You are on page 1of 11

Article

pubs.acs.org/JPCC

Direct Synthesis of Single-Phase p‑Type SnS by Electrodeposition


from a Dicyanamide Ionic Liquid at High Temperature for Thin Film
Solar Cells
Marc Steichen,*,† Rabie Djemour,† Levent Gütay,† Jérôme Guillot,‡ Susanne Siebentritt,†
and Phillip J. Dale†

Laboratory for Photovoltaics, University of Luxembourg, 41 rue du Brill, L-4422 Belvaux, Luxembourg

Département de Science et Analyse des Matériaux, Centre de Recherche Public Gabriel Lippmann, 41 rue du Brill, L-4422 Belvaux,
Luxembourg
*
S Supporting Information

ABSTRACT: Thin film solar cells based on nontoxic and earth-abundant elements are
necessary for future-generation photovoltaic devices. Tin monosulfide is a promising candidate
that can be used as an absorber material in thin film photovoltaics. In this paper, we introduce
the direct synthesis of stoichiometric and single-phase p-type SnS films via the
electrodeposition from the ionic liquid 1-butyl-3-methylimidazolium dicyanamide ([C4mim]-
[DCA]) containing elemental sulfur and SnCl2 at high temperature. The electrochemical
behavior is studied, and a deposition mechanism of tin monosulfide from the elemental sulfur
saturated ionic liquid is proposed. XRD, XPS, and Raman spectroscopy demonstrate the sole
presence of α-SnS without any secondary sulfide phases (e.g., SnS2, Sn2S3). XPS depth
profiling confirmed the phase purity and disproved the presence of organic contamination in
the as-deposited films. Photoelectrochemical measurements affirmed the p-type conductivity
of the SnS films. The as-deposited layers have an indirect optical band gap at 1.17 eV and high
optical absorption (α ≥ 104 cm−1) at photon energy above 1.4 eV. First solar cells with a
standard thin film substrate cell configuration are presented.

1. INTRODUCTION Various formation methods have been used to grow SnS


Cost-effective, sustainable, and environmentally friendly absorber layers. The first deposition methods are based on
absorber materials are required for thin film photovoltaic vacuum and thermal techniques, such as electron beam
(PV) technologies to establish as a major energy supplier with evaporation,13 thermal evaporation,14,15 RF sputtering,16 spray
terawatt production scale for a low-carbon future.1 The two pyrolysis,7,17 chemical vapor deposition18,19 (CVD), and atomic
most promising thin film PV materials CdTe and Cu(In,Ga)Se2 layer deposition12 (ALD). The second deposition route
could be limited by the scarcity and supply issues of indium and involves solution-based techniques, such as chemical bath
tellurium, restricting large-scale production and increasing the deposition,20 successive ionic layer adsorption and reaction21
production costs.2 New earth-abundant materials could allow (SILAR), and electrodeposition.9,22−30 Besides the formation of
significantly decreasing the costs per kilowatt hour. The the orthorhombic SnS phase, the secondary sulfide phases SnS2
quaternary Cu2ZnSnS4 is such a promising material allowing and Sn2S3 are often codeposited.14,19,23,31 The incorporation of
the fabrication of efficient solar cells.3 However, recent these secondary phases is undesired, as they are n-type
investigations showed issues related to its stability along with semiconductors and are thus detrimental for the p−n
the presence of secondary phases detrimental for solar cell heterojunction formation in the solar cells. High-temperature
performances.4,5 For these reasons, simpler binary absorber nonequilibrium techniques are restricted by the high vapor
materials have been targeted as possible candidates for future pressure of sulfur, which evaporates from the deposits, leaving
thin film PV applications. Tin monosulfide (tin(II) sulfide, α- behind nonstoichiometric films. Nevertheless, pure SnS films
SnS), composed solely of nontoxic and earth-abundant can be formed at high deposition temperatures depending on
elements, is such a potential material and presents suitable the growth method.12,17,18,32 Electrodeposition is an interesting
properties for thin film PV applications: SnS has an indirect low-cost and low-energy growth method for chalcogenide
band gap at around 1.1 eV and a high absorption coefficient (α absorber layers33 as it is compatible with large-scale production
> 104 cm−1) above the direct band gap at 1.3 eV.6−9 Further, it and economical in material usage. The electrochemical
shows intrinsic p-type conductivity and appropriate charge
carrier concentrations of 1013−1016 cm−3.10−12 Despite these Received: November 23, 2012
properties, SnS-based thin film solar cells have only reached low Revised: February 11, 2013
power conversion efficiencies with a maximum of 1.3%.7 Published: February 11, 2013

