1:6 to 1:15 M to optimize the source
et al. (2003) confirmed the formation of single phase SnS by following concentration ratio between Sn:S in the formation of single phase of
CTAB and oxalic acid assisted DI water based reaction for 10 h. Gedi SnS.
et al. (2016) prepared orthorhombic structured single phase SnS by The role of CTAB as a cationic surfactant in the structural and
following CBD method with tartaric acid was used as a complexing morphological properties of SnS is studied by varying the concentration
agent. Chalapathi et al. (2016) prepared cubic structured SnS by fol- of CTAB from 0.01 M to 0.5 M. Initially, 0.1 M of SnCl2, 1.2 M of Na2S
lowing CBD method with EDTA was used as the complexing agent. Baby and 0.01 M of CTAB is prepared in 10 ml of EG separately. Then by
and Mohan (2017a) reported the phase optimization study of SnS NPs following the similar experimental setup as explained above, CTAB and
prepared by PVP assisted polyol synthesis method. According to the Na2S solutions were added drop by drop into SnCl2 solution and as a
best of our knowledge, it is very clear that a detailed analysis to confirm result the solution color changes to dark brown. Then reaction is
the phase purity of orthorhombic structured SnS NPs using surfactant stopped after 10 min, washing several times with methanol and then a
(CTAB) without the formation of secondary phases such as SnS2 and thick film is prepared using drop casting technique. Then the experi-
Sn2S3, with a band gap ∼1.5 eV along with the support of Hall mea- ment is repeated by increasing the concentration of CTAB as 0.025,
surement study were not clearly investigated so far which are indeed 0.05, 0.1, 0.15, 0.2, 0.3 and 0.5 M. The effect of reaction time (Rt) is
important aspects towards enhancing the efficiency of SnS. In this work, also studied by varying the Rt as 10, 20, 40 and 60 min. The complete
CTAB assisted polyol method is followed for the synthesis of single experimental procedure is drawn as a schematic diagram as shown in
phase SnS NPs. It involves ethylene glycol (EG) as the reducing agent Fig. 1.
and cetyltrimethylammoniumbromide (CTAB) as the cationic surfac- The chemical structure of as deposited drop casted films were stu-
tant, controls the formation growth and agglomeration of NPs. died by using confocal Raman spectrometer (ReinshawinVia Raman
The main objectives of this work are following, 1. To synthesis Microscope, UK) with an excitation wavelength of 514 and 785 nm
single phase SnS with an energy band gap near to 1.5 eV using a simple using Ar+ and semiconductor lasers respectively. Raman analysis was
wet chemistry route by tailoring certain parameters such as (a) Sn:S carried out in the scan range of 50–350 cm−1 with acquisition time and
source concentration ratio, (b) CTAB concentration and (c) Reaction power as 30 s and 1.25mW for 514 nm and 150 s and 0.0025 mW for
time. 2. To investigate the single phase formation and the possibility of 785 nm excitation wavelengths respectively. The crystal structure of as
formation of secondary phases by XRD, Raman, XPS, EDAX, and deposited drop casted films were studied by using X-ray diffraction
(XRD) (Panalytical-X’Pert PRO). The surface morphology and electron

374
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

diffraction patterns of as prepared samples were studied by using high along with SnS. By increasing to 1:12 M leads to the complete absence
resolution transmission electron microscope (HR-TEM) (JEOL/JEM of the peak at 314.4 cm−1 confirming the complete suppression of SnS2
2100). The elemental composition of the samples were studied by EDAX phase. Also by increasing Sn:S ratio from 1:6 to 1:12 M leads to a blue
analysis by using VEGA 3, TESCAN. The oxidation states of Sn and S shift in the Raman peak from 64.7 to 66 cm−1 corresponds to Sn2S3
and the chemical bonding of SnS were analysed using X-ray photo- phase whereas the Raman peak corresponds to SnS is observed at
electron spectroscopy (XPS) (PHI 5000 Versa Probe II, FEI Inc) in the 70 cm−1. Further increase in the Sn:S ratio to 1:15 M leads to a red shift
binding energy range of 0–1000 eV and the binding energies in spec- in the Raman peak towards 64.7 cm−1 corresponds to Sn2S3 phase.
trometer were calibrated using C 1s line. The absorbance spectra of the Hence for Sn:S source concentration ratio of 1:12 M, chemical structural
colloidal solutions prepared after dispersing the NPs in the methanol analysis shows a trend to form single phase of SnS, which is not com-
medium was studied using UV–Vis-NIR spectrophotometer (Shimadzu pletely achieved in the absence of any surfactant. Also it can be clearly
UV-3600 Plus) from the wavelength range of 300–1200 nm. The seen that Raman modes of SnS2 is highly active with an excitation
emission spectra of the prepared colloidal solutions were recorded with wavelength of 514 nm whereas Raman modes of SnS and Sn2S3 phases
an excitation wavelength of 520 nm using spectrofluorometer (FLUO- are highly active by using an excitation wavelength of 785 nm (Baby
ROLOG-FL3-11), equipped with Xenon lamp-450 W. The electrical re- and Mohan, 2017a).
sistivity, mobility, carrier concentration, etc. were studied using Agilent The effect of surfactant in the single phase formation of SnS is
device analyser-B1500A Hall measurement system. studied by varying the concentration of CTAB from 0.01 to 0.3 M by
keeping Sn:S molarity ratio 1:12 M for a Rt of 10 min using 514 and
3. Results and discussions 785 nm as excitation wavelength as shown in Fig. 3(a) and (b) re-
spectively. By using 514 nm as an excitation wavelength, formation of
3.1. Chemical structural analysis SnS2 (∼314 cm−1) as a secondary phase is confirmed for samples
prepared with higher concentration of CTAB (> 0.15 M) while single
Raman spectroscopy is an efficient, sensitive and nondestructive phase of SnS is obtained for lower concentration (up to 0.1 M). By using
technique to study the phase formation of different tin sulphide phases 785 nm as an excitation wavelength, single phase of SnS is obtained for
such as SnS, SnS2 and Sn2S3 by their distinct Raman shift. The pre- samples prepared with 0.025 M of CTAB while increasing the CTAB
dominant Raman active modes of SnS are Ag (40, 95, 192 and concentration leads to the formation of Sn2S3 (∼63 cm−1) as a sec-
218 cm−1), B2g (70, 85 and 290 cm−1) and B3g (49 and 164 cm−1) ondary phase along with SnS. Hence chemical structural analysis con-
(Chandrasekhar et al., 1977). Similarly the Raman active modes of firmed the formation of single phase of SnS for samples prepared with
Sn2S3 is observed at 63, 183, 154 and 308 cm−1 and for SnS2 205 and 0.025 M of CTAB whereas by increasing the CTAB concentration up to
315 cm−1 (Chandrasekhar and Mead, 1979; Smith et al., 1977). 0.1 M leads to the formation of Sn2S3 as an additional phase along with
Fig. 2(a) and (b) shows the Raman spectra of as deposited samples SnS and further increase results in the formation of both SnS2 and Sn2S3
prepared by varying the source concentration ratio of Sn:S from 1:6 M as an additional phase along with SnS.
to 1:15 M for a Rt of 10 min without any surfactants, using an excitation The influence of Rt on the single phase formation of SnS is studied
wavelength of 514 and 785 nm respectively. For samples prepared with by varying the Rt as 10, 20, 40 and 60 min as shown in Fig. 4. By
Sn:S molarity ratio up to 1:9 M, Raman analysis confirmed the forma- increasing the Rt to 20 min (Fig. 4(a)), the secondary phases are de-
tion of SnS2 and Sn2S3 (314.4 and 64.7 cm−1) as secondary phases creasing for sample prepared with 0.025 M of CTAB and all the peaks

Fig. 2. Raman spectra of as deposited sample prepared by varying Sn:S ratio from 1:6 to 1:15 M for the reaction time of 10 min without CTAB using excitation
wavelength of (a) 514 nm and (b) 785 nm.