© 2013 American Chemical Society 4383 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

synthesis of tin monosulfide in aqueous solutions suffers from type sulfide semiconductors have been formed by electro-
the commonly encountered issues in this media; Sn2+ can be deposition from ionic liquids.
oxidized to Sn4+, leading to the formation of secondary SnS2 In this report, we present the direct electrochemical synthesis
and Sn2S3 phases,22,23 and the traces of dissolved oxygen in of stoichiometric and single-phase p-type SnS thin films from
nonbuffered electrolytes results in the coformation of stable tin the ionic liquid 1-butyl-3-methylimidazolium dicyanamide
oxide phases.9,34 Recently, Mathews et al.23 have reported on ([C4mim][DCA]) containing elemental sulfur and SnCl2 on
the electrochemical synthesis of SnS from an aqueous solution Mo-coated glass at temperatures between 80 and 120 °C. The
containing SnCl2 and Na2S2O3 on FTO-coated glass at pH 2.5 SnS films have been characterized for their elemental
and 45 °C. Pulse electrodeposition was used to dissolve any composition, surface morphology, crystal structure, phase
weakly bonded SnS and Sn formed during the deposition purity, and optoelectronic properties. First, SnS thin film
process. Crystalline orthorhombic SnS was grown in the solar cells have shown that the direct electrochemical synthesis
presence of secondary tin sulfide phases. Brownson et al.25,35 route of tin monosulfide absorbers from ionic liquids at high
investigated the galvanostatic electrodeposition of tin mono- temperature have potential for future low-cost PV applications.
sulfide from an aqueous solution of SnCl2 and Na2S2O3
containing a tartaric acid additive on FTO-coated glass at pH 2. EXPERIMENTAL SECTION
1.5−2.5 and temperatures of 50−90 °C. The addition of tartaric 2.1. Reagents and Materials. Anhydrous SnCl2 (Alfa
acid resulted in the formation of a new polymorph (δ-SnS), Aesar, 99.995%) and elemental sulfur (Alfa Aesar, 99.999%)
which can be transformed into α-SnS by thermal annealing at T were used as received. The ionic liquid 1-butyl-3-methylimida-
> 350 °C. zolium dicyanamide, [C4mim][DCA], was purchased from Io-
High deposition temperatures are expected to result in the li-tec (Germany) in the highest available purity (>99%). The
growth of chalcogenide films with enhanced crystal quality. We solvent was further purified prior to usage under vacuum (10−3
have recently reported on the direct potentiostatic electro- bar) together with P2O5 (Sigma-Aldrich) desiccant for 48 h to
deposition of p-type trigonal selenium from the ionic liquid 1- reduce water levels below 100 ppm, determined by coulometric
ethyl-3-methylimidazolium tetrafluoroborate/chloride at tem- Karl Fischer titration. All chemicals have been handled in a
peratures above 100 °C36 and demonstrated that the nitrogen-filled glovebox (MBraun Unilab) with water and
crystallinity of the t-Se was improved by increasing the oxygen levels below 1 ppm.
deposition temperature from 100 to 160 °C. Ionic liquids The S8-saturated deposition baths were prepared by
(IL) are defined as liquids consisting of molecular cations and dissolving elemental sulfur together with SnCl2 at 100 °C.
anions that have a melting point below 100 °C.37,38 ILs exhibit The mixture was then stirred for 2 h at 100 °C and left
intrinsic high thermal stability and give the opportunity to overnight at room temperature. Some of the S8 remained
deposit at temperatures in excess of those possible in water (T undissolved and was kept in the electrochemical cell to ensure
≥ 100 °C), allowing the direct synthesis of crystalline device- the saturation of the deposition bath. The solution was yellow-
quality PV chalcogenide absorbers without further thermal orange, and the saturation concentration of elemental sulfur in
annealing requirements.33 [C4mim][DCA] was estimated to be 10 mM by successive
Boros et al. have shown that ionic liquids are good solvents addition experiments in agreement with reported values.40
for elemental chalcogens.39 The authors showed that the 2.2. Electrochemical Deposition and Characterization
solubility of elemental sulfur strongly dependent on the Lewis of SnS Thin Films. The electrochemical experiments were
basicity and the nucleophilicity of the anion. Manan et al. have performed in the glovebox in a thermally controlled three-
recently investigated the electrochemistry of elemental sulfur electrode cell (±2 °C), connected to an Ecochemie μAutolab
and polysulfides in several ionic liquids on Au, Pt, and GC type III potentiostat. The counter electrode was a large surface
electrodes at room temperature.40 In-situ UV−vis spectroscopy platinum foil, and the reference electrode was a silver wire
in elemental sulfur saturated 1-butyl-3-methylimidazolium directly immersed in the ionic liquid. The working electrode
dicyanamide allowed to identify S62− and S42− as the main consisted of a Mo layer sputtered on soda lime glass with a
cathodic species at room temperature. geometrical surface area of 2 × 2 cm2. The cathodic
Only a few reports on the electrodeposition of sulfide electrochemical synthesis of the SnS films was performed
compound semiconductors from ionic liquids are presented to under potentiostatic control at temperatures between 80 and
date. Dale et al.41 have explored the electrodeposition of n-type 120 °C. Prior to use, the substrates were etched in an aqueous
CdS and ZnS films from the deep eutectic based ionic liquid 30% ammonia solution for 5 min, ultrasonically cleaned in
chlorine chloride:urea. Izgorodin et al.42 have studied the water and ethanol, and dried under a nitrogen flow. After the
electrodeposition of thin CdS thin films on FTO coated glass deposition, the films were allowed to naturally cool to room
from methyltributylphosphonium tosylate containing Na2S2O3 temperature and were sequentially washed with ultrapure water
and CdCl2 at 130−150 °C. The authors proposed the direct and ethanol to remove the remaining ionic liquid and dried
reduction of S2O32− to S2−, which then reacts with Cd2+ to form under nitrogen.
CdS. Elemental sulfur could not be directly used as sulfide X-ray diffraction (XRD) measurements were collected on a
precursor as too negative potentials were needed for the Bruker Discover D8 diffractometer with a Cu Kα1/2 source (λα1
reduction in the phosphonium ionic liquid. Very recently, Chen = 1.54056 Å) and a scintillation detector in grazing incidence
et al.43 demonstrated the electrodeposition of CuS from the configuration (3°). Scanning electron microscopy (SEM) was
ionic liquid 1-ethyl-3-methylimidazolium bis- performed on a Hitachi SU-70 equipped with an Oxford
(trifluoromethanesulfonyl imide) ([C2mim][Tf2N]) containing Instruments INCA X-MAX analyzer for energy-dispersive X-ray
Cu(Tf2N)2 and elemental sulfur on Pt at high temperature (T (EDX) and wavelength-dispersive X-ray (WDX) spectroscopy
≥ 120 °C). Copper(II) sulfide films have been formed at low to distinguish the overlapping S Kα and Mo Lα peaks. Raman
overpotential (E = +0.25 V vs Ag) in the regime of Cu0 spectroscopy was performed on a confocal home-built setup
deposition and positive of the sulfur reduction. So far, no p- equipped with a piezo-element for scanning measurements. All
4384 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