375
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Fig. 3. Raman spectra of as deposited samples prepared by varying the concentration of CTAB from 0.025 to 0.3 M for a recation time of 10 min using excitation
wavelength of (a) 514 and (b) 785 nm.

observed matching with SnS phase. Further increase in CTAB con- (Chandrasekhar et al., 1977; Baby and Mohan, 2017a). The peak at 95,
centration leads to a red shift in these two peaks showing the formation 190 and 220 cm−1 corresponds to the Ag (c(aa)c) mode which arise due
of Sn2S3 phase. Hence Raman analysis confirmed the formation of to the incident and scattered direction of radiation parallel to c - axis
single phase of SnS only for sample with 0.025 M of CTAB. However and the polarization of incident and scattered photons along a axis
further increase in reaction time towards 40 and 60 min leads to the (Chandrasekhar et al., 1977; Baby and Mohan, 2017a). The high intense
formation of single phase of SnS irrespective of variation in the CTAB peak at 220 cm−1 corresponds to the vibration of Sn atom against the S
concentration (Fig. 4(b) and (c)). The FWHM value is decreasing sys- atoms (Chandrasekhar et al., 1977; Baby and Mohan, 2017a). The less
tematically by increasing the Rt from 10 to 60 min indicating the im- intense peaks at 95 cm−1 and 190 cm−1 corresponds to the ‘breathing’
proved crystallinity of the sample. Chemical structural analysis con- and ‘waving’ modes of the layer (Chandrasekhar et al., 1977; Baby and
firmed that improved phase stability of SnS is achieved by increasing Mohan, 2017a). The less intense peak at 285 cm−1 corresponds to the
the Rt. Also by increasing the Rt from 10 min to 60 min, the Raman B2g (b(ac)b) mode which arise due to the vibration of Sn atom against S
peak at 164 cm−1 corresponds to the B3g (c(ab)c) mode become pro- atom (Chandrasekhar et al., 1977; Baby and Mohan, 2017a). The ob-
minent indicating the predominant vibration of Sn atom against S atom served Raman peak at 74 cm−1 for samples with Rt of 40 min is also

Fig. 4. Raman spectra using an excitation wavelength of 785 nm for as deposited samples prepared by varying the concentration of CTAB from 0.025 to 0.2 M for a
recation time of (a) 20 min (b) 40 min and (c) 60 min.

376
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

And the calculated values are enclosed in Table 1.


Williamson-Hall relation given by (Williamson and Hall, 1953)
β cosθ K 4ε sinθ
= +
λ D λ (4)
For an orthorhombic crystal, young’s modulus related to their
elastic compliances Sij as
1
= S11 l14 + (2S12 + S66 ) l12 l 22 + S22 l 24 + (2S23 + S44 ) l 22 l32 + S33 l34
E
+ (2S13 + S55) l12 l32

where l1, l2 and l3 are direction cosines (Wen et al., 2017). S11, S12, S66,
S22, S23, S44, S33, S13 and S55 are the elastic compliances of SnS which can be
calculated from the elastic constants (Lubarda and Chen, 2008; Koc et al.,
2015; Kafashan, 2018). The calculated values for the elastic compliances
of SnS are S11 = 13.549 × 10−12, S12 = −0.129 × 10−12, S13 =
−6.5665 × 10−12, S33 = 15.8052 × 10−12, S44 = 27.5710 × 10−12 and
Fig. 5. XRD pattern of 350 °C annealed samples prepared with 0.025 and 0.2 M
S66 = 25.9538 × 10−12 (TPa−1).
of CTAB with reation time of 10 and 40 min.
The energy density of a crystal can be calculated using uniform
deformation energy density model (UDEDM). In UDEDM, the aniso-
corresponds to B2g mode appeared due to the compressive mode of the tropic nature of the crystallite is considered with the assumption that
layer along c axis (Chandrasekhar et al., 1977; Baby and Mohan, the deformation energy is uniform in all crystallographic direction. For
2017a). an elastic systems that follows hooks law, the energy density u can be
calculated as u = ε2Yhkl/2 (Aly et al., 2017; Zak et al., 2011). Then equ.
3.2. Crystal structural analysis 4 can be rewritten as
1/2
The crystal structure of the drop casted films were studied using β cosθ K 4sinθ ⎛ 2u ⎞
= + ⎜ ⎟

XRD technique after annealed the samples under vacuum (10−3 mbar) λ D λ ⎝ Yhkl ⎠ (6)
at 350 °C for 90 min as shown in Fig. 5. As the same peak orientation Then the anisotropic energy density (u) can be calculated from the
was observed for all samples prepared by varying the CTAB con- slop of the graph plotted between 4sinθ/(Yhkl/2)1/2 vs βcosθ and the
centration from 0 to 0.2 M and Rt as 10, 20 and 40 min, only a couple of crystallite size from the y intercept as shown in Fig. 6 and the calculated
important results were included in Fig. 5 whereas the microstructural values are enclosed in Table 2. Stress is calculated from the Hook’s law,
properties were discussed for other samples and enclosed in Table 1. All keeping the linear relation between stress and strain given by, σ = Yε,
the diffraction peaks were compared with the standard JCPDS card No. where σ is the stress of the crystal, ε is anisotropic micro strain which
00-039-0354 and indexed as (1 1 0), (1 2 0), (1 0 1), (1 1 1), (1 3 1), will depend upon the crystallographic directions and Y is the young’s
(2 1 0), (0 0 2), (2 1 1), (1 1 2), (1 2 2), (2 3 1), (0 4 2), (2 5 1) and (1 5 2) modulus (Y = 105.3 GPa) (Aly et al., 2017; Mote et al., 2012; Koc et al.,
confirming the formation of single phase polycrystalline SnS NPs with 2015).
orthorhombic crystal structure. The preferred orientation along (1 1 1) The observed micro strain could be possibly due to the presence of
plane predicts the formation of Sn rich SnS NPs (Baby and Mohan, the above mentioned intrinsic defects in the polycrystalline sample.
2017a). Crystallite size calculated from W-H method, increases with increasing
The lattice constants of orthorhombic structure SnS phase can be Rt due to increase in crystallization of SnS structure. The dislocation
calculated using the equation density is related with the crystal growth defects and it is defined as the
1 h2 k2 l2 length of dislocation line per unit volume which can be calculated from
= 2 + 2 + 2 the crystallite size using Williamson and Smallman’s formula
d2 a b c (1)