Raman measurements were line scans of 10 μm performed with


an excitation wavelength of 514 nm at 2 mW. The excitation
diameter of the laser area is 1 μm. The Raman analysis depth is
estimated to be 100 nm, based on the calculated absorption
coefficient (section 3.6.2.).
X-ray photoelectron spectroscopy (XPS) analyses were
realized on a Kratos Axis-Ultra DLD instrument using a
monochromatic Al Kα X-ray source (hν = 1486.6 eV, 150 W)
at a pass energy of 160 and 20 eV for survey and high-
resolution spectra, respectively. In order to obtain depth
profiles, the samples were etched prior to the analysis using an
Ar beam (3 kV, 0.2 mA) at an incident angle of 45°. The XPS
analysis spot was 110 μm in diameter.
Photoelectrochemical measurements were carried out in a
three-electrode setup in 0.2 M Eu(NO3)3 aqueous solution with
platinum counter and Ag|AgCl (sat.) reference electrodes. The
absorber layers were illuminated with a white light LED source
chopped at constant frequency on a home-built apparatus
during potential sweep and potentiostatic measurements.
The optical absorption properties of the samples were
determined by optical transmission measurements, performed
on a PerkinElmer Lambda 950 UV−vis spectrometer equipped
with an Ulbricht sphere. The recorded optical transmission data
were evaluated by an optical modeling tool kit in the program
Diplot44 as described in detail here.45 The modeling includes
the optical absorption behavior of the absorber material as such
and the geometrical properties and structure of the sample.
After the electrodeposition of the p-type SnS absorber layer,
photovoltaic devices were completed by chemical bath
deposition (CBD) of a ca. 50 nm thick CdS buffer layer,
followed by radio-frequency (RF) sputtering of an intrinsic
ZnO (i-ZnO) layer and a n-type Al2O3-doped ZnO (Al:ZnO)
window layer. Finally, Ni−Al front contacts were formed on
top via e-beam evaporation.
The final device configuration is Mo/SnS/CdS/i-ZnO/ Figure 1. Cyclic voltammetry (CV) of the ionic liquid [C4mim]-
Al:ZnO/Ni−Al. Current−voltage (I−V) characteristics of the [DCA] containing (a) saturated S8, (b) 20 mM SnCl2, and (c)
devices have been measured on a home-built I−V setup, saturated S8−20 mM SnCl2 on a Mo electrode, T = 100 °C, vscan = 20
equipped with a halogen lamp as light source. Under mV s−1. Pt counter electrode, Ag pseudoreference electrodes. The
illumination, the system is calibrated to 100 mW cm−2 at 25 arrows indicate the scan direction.
°C. The devices have been further characterized by measure-
ments as a function of illumination wavelength to generate liquid at room temperature. In-situ UV−vis spectroscopy
external quantum efficiency (EQE) spectra. The EQE spectra showed that S62− is the main reduction product at c1 and
have been recorded on a home-built apparatus and allow the S42− is the main species formed at the cathodic peak c2 on Au
extraction of the bandgap of the SnS. electrode at room temperature. At this temperature, the ionic
liquid prevents the further dissociation reactions of the dianions
3. RESULTS AND DISCUSSION S62− and S42− into lower polysulfides Sx2− (x < 4). The
3.1. Cyclic Voltammetry Studies. To understand the formation of the blue radical S3•− is not observed in
electrochemistry of S8 and SnCl2 in [C4mim][DCA] at high [C4mim][DCA] at room temperature, contrary to reports
temperature, cyclic voltammetry (CV) studies have been from other ionic liquids39 and molecular solvents, such as
performed. Figure 1 shows the CV of the ionic liquid dimethylformamide46 and acetonitrile,47 where the decom-
[C4mim][DCA] containing (a) saturated S8, (b) 20 mM position of S42− into S3•− and other Sx2− (x < 4) is observed.
SnCl2, and (c) saturated S8−20 mM SnCl2 on a Mo electrode The coloration of the S8-saturated ionic liquid [C4mim][DCA]
at 100 °C. The CV in Figure 1a is started at the open circuit is yellow-orange at 100 °C; thus, the formation of large
potential (OCP) at +0.05 V and scanned into the cathodic amounts of the radical S3•− is also excluded at higher
direction until the turning potential at −1.2 V with a scan rate temperatures.
of 20 mV s−1. Two distinct reduction peaks c1 and c2 are clearly Figure 1b presents the CV of [C4mim][DCA] containing 20
observed at −0.33 and −0.54 V, respectively. On the backward mM SnCl2 on a Mo at 100 °C. The potential is scanned in the
scan, two anodic current peaks a2 and a1 are recorded at −0.37 negative direction from the OCP at +0.05 V, and deposition of
and +0.01 V. metallic tin (peak cSn) occurs at −0.63 V, which is more
The anodic peak a1 is related to the cathodic peak c1 (and a2 negative than the reduction waves of elemental sulfur under the
is related to c2) as indicated by experiments realized with same conditions (curve a). On the reverse scan, a typical
different cathodic turning potentials. Recently, Manan et al.40 nucleation loop is observed and is followed by the stripping
studied the electrochemistry of elemental sulfur in this ionic peak at −0.16 V, presenting a shoulder at −0.30 V, which could
4385 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

be explained by different tin speciation during the stripping films deposited at E ≤ −0.8 V are also very Sn-rich. Even
process.48 The further oxidation of Sn2+ to Sn4+ is not observed though, electroactive sulfides are formed at these potentials;
on the Mo electrode. Control experiments have been their formation rate is too low to keep up with the Sn metal
performed on platinum (Pt) and glassy carbon (GC) deposition, which is favored compared to the kinetically
electrodes; as in the case of Mo, no further oxidation of Sn2+ hindered SnS compound formation reaction.49
to Sn4+ was recorded on Pt. However, large oxidation currents Figure 2 shows (A) the XRD diffractograms and (B) the
were observed on a GC electrode (cf. Supporting Information). corresponding SEM top-view and cross-section images of the
These results agree with previous studies48 realized in a similar as-deposited SnS films formed at (a) E = −0.7 V and (b) E =
ionic liquid [C2mim][DCA]−50 mM SnCl2 at 40 °C, where −0.8 V.
the oxidation of Sn2+ was absent on Pt, but was present on GC The reference pattern (JCPDS card 39-0354) of ortho-
electrodes. rhombic SnS with the lattice constants a = 0.4329 nm, b =
Figure 1c shows the CV of the S8-saturated ionic liquid 1.1192 nm, and c = 0.3984 nm is included for comparison. For
[C4mim][DCA] containing 20 mM SnCl2 on the Mo electrode both samples, all the diffraction peaks correspond to the
at 100 °C. Starting the cathodic scan from the OCP at +0.10 V, orthorhombic SnS phase. The average coherence lengths D of
a first cathodic wave is observed at −0.40 V similar to the the crystals have been estimated from the main (111) Bragg
response c1 in the CV of Figure 1a. A second large cathodic peak of orthorhombic SnS using the Scherrer formula:
current cSnS is recorded at −0.67 V. On the backward scan, no

stripping peak is recorded, suggesting the formation of a stable D=
sulfide compound. β cos(θ)
3.2. Electrochemical Synthesis and Characterization where D is the average coherence length (nm), k = 0.9 is the
of As-Deposited SnS Thin Films. On the basis of the above- shape factor (spherical grain shape), λ = 1.540 56 Å is the
mentioned CV studies, the electrodeposition of SnS films has wavelength of the incident radiation, β is the full width at half-
been performed under potentiostatic control from S8-saturated maximum fwhm (rad), and θ is the Bragg angle (rad). The
[C4mim][DCA] containing 20 mM SnCl2 on Mo at 100 °C. FWHMs have been determined by fitting Gaussian profile
Table 1 shows the WDX/EDX analysis of films deposited in the peaks to the experimental data. As the crystals are not exactly of
spherical symmetry, the values have to be considered as
Table 1. WDX/EDX Analysis of the As-Deposited Thin approximations. Similar average coherence length D = 20.5 ±
Films Deposited on Mo from [C4mim][DCA]−S8 sat.−20 0.4 nm have been extracted for both films. These values are
mM SnCl2 at Different Potentials for 30 min at 100 °Ca smaller than the average grain sizes observed in the SEM,
Edep (V) WDX/EDX (at. %) Sn/S XRD detected phases showing that the SnS crystallites are subject to crystal defects.
−0.5 2.49 ± 0.09 SnS, Sn
However, the estimated average coherence lengths are
−0.6 1.25 ± 0.08 SnS, Sn
significantly higher than values previously reported for
−0.7 0.98 ± 0.05 SnS
electrodeposited SnS films from aqueous23 and nonaqueous26
−0.8 1.05 ± 0.06 SnS
solutions and are close to the average grain size values reported
−0.9 5.1 ± 0.1 Sn, SnS
by thin SnS films grown by ALD at 120 °C,12 showing that
a higher temperatures are also favorable for crystal growth during
SEM images are provided as Supporting Information.
electrodeposition. The SEM top-view image in Figure 2Ba
shows the SnS film deposited at E = −0.7 V. The film presents
potential range from −0.5 to −0.9 V. Stoichiometric films are a good surface coverage. The film is compact and exposes no
grown at −0.8 V < E < −0.7 V in the region of the peak cSnS. visible pinholes to the Mo back-contact electrode. The film is
On either side of this potential range, Sn0 is incorporated in the formed by larger clusters composed of agglomerated
films. orthorhombic SnS crystals with grain sizes ranging from 100
At lower applied deposition potentials (E ≥ −0.6 V) the to 150 nm. The cross-section image confirms these
polysulfide species formed at the peak c1 do not formed rapidly observations; uniformly dispersed and well-defined crystals
enough to participate in the formation of pure SnS films. On are visible throughout the 800 nm thick film to form a compact
the other side of the stoichiomentric SnS potential range, the single-phase SnS layer. The surface roughness of the layer is

Figure 2. (A) XRD diffractograms of SnS films deposited at (a) E = −0.7 V and (b) E = −0.8 V from [C4mim][DCA]−sat. S8−20 mM SnCl2 on
Mo (∗). The JCPDS pattern 39-0354 (Herzenbergite SnS) is included for comparison. (B) SEM top-view and cross-section images of the two
corresponding SnS films.