and the values are enclosed in Table 1 which is compared with the δ = 1/D2 (7)
std JCPDS card no 00-039-0354. Increasing the Rt as well as the CTAB where D is the crystallite Size and the values are enclosed in Table 2
concentration results an expansion in the a and b direction and com- which is well agreement with the reported values. Increasing the Rt as
pression along c direction. The crystallite size of these samples were well as the CTAB concentration enhance the crystallisation of the
studied from XRD analysis using Debye-Scherer formula sample which in turn decrease the Dislocation density as evident from
Kλ Fig. 5 (Chaki et al., 2014). The number of unit cells in SnS nanocrystals
D= can be calculated using the following equation (Aly et al., 2017),
βcosθ (2)
n = πD3 /(6V) (8)
where D is crystallite size, K is shape factor (∼0.9), λ is wavelength of
Cukα radiation and β is full width at half maximum (in radians) and the where V is the unit volume and the calculated values are enclosed in
calculated values are enclosed in Table 1. The crystallite size is slightly Table 2. Increase in the Rt as well as CTAB concentration results an
decreased in CTAB based reaction as compared to EG based reaction, increase in the crystal size number per unit volume which is due to the
whereas by increasing the Rt from 10 to 60 min leads to a slight in- slight increase in the crystallite size by increasing the Rt and CTAB
crease in the crystallite size. Micro strain of the samples, which arise concentration.
due to dislocations, grain boundary triple junction, contact or sinter
stresses, stacking faults, coherency stresses, etc. is calculated from XRD 3.3. XPS analysis
using the following equation (Mote et al., 2012; Baby and Mohan,
2017a; Gedi et al., 2016), Fig. 7 shows a wide range XPS spectra of SnS NPs synthesized with
0.025 M of CTAB, attributed to the presence of C, O, Sn and S. The
β cosθ
Strain(ε) = observed XPS peaks were calibrated using C1s at 284.5 eV. The XPS
4 (3)
spectra shows characteristic peaks of Sn 4d, S 2p, S 2, C 1 s, Sn 3d, O 1 s,

377
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Table 1
2θ, FWHM, d-spacing, crystallite size, micro strain, lattice parameters and unit cell volume values obtained from the XRD analysis of annealed SnS samples.
Sample code 2theta (°) FWHM (β) d (A°) D (nm) D average (nm) ε * 10−3 ε average * 10−3 a (A°) b (A°) c (A°) V = abc (A°)3

00-039-0354 4.3291 11.1923 3.9838 193.03

Sn:S – 1:12 26.03 0.52836 3.4204 15 15 2.3209 2.2877 4.2966 11.2992 3.9861 193.52
31.59 0.56242 2.8291 15 2.3595
38.98 0.55101 2.3088 15 2.2647
51.09 0.51803 1.7863 16 2.0378
48.74 0.61831 1.8668 14 2.4556

CTAB 0.025 M – 10 min 25.95 0.59348 3.4204 14 14 2.5048 2.39 4.2932 11.3180 3.9903 193.89
31.55 0.59302 2.8299 14 2.4403
38.93 0.57628 2.3105 15 2.3597
51.09 0.62682 1.7863 15 2.3445

CTAB 0.025 M – 20 min 25.98 0.5806 3.4269 17 16 2.3568 2.3626 4.3079 11.3111 3.9837 194.11
31.49 0.5535 2.8317 15 2.3225
38.81 0.54712 2.311 16 2.2676
48.84 0.6038 1.8632 14 2.3971
51.09 0.632 1.7863 16 2.4715

CTAB 0.025 M – 40 min 25.97 0.45139 3.4282 19 17 1.9177 2.4397 4.3109 11.3318 3.9733 194.10
31.49 0.54121 2.8291 15 2.2538
38.94 0.4261 2.311 20 3.5033
51.08 0.5384 1.7866 19 2.1268
48.68 0.6034 1.8689 14 2.3970

CTAB 0.2 M – 10 min 30.54 0.6281 2.9248 12 13 2.7675 2.4164 4.2999 11.2869 3.99 193.64
31.52 0.5745 2.8369 12 2.411
38.91 0.6286 2.3128 13 2.6211
51.02 0.6375 1.7886 13 2.5309
64.1 0.6492 1.4516 14 2.3992

CTAB 0.2 M – 20 min 26.01 0.61472 3.423 13 14 2.6114 2.5728 4.2970 11.3244 3.9886 194.09
31.60 0.65858 2.8291 13 2.7629
38.95 0.5822 2.3105 14 2.3932
48.78 0.63499 1.8654 14 2.5215

CTAB 0.2 M – 40 min 31.43 0.63859 2.8299 13 15 2.6807 2.6906 4.3065 11.3520 3.9781 194.45
38.91 0.62867 2.3128 14 2.6211
51.02 0.63757 1.7886 15 2.5309
64.1 0.64922 1.4516 15 2.3992

Sn 3p and Sn 3 s. The observed peaks at 285 and 531 eV correspond to oxidation state and S-Sn4+ bonding respectively and Sn2S3 has both
C1s and O1s respectively due to the surface contamination during drop Sn2+ and Sn4+ oxidation states (Reddy et al., 2015a; Baby and Mohan,
casting (Reddy et al., 2015a). Fig. 8(a) and (b) shows the high resolu- 2017a). Hence the observed Sn3d5/2 peaks at 486 and 161 eV attributed
tion XPS spectra of Sn 3d and S 2p states of CTAB 0.025 M – 10 min. The to the presence of Sn2+ oxidation state and S-Sn2+ bond formation
observed separation of 8.5 eV between the Sn3d5/2 and Sn3d3/2 is well which confirms the formation of single phase of SnS without any sec-
matching with standard values (Reddy et al., 2015a; Baby and Mohan, ondary phases which is in good agreement with Raman and XRD re-
2017a). Among different phases of tin sulphides, SnS has Sn3d5/2 peak sults. Fig. 8(c) and (d) shows Sn 3d and S 2p states high resolution XPS
at 485.9 eV and S2p3/2 peak at 161 eV attributed to the presence of spectra corresponds to sample CTAB 0.05 M – 10 min. The observed
Sn2+ oxidation state and S-Sn2+ bonding respectively, SnS2 has Sn3d5/2 doublet at 485.4 and 486.4 eV confirms the presence of Sn2+ and Sn4+
peak at 487 eV and S2p3/2 peak at 162 eV indicate the presence of Sn4+ oxidation states respectively. High resolution XPS spectra of S2p3/2

Fig. 6. The plot between (β cosθ) vs (4 sinθ/(Ehkl/2)1/2) for samples prepared with CTAB 0.025 M for reaction time of (a) 10 min and (b) 40 min.

378
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Table 2
Calculated crystallite size (D), micro strain (ε), Stress (σ), and energy density (u) of SnS nanoparticles obtained using Williamson-hall method.
Sample code D (nm) ε UDEDM δ * 10−3 (nm−2) n

−3
D (nm) ε σ (Mpa) U (KJm )

CTAB 0.025 M – 10 min 14 2.39 14 1.7640 186 163 5.40 6.79


CTAB 0.025 M – 40 min 17 2.4395 18 2.2612 238 269 3.02 16.25

Fig. 7. A wide range XPS spectra for sample CTAB 0.025 M – 10 min. Fig. 9. HR-TEM image shows surface morphology of Sn-S NPs prepared in EG
based reaction with the molarity ratio of 1:12 (Sn:S).

Fig. 8. High resolution XPS spectra of (a) and (c) for Sn3d and (b) and (d) for S2p oxidation states for sample CTAB 0.025 M – 10 min and CTAB 0.15 M – 10 min
respectively.

379
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Fig. 10. HR-TEM images show surface morphology of SnS NPs prepared using 0.025 M of CTAB for Rt of (a) 10 min and (c) 40 min. (b) SAED pattern of poly-
crystalline SnS NPs of sample CTAB 0.025–10 min. (d) High resolution image of CTAB 0.025–40 min shows the inter planar spacing of SnS NPs.

consist of 2 peaks at 161 and 161.7 eV which attributed to the bond comparing with JCPDF card No. 00-039-0354. While for sample with
formation of S-Sn2+ and S-Sn4+ respectively. Hence XPS analysis CTAB 0.2 M – 10 min (Fig. 11(b)) confirmed the formation of Sn2S3
confirms the formation of Sn2S3 and SnS2 as secondary phases along [JCPDF card No. 01-073-0031] as a secondary phase along with SnS as
with SnS for sample CTAB 0.05 M – 10 min as evident from Raman evident from Raman analysis.
analysis. Surface morphology analysis confirmed the improved homogeneity
in size and shape of as prepared SnS NPs by using CTAB as cationic
surfactant, compared to EG base reaction and it is further enhanced by
3.4. Surface morphology increasing the CTAB concentration from 0.025 M to 0.2 M. Also by in-
creasing the CTAB concentration from 0.025 to 0.2 M, length and width
Fig. 9 shows the surface morphology of SnS NPs prepared in EG of the NRs is decreasing from 250 to 180 nm and 30 to 10 nm respec-
without CTAB for Rt of 10 min. It shows highly inhomogeneous size tively with an increase in aspect ratio from 5.5 to 9.2, indicating more
formation of SnS nanorods with length and width ranging from 60 to controlled growth process by increasing the CTAB concentration due to
100 nm and 10–15 nm respectively with an average aspect ratio ∼ 6.6. the improved stabilization of NPs by increasing the surfactant con-
Whereas in EG-CTAB based reactions, with 0.025 M of CTAB for Rt of centration which enables the effective surfactant adsorption on NPs
10 min (Fig. 10(a)), the formation of SnS nanorods (with length surface prevents their unlimited growth (Niu and Li, 2014). Whereas by
250 nm, width ∼ 35 nm and aspect ratio ∼ 5.5) and spherical shape of increasing the Rt from 10 to 40 min leads to the formation of highly
particles (with size of 150 nm in diameter) is confirmed. By increasing homogenous NRs with increase in length and width from 180 to 220 nm
Rt to 40 min, the reaction leads to the formation of highly homogenous and 10 to 25 nm respectively with a slight increase in aspect ratio from
SnS nanorods with length 220–250 nm, width 25–35 nm and average 9.2 to 9.6. Hence surface morphology analysis confirmed that CTAB
aspect ratio 8 (Fig. 10(c)). By increasing CTAB concentration to 0.2 M concentration and Rt has a significant effect in the formation of highly
and with Rt 10 min leads to the formation of SnS nanorods with length homogenous size and shape SnS NPs.
180–200 nm, width 10–12 nm and aspect ratio ∼ 9.2 (Fig. 11(a)).
Further increase in Rt to 40 min leads to the formation of highly
homogenous SnS nanorods with length 220–240 nm, width 25–30 nm 3.5. Formation mechanism of NRs
and aspect ratio ∼ 9.6 (Fig. 11(c)). The high resolution image
(Fig. 10(d)) shows that the observed nanorods have 0.28 and 0.35 nm In SnS, Sn is having +2 oxidation state with sterically active and
as the interplanar spacing corresponds to hkl planes (1 1 1) and (1 2 0) chemically inactive inert lone pair in 5 s orbital (Burton and Walsh,
as the preferential growth direction respectively confirming the or- 2012; Tripathi and Mitra, 2013). This inert lone pair of Sn(II) forces the
thorhombic crystal structure of SnS as evident from XRD analysis (Pol crystal to grow in the [1 0 1] direction with successive layers between
et al., 2008). It is already discussed that SnS exhibit layered structure two SnS molecule along one direction which leads to the formation of
with each Sn atom is coordinated six S atoms results an highly distorted NRs (Tripathi and Mitra, 2013). The surface energies (σ) low indexed
octahedral geometry which leads to three short and three long SneS lattice planes of orthorhombic SnS crystal structure increases in the
bonds with ∼0.27 and ∼0.34 nm respectively (Lewis et al., 2014; order of σ(100) < σ(001) < σ(010) < σ(111) (Reddy, 2013). As a result,
Suryawanshi et al., 2014). The observed two lattice fringes in HRTEM in order to obtain lowest free energy, crystal growth is maximum along
image may arise due to this distorted octahedral geometry (Reddy et al., low energy facets and minimum along high energy facets. XRD analysis
2015b). The selected area electron diffraction pattern (SAED) images confirmed the preferential growth direction of SnS along [1 1 1] plane
confirmed the formation of single phase orthorhombic polycrystalline with [1 0 0] and [0 0 1] as the base planes due to the low surface energy
SnS NPs for sample prepared with EG-CTAB based reactions with CTAB of these planes (Reddy, 2013). Hence, atoms are added along the [1 0 0]
0.025–10 min (Fig. 10(b)) and CTAB 0.2 M – 40 min (Fig. 11(d)) after and [0 0 1] direction with a tampering end to minimize lattice face

Fig. 11. HR-TEM images shows the surface morphology of Sn-S NPs prepared using 0.2 M of CTAB for Rt of (a) 10 and (d) 40 min. SAED pattern of polycrystalline Sn-
S NPs of samples (b) CTAB 0.2 M – 10 min and (d) CTAB 0.2 M – 40 min. The symbol * indicates for the sample CTAB 0.2 M – 10 min corresponds to Sn2S3 phase.

380
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

3.7. Optical properties

The optical absorbance spectra of all samples dispersed in methanol


were studied in the wavelength range of 300–1200 nm using UV–Vis-
NIR absorption spectrophotometer. Fig. 14 shows the absorption
spectra of the samples prepared in (a) EG and (b) CTAB based reaction
for Rt of 10 min. Fig. 15(a), (b) and (c) shows the absorption spectra of
Fig. 12. Growth mechanism of SnS nanorods. samples prepared in CTAB based reaction for Rt of 20, 40 and 60 min
respectively. For samples prepared from EG based reactions without the
surface energy. Initial nucleation of SnS monoatomic layers happens addition of CTAB, by increasing Sn:S source concentration ratio from
along the [1 0 1] axis direction. Due to the layered structure of SnS 1:6 to 1:12 M, the absorption maxima is red shifted from 415 to 450 nm
perpendicular to [0 1 0] direction with van der Waals interaction be- and the absorption band edge is shifted as 680, 662 and 747 nm which
tween the layers induces slowest growth rate along [0 1 0] direction, may be due to the complete absence of SnS2 (Eg = 2.2 eV) phase and
leads to the formation of rod like structure as shown in Fig. 12 (Tan also due to the suppression of Sn2S3 (Eg = 1.78 eV) phase as evident
et al., 2016). from Raman analysis (Baby and Mohan, 2017a). The band gap of SnS
(Eg = 1.38 eV) is very low as compared to the values of other secondary
phases (Burton et al., 2013). Further increase of Sn:S ratio to 1:15 M
3.6. Elemental composition leads to a blue shift in the absorption maxima and band edge as 430 and
714 nm respectively, may be due to the formation of larger quantity of
The quantitative elemental analysis of samples CTAB 0.025 M – Sn2S3 NPs along with SnS phase.
40 min and CTAB 0.15 M – 10 min were carried out using EDAX. EDAX Whereas the incorporation of CTAB (0.025 M) as a cationic surfac-
mapping analysis as shown in Fig. 13(a) shows Sn/S ratio 1 for sample tant in the reaction leads to a red shift in the absorption maxima to
CTAB 0.025 M – 40 min indicating the formation of single phase of SnS 455 nm as shown in Fig. 14(b) due to the controlled growth of SnS NPs
NPs as evident from Raman analysis (Burton et al., 2013). EDAX as evident from surface morphological analysis. Further increase in Rt
mapping analysis of sample CTAB0.15 M – 10 min as shown in as 20, 40 and 60 min as shown in Fig. 15(a), (b) and (c), leads to a
Fig. 13(c) shows the formation of sulphur rich SnS phase with Sn/S systematic red shift in the absorption maxima as 465, 473 and 549 nm
ratio 0.89 predicting the formation of Sn2S3 as a secondary phase along respectively showing the improved homogeneity in size and shape of
with SnS for sample CTAB 0.15 M – 10 min as observed from Raman the prepared NPs. For samples prepared with Rt of 10 and 20 min, by
analysis (Burton et al., 2013). Fig. 13(b) and (d) shows EDAX point varying CTAB concentration from 0.025 M to 0.2 M, the absorption
analysis of samples CTAB 0.025 M – 40 min and CTAB 0.15 M – 10 min band edge is blue shifted from 800 to 740 nm and 850 to 810 nm re-
which is in good agreement with EDAX mapping analysis. spectively. The observed drastic blue shift in the absorption band edge
by increasing the CTAB concentration from 0.025 M can be correlated

Fig. 13. (a) EDAX mapping and (b) elemental analyses (both qualitative and quantitative) of SnS NPs prepared with 0.025 M of CTAB for a Rt of 40 min (CTAB
0.025 M – 40 min) reveals Sn/S ratio equal to 1. (c) EDAX mapping and (d) elemental analyses (both qualitative and quantitative) of Sn-S NPs prepared with 0.15 M
of CTAB for the Rt of 10 min (CTAB 0.15 M – 10 min) shows Sn/S ratio equal to 0.89. The insert table shows the calculated weight and atomic percentage of Sn and S.

381
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Fig. 14. Optical absorption spectra of Sn-S NPs prepared in (a) EG based reaction and (b) EG-CTAB (0.025–0.2 M) based reaction for the Rt of 10 min.

Fig. 15. Optical absorption spectra of Sn-S NPs prepared in EG-CTAB (0.025–0.2 M) based reaction for the Rt of (a) 20 min (b) 40 min and (c) 60 min. (d) Tauc plot
shows the energy band gap of SnS NPs prepared with 0.025 M of CTAB by varying the Rt from 10 to 60 min.

with the variation in particle size as well as due to the formation of with Rt of 40 and 60 min, the absorption band edge is slightly blue
Sn2S3 as an additional phase along with SnS as evident from chemical shifted from 895 to 870 nm and 1035 to 1010 nm respectively. Che-
structural analysis (Guo et al., 2015). Whereas for samples prepared mical structural analysis confirmed the formation of single phase of SnS

382
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Table 3 secondary phases such as SnS2 and Sn2S3. By increasing Rt from 10 to


Calculated direct band gap energy values. 60 min, band gap value is systematically decreasing which is attributed
Sample code Band gap (eV) to the variation in particle size as evident from morphological analysis.
Chemical structural analysis confirmed the formation of single phase of
10 min 20 min 40 min 60 min SnS for all samples prepared using Rt of 40 and 60 min and hence band
gap of single phase SnS NPs were successfully tuned to 1.5 eV by
Sn:S – 1:6 M 2.15 – – –
Sn:S – 1:9 M 2.30 – – – varying Rt and CTAB concentration.
Sn:S – 1:12 M 1.92 – – – Fig. 16 shows emission spectra in the range of 700–900 nm for the
Sn:S – 1:15 M 1.98 – – – sample CTAB0.025 M – 10 min with an excitation wavelength of
CTAB 0.025 M 1.69 1.59 1.45 1.34 520 nm. The observed sharp red emission peak at 782 nm may be at-
CTAB 0.05 M 2.05 1.72 1.58 1.51
tributed to the free carrier or excitonic band edge transition between
CTAB 0.1 M 2.02 1.84 1.75 1.49
CTAB 0.15 M 2.18 1.88 1.76 – the valence band and conduction band (Baby and Mohan 2017b; Devika
CTAB 0.2 M 2.2 1.82 1.72 – et al., 2010; Kafashan et al., 2016) which is in good agreement with the
CTAB 0.3 M 1.92 1.72 1.6 – direct energy band gap value obtained from the Tauc plot.

3.8. Electrical properties

The electrical properties of CTAB0.15 M – 10 min and CTAB


0.025 M – 40 min were studied using Hall effect experiment on a thick
film prepared with the thickness of 5 µm using drop casting technique.
The drop casted films were annealed under vacuum at temperature
200 °C for 60 min and the values of resistivity, carrier concentration,
mobility and polarity were calculated as shown in Table 4. Chaki et al.
(2014) prepared different shaped SnS NPs using wet chemical route and
obtained resistivity 659 × 103 Ω cm for NPs, 406 × 103 Ω cm for nano
whiskers and 690 × 102 Ω cm for nanoribbons with carrier concentra-
tion of 5.20 × 1011, 8.43 × 1011 and 4.99 × 1012 cm−3 respectively.
Guneri et al. (2010) prepared SnS NPs with needle shape like structure
by CBD method and obtained resistivity of 253 × 103 Ω cm. Chalapathi
et al. (2016) prepared cubic structured SnS NPs with resistivity and
carrier concentration in the range of 105 × 102 Ω cm and
7.93 × 1012 cm−3 respectively. Gedi et al. (2016) reported resistivity
value of 38 Ω cm and carrier concentration 3.4 × 1015 cm−3 for SnS
films prepared by CBD method. In this work, sample CTAB0.025 M –
Fig. 16. The emission spectra of sample CTAB0.025 M – 10 min shows the band 40 min showed the formation of single phase SnS NRs, exhibiting p type
edge emission of SnS in the range of 700–900 nm with an excitation wavelength conductivity with resistivity 272 × 102 Ω cm and carrier concentration
of 520 nm. 1.14 × 1013 cm−3 and hall mobility 20.31 cm2 V−1 S−1. The value is
found to be slightly higher as compared to Gedi et al. work could be
for all these samples irrespective of the CTAB concentration. And hence attributed to the formation of SnS nanorods and however comparable
the observed slight blue shift in the absorption band edge is only due to with the value of nanoribbons prepared by Chaki et al group. For the
the variation in particle size. sample CTAB 0.05 M – 10 min where Raman and XPS analysis con-
Direct band gap value of all the samples were calculated from x- firmed the presence of SnS2 and Sn2S3 as additional phases along with
intercept of Tauc plot and the values are enclosed in Table 3. Fig. 15(d) SnS, showing n-type conductivity with resistivity value of 13,860 Ω cm,
shows Tauc plot of samples prepared with 0.025 M of CTAB by varying carrier concentration 0.45 × 1013 cm−3 and electron mobility
Rt from 10 to 60 min. The observed high band gap value for samples 100.21 cm2 V−1 S−1.
prepared without CTAB is due to the formation of SnS2 and Sn2S3
phases along with SnS phase as evident from chemical structural ana- 4. Conclusion
lysis. By increasing the Sn:S ratio to 1:12 M, band gap value is red
shifted from 2.3 to 1.92 eV due to the complete suppression of SnS2 The optimization study towards the formation of single phase of SnS
phase which is having comparatively higher band gap value among NRs from CTAB assisted polyol synthesis method, by varying the Sn:S
different tin sulphide phases. Whereas by using CTAB as a surfactant, source concentration, CTAB concentration and reaction time has been
for Rt of 10 min, band gap value is red shifted to 1.69 eV due to the investigated for the first time. Raman and XPS analyses showed the
formation of single phase of SnS which is having comparatively low formation of single phase SnS with 0.025 M of CTAB and lead to form
band gap value among different tin sulphide phases. The observed blue SnS2 and Sn2S3 for the higher CTAB concentration. The single phase of
shift of 0.31 eV as compared to bulk band gap value of SnS could be SnS is observed irrespective of CTAB concentration with the reaction
attributed to the quantum confinement effect (Nair et al., 2005; Baby time of 40 min. XRD and SAED analyses confirmed the formation of
and Mohan, 2017a). Whereas further increase in CTAB concentration polycrystalline orthorhombic SnS for 0.025 M of CTAB. The structural
leads to an increase in band gap value due to the formation of properties such as crystallite size, strain, stress and energy density

Table 4
Electrical properties of Sn-S NPs.
Sample Code Resistivity × 102 (Ω cm) Polarity(p/n) Carrier concentration × 1013 (cm−3) Mobility (cm2 V−1 S−1) Phase present

CTAB 0.15 M – 10 min 138.6 n 0.45 100.21 SnS, Sn2S3, SnS2


CTAB 0.025 M – 40 min 272 p 1.14 20.31 SnS

383
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

values are calculated from Williamson-Hall analysis by using UDEDM. of P-type SnS thin films prepared by chemical bath deposition. Chalcogen. Lett. 7,
HR-TEM analysis confirmed the formation of SnS nanorods with an 685–694.
Guo, S., Fidler, A.F., He, K., Su, D., Chen, G., Lin, Q., Pietryga, J.M., Klimov, V.I., 2015.
improved homogeneity in size and shape with the increase of CTAB Shape controlled narrow-gap SnTe nanostructures: From nanocubes to nanorods and
concentration and Rt. EDAX mapping showed Sn/S ratio as 1 and 0.89 nanowires. J. Am. Chem. Soc. 137, 15074–15077.
for CTAB concentration of 0.025 and 0.15 M respectively, which is Hirai, Y., Kurokawa, Y., Yamada, A., 2013. Numerical study of Cu(In, Ga)Se2 solar cell
performance toward 23% conversion efficiency. Jpn. J. Appl. Phys. 53,
corroborating with Raman and XPS analyses. Band gap calculations 012301–12305.
showed the successful tuning by varying the reaction time for 0.025 M Kafashan, H., Sheini, F.J., Kahrizsangi, R.E., Yousefi, R., 2016. Nanostructured SnS1-xTex
of CTAB, which is close to the optimum value of 1.5 eV. Hall mea- thin films: effect of Te concentration and physical properties. J. Alloys Comp. 681,
595–605.
surement showed p-type conductivity with resistivity 27,200 Ω cm for Kafashan, H., 2018. Structural characterizations of pure SnS and In-doped SnS thin films
pure SnS whereas n-type conductivity is observed when the secondary using isotropic and anisotropic models. Mater. Res. Exp. 5, 046417.
phases present along with SnS phase. Klochko, N.P., Lukianova, O.V., Kopach, V.R., Tyukhov, I.I., Volkova, N.D., Khrypunov,
G.S., Lyubov, V.M., Kharchenko, M.M., Kirichenko, M.V., 2016. Development of a
new thin film composition for SnS solar cell. Sol. Energy 134, 156–164.
Acknowledgements Koc, H., Simsek, S., Palaz, S., Oltulu, O., Mamedov, A.M., Ozbay, E., 2015. Mechanical,
electronic, and optical properties of the A4B6 layered ferroelectrics: ab initio calcu-
The authors acknowledge SERB-DST, India (SR/FTP/PS-131/2012 lation. Phys. Stat. Solidi C 6, 651–658.
Lewis, D.J., Kevin, P., Osman, B., Muryn, C.A., Malik, M.A., O’Brien, P., 2014. Routes to
& EEQ/2016/000228) for the financial support. Authors thank ACMS- tin chalcogenide materials as thin films or nanoparticles: a potentially important class
IIT Kanpur for providing XPS facility. Authors are grateful to Dr. D. of semiconductor for sustainable solar energy conversion. Inorg. Chem. Front. 1,
Selvakumar, Scientist-F, DEBEL (DRDO), Bangalore for EDAX analysis. 577–598.
Liu, Y., Hou, D., Wang, G., 2003. Synthesis and characterization of SnS nanowires in
Authors thank STIC-CUSAT, Kochi for providing HR-TEM facility. cetyltrimethylammoniumbromide (CTAB) aqueous solution. Chem. Phys. Lett. 379,
Authors thank CeNSE –IISC, Bangalore for providing Hall Measurement 67–73.
facility. BHB sincerely thanks UGC-BSR for providing research fellow- Loferski, J.J., 1956. Theoretical considerations governing the choice of the optimum
semiconductor for photovoltaic solar energy conversion. J. Appl. Phys. 27, 777–784.
ship. Lu, J., Nan, C., Li, L., Peng, Q., Li, Y., 2013. Flexible SnS nanobelts: facile synthesis,
formation mechanism and application in Li-ion batteries. Nano Res. 6, 55–64.
References Lubarda, V.A., Chen, M.C., 2008. On the elastic moduli and compliances of transversely
isotropic and orthotropic materials. J. Mech. Mater. Struct. 3, 153–170.
Malone, B.D., Gali, A., Kaxiras, E., 2014. First principles study of point defect in SnS.
Albers, W., Haas, C., Vink, H.J., Wasscher, J.D., 1961. Investigations on SnS. J. Appl. Phys. Chem. Chem. Phys. 16, 26176–26183.
Phys. 32, 2220–2225. Mote, V.D., Purushotham, Y., Dole, B.N., 2012. Williamson-Hall analysis in estimation of
Ali, S., Wang, F., Zafar, S., Iqbal, T., 2017. Hydrothermal synthesis, characterization and lattice strain in nanometer-sized ZnO particles. J. Theor. Appl. Phys. https://doi.org/
Raman vibrations of chalcogenide SnS nanorods. IOP Conf. Ser. Mater. Sci. Eng. 275, 10.1186/2251-7235-6-6.
012007. Nair, S.S., Mathews, M., Anantharaman, M.R., 2005. Evidence for blue shift by weak
Aly, K.A., Khalil, N.M., Algamal, Y., Saleem, Q.M.A., 2017. Estimation of lattice strain of exciton confinement and tuning of bandgap in superparamagnetic nanocomposites.
zirconia nano-particles based on Williamson-Hall analysis. Mater. Chem. Phys. 193, Chem. Phys. Lett. 406, 398–403.
182–188. Niu, Z., Li, Y., 2014. Removal and utilization of capping agents in nanocatalysis. Chem.
Avellaneda, D., Nair, M.T.S., Nair, P.K., 2008. Polymorphic tin sulfide thin films of zinc Mater. 26, 72–83.
blende and orthorhombic structures by chemical deposition. J. Electrochem. Soc. Panda, S.K., Datta, A., Dev, A., Gorai, S., Chaudhari, S., 2006. Surfactant-assisted
155, D517. synthesis of SnS nanowires grown on tin foils. Cryst. Growth Des. 6, 2177–2181.
Baby, B.H., Mohan, D.B., 2017a. The formation of α-phase SnS nanorods by PVP assisted Park, H.H., Heasley, R., Sun, L., Steinmann, V., Jaramillo, R., Hartman, K., Chakraborty,
polyol synthesis: phase stability, micro structure, thermal stability and defects in- R., Sinsermsuksakul, P., Chua, D., Buonassisi, T., Gordon, R.G., 2015. Co-optimiza-
duced energy band transitions. Mater. Chem. Phys. 192, 317–329. tion of SnS absorber and Zn(O, S) buffer materials for improved solar cells. Prog.
Baby, B.H., Mohan, D.B., 2017b. The formation of α-phase SnS nanostructure from a Photovolt: Res. Appl. 23, 901–908.
hybrid, multi-layered S/Sn/S/Sn/S thin films: phase stability, surface morphology Pejjai, B., Reddy, V.R.M., Gedi, G., Park, C., 2017. Status review on earth-abundant and
and optical studies. Appl. Surf. Sci. 423, 1111–1123. environmentally green Sn-X (X = Se, S) nanoparticle synthesis by solution methods
Banai, R.E., Horn, M.W., Brownson, J.R.S., 2016. A review of tin (II) monosulfide and its for photovoltaic applications. Int. J. Hydrogen Energy 42, 1–42.
potential as a photovoltaic absorber. Sol. Energy Mater. Sol. Cells 150, 112–129. Peng, H., Jiang, L., Huang, J., Li, G., 2007. Synthesis of morphologically controlled tin
Banu, S., Ahn, S.J., Eo, Y.J., Gwak, J., Cho, A., 2016. Tin monosulfide (SnS) thin films sulfide nanostructures. J. Nanopart. Res. 9, 1163–1166.
grown by liquid-phase deposition. Sol. Energy 145, 33–41. Pol, V.G., Pol, S.V., Gedanken, A., 2008. Core-shell nanorods of SnS-C and SnSe-C:
Bhorde, A., Pawbake, A., Sharma, P., Nair, S., Funde, A., Bankar, P., More, M., Jadkar, S., synthesis and characterization. Langmuir 24, 5135–5139.
2018. Solvothermal synthesis of tin sulfide (SnS) nanorods and investigation of its Reddy, K.T.R., Reddy, N.K., Miles, R.W., 2006. Photovoltaic properties of SnS based solar
field emission properties. Appl. Phys. A 124, 133 (1–8). cells. Sol. Energy Mater. Sol. Cells 90, 31041–33046.
Biswas, S., Soumitra, K., Subhadra, C., 2007. Thioglycolic acid (TGA) assisted hydro- Reddy, N.K., Mudusu, D., Gunasekhar, K.R., 2013. Influence of seed layer orientation on
thermal synthesis of SnS nanorods and nanosheets. Appl. Surf. Sci. 253, 9259–9266. the growth and physical properties of SnS nanostructures. Appl. Phys. A 116,
Burton, L.A., Walsh, A., 2012. Phase stability of the earth abundant tin sulphides SnS, 1193–1197.
SnS2 and Sn2S3. J. Phys. Chem. 116, 24262–24267. Reddy, N.K., 2013. Growth-temperature dependent physical properties of SnS nanocrys-
Burton, L.A., Colombara, D., Abellon, R.D., Grozema, F.C., Peter, L.M., Savenije, T.J., talline thin films. ECS J. Solid State Sci. Technol. 2, P259–P263.
Dennler, G., Walsh, A., 2013. Synthesis, characterization and electronic structure of Reddy, V.R.M., Gedi, S., Park, C., Miles, R.W., 2015a. Development of sulphurized SnS
single-crystal SnS, Sn2S3, and SnS2. Chem. Mater. 25, 4908–4916. thin film solar cells. Curr. Appl. Phys. 15, 588–598.
Chaki, S., Chaudhary, M., Deshpande, M.P., 2014. Synthesis and characterization of Reddy, N.K., Devika, M., Gopal, E.S.R., 2015b. Review on tin (II) sulfide (SnS) material:
different morphological SnS nanomaterials. Adv. Nat. Sci. Nanosci. Nanotechnol. 5, synthesis, properties and applications. Crit. Rev. Solid State Mater. Sci. 40, 359–398.
045010. Ren, L., Jin, Z., Cai, S., Yang, J., Hong, Z., 2012. Green synthesis by diethylene glycol
Chalapathi, U., Poornaprakash, B., Park, S.H., 2016. Chemically deposited cubic SnS thin based solution process and characterization of SnS nanoparticles. Cryst. Res. Technol.
films for solar cell applications. Sol. Energy 139, 238–248. 47, 461–466.
Chandrasekhar, H.R., Humphreys, R.G., Zwick, U., Cardona, M., 1977. Infrared and Sinsermsuksakul, P., Heo, J., Noh, W., Hock, A.S., Gordon, R.G., 2011. Atomic layer
Raman spectra of the IV-VI compounds SnS and SnSe. Phys. Rev. B 15, 2177–2183. deposition of tin monosulfide thin films. Adv. Energy Mater. 1, 1116–1125.
Chandrasekhar, H.R., Mead, D.G., 1979. Long-wavelength phonons in mixed-valence Sinsermsuksakul, P., Sun, L., Lee, S.W., Park, H.H., Kim, S.B., Yang, C., Gordon, R.G.,
semiconductor SnIISnIVS3. Phys. Rev. B 19, 932–937. 2014. Overcoming efficiency limitations of SnS-based solar cells. Adv. Energy Mater.
Cheng, S., Conibeer, G., 2011. Physical properties of very thin SnS films deposited by 4, 1400496.
thermal evaporation. Thin Solid Films 520, 837–841. Smith, A.J., Meek, P.E., Liang, W.Y., 1977. Raman scattering studies of SnS2 and SnSe2. J.
Devika, M., Reddy, N.K.R., Ramesh, K., Ganeshan, R., Gunasekhar, K.R., Gopal, E.S.R., Phys. C: Solid State Phys. 10, 1321–1333.
Reddy, K.T.R., 2007. Thickness effect on the physical properties of evaporated SnS Sohila, S., Rajalakshmi, M., Muthamizhchelvan, C., Kalavathi, S., et al., 2011a. Synthesis
film. J. Electrochem. Soc. 154, H67–H73. and characterization of SnS nanosheets through simple chemical route. Mater. Lett.
Devika, M., Reddy, N.K, Prashantha, M., Ramesh, K., Reddy, S.V., Hahn, Y.B., 65, 1148–1150.
Gunasekhar, K.R., 2010. The physical properties of SnS films grown on lattice-mat- Sohila, S., Rajalakshmi, M., Ghosh, C., Arora, A.K., Muthamizhchelvan, C., 2011b. Optical
ched and amorphous substrates. Phys. Stat. Solidi A 207, 1864–1869. and Raman scattering studies on SnS nanoparticles. J. Alloys Compd. 509,
Gedi, S., Reddy, V.R.M., Kotte, T.R.R., Kim, S.H., Jeon, C.W., 2016. Chemically synthe- 5843–5847.
sized Ag-doped SnS films for PV applications. Ceram. Int. 42, 19027–19035. Steinmann, V., Jaramillo, R., Hartman, K., Chakraborty, R., Brandt, R.E., Poindexter, J.R.,
Green, M.A., Emery, K., Hishikawa, Y., Warta, W., Dunlop, E.D., 2015. Solar cell effi- Lee, Y.S., Sun, L., Polizzotti, A., Park, H.H., Gordon, R.G., Buonassisi, T., 2014. 3.88%
ciency tables (version 45). Prog. Photovolt. 23, 1–9. Efficient tin sulfide solar cells using congruent thermal evaporation. Adv. Mater. 26,
Guneri, E., Gode, F., Ulutas, C., Kirmizigul, F., Altindemir, G., Gumus, C., 2010. Properties 7488–7492.

384
B.H. Baby, D. Bharathi Mohan Solar Energy 174 (2018) 373–385

Sucharitakul, S., Kumjar, U.R., Sankar, R., Chou, F.C., Chen, Y.T., Wang, C., He, C., Gao, 100, 032104.
X.P.A., 2016. Screening limited switching performance of multilayer 2D semi- Wang, W., Winkler, M.T., Gunawan, O., Gokmen, T., Todorov, T.K., Zhu, Y., Mitzi, D.B.,
conductor FETs: the case for SnS. Nanoscale 8, 19050–19057. 2014. Device characteristics of CZTSSe thin-film solar cells with 12.6% efficiency.
Suryawanshi, S.R., Warule, S.S., Patil, S.S., Patil, K.R., More, M.A., 2014. Adv. Energy Mater. 4, 1301465 (1–5).
Vapor−liquid−solid growth of one-dimensional tin sulfide (sns) nanostructures with Wen, Y., Wang, L., Liu, H., Song, L., 2017. An initio study of the elastic and mechanical
promising field emission behaviour. ACS Appl. Mater. Interfaces 6, 2018–2025. properties of B19 TiAl. Crystals 7. https://doi.org/10.3390/cryst7020039.
Tan, Q., Wu, C.F., Sun, W., Li, J.F., 2016. Solvothermally synthesized SnS nanorods with Williamson, G.K., Hall, W.H., 1953. X-ray line broadening from filed aluminium and
high carrier mobility leading to thermoelectric enhancement. RSC Adv. 6, wolfram. Acta. Metall. 1, 22–31.
43985–43988. Yang, H., Kim, C.E., Giri, A., Soon, A., Jeong, U., 2015. Synthesis of surfactant-free SnS
Tarkas, H.S., Marathe, D.M., Mahajan, M.S., Muntaser, F., Patil, M.B., Tak, S.R., Sali, J.V., nanoplates in an aqueous solution. RSC Adv. 5, 94796–94801.
2017. Synthesis of tin monosulfide (SnS) nanoparticles using surfactant free micro- Yue, G.H., Wang, L.S., Wang, X., Chen, Y.Z., Peng, D.L., 2009. Characterization and op-
emulsion (SFME) with the single microemulsion scheme. Mater. Res. Express 4, tical properties of the single crystalline sns nanowire arrays. Nanoscale Res. Lett. 4,
025018. 359–363.
Tripathi, A.M., Mitra, S., 2013. Tin sulfide (sns) nanorods: structural, optical and lithium Zak, A.K., Majid, W.H.A., Abrishami, M.E., Yousefi, R., 2011. X-ray analysis of ZnO na-
storage property study. RSC Adv. 4, 10358–10366. noparticles by Williamson-Hall and size-strain plot methods. Solid State Sci. 13,
Vidal, J., Lany, S., d’Avezac, M., Zunger, A., Zakutayev, A., 2012. Band structure, optical 251–256.
properties and defect physics of the photovoltaic semiconductor SnS. Appl. Phys. Lett.

385

You might also like