4386 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393


The Journal of Physical Chemistry C Article

Figure 3. (A) XRD diffractograms of three films deposited at different temperatures from [C4mim][DCA]−sat. S8−20 mM SnCl2 on Mo (∗). (B)
Corresponding high-resolution XRD of the main (111) Bragg peaks. The JCPDS patterns 39-0354 (Herzenbergite SnS) and 04 0673 (Sn) are
included for comparison.

estimated to be ∼100 nm. The films were adherent and brown- [C4mim][DCA] containing 20 mM SnCl2 at constant over-
matte black in appearance. The films deposited at E = −0.8 V potentials. At T < 80 °C, no SnS could be grown, solely metallic
present a much rougher surface morphology and were black Sn0 was deposited. Only at T ≥ 80 °C was α-SnS formation
and matte in appearance. Needle-like elongated, dendritic SnS possible. As shown on the XRD diffractogram in Figure 3,
crystallites are randomly grown on the top surface. The cross- however, the codeposition of SnS with Sn was observed. At T =
sectional image adds more information to the growth of the 100 °C, pure orthorhombic SnS films are confirmed by the
SnS layer. The dendritic morphology of the SnS crystals XRD analysis.
observed in the top-view image of the film grown at −0.8 V is At a deposition temperature of 120 °C, well-crystallized α-
confirmed; however, a compact, dense, and well-crystallized SnS was deposited; nevertheless, we have observed that the
SnS film of 500−600 nm is revealed on the Mo substrate. deposition bath was unstable at T ≥ 120 °C. A yellow
This compact underlayer has been systematically observed amorphous SnSx precipitate formed after 1 h at these
for all SnS films. In a first stage, after the initial nucleation on temperatures, depleting the deposition solution of sulfur and
the Mo surface, a thin compact SnS layer (cf. Supporting leading to the codeposition of Sn0 together with orthorhombic
Information) is quickly formed and covers the Mo electrode SnS. No deposition of single-phase SnS films was possible at T
surface completely after 1 min. Depending on the applied > 120 °C. The precipitation of colloidal SnS2 has been
overpotential, the SnS layer either continues to grow thicker previously reported in ethylene glycol of elemental sulfur and
compactly (−0.7 V) or enters a dendritic growth regime (−0.8 SnCl2 at 70 °C.26 In comparison to this organic solvent,
V). The presence of the thin compact SnS layer at the early [C4mim][DCA] appears to stabilize elemental sulfur better up
growth stage can be explained by the fact that α-SnS is a p-type to 120 °C. At higher temperatures, the imidazolium cation may
semiconductor with several orders of magnitude fewer electrons undergo reactions with elemental sulfur in the presence of a
than a metal. The formation reaction for SnS is caused by a base, as reported previously.51
reduction current, meaning that electrons are flowing from the Table 2 summarizes the estimated coherence length D from
the main Bragg peak (111) of the SnS films deposited at 80,
electrode to reduce the species in solution. The further
reduction of SnS onto a metallic electrode covered with SnS
nuclei will occur more rapidly on the bare metallic surface of Table 2. Evaluation of the As-Deposited Phases and
the electrode due to its disproportionally higher electron Coherence Length D of SnS on Mo from [C4mim][DCA]−
S8 sat.−20 mM SnCl2 at E = −0.7 V for 30 min at Different
density compared to the p-type SnS. Once the compact p-type
Temperaturesa
layer formed, the electrodeposition of SnS continues in a slower
second stage on the p-type film due to the lack of available Tdep (°C) XRD detected phases DSnS (nm)
electrons for electrochemical reduction,50 as evidenced by the 80 SnS, Sn 15.1 ± 0.6
low current densities in ambient light conditions (j ≈ 0.2 mA 100 SnS 20.5 ± 0.4
cm−2). One way to generate more electrons in a p-type 120 SnS, Sn 25.0 ± 0.6
semiconductor is to illuminate its surface with photons of a
The evaluation is realized on the main (111) Bragg peak.
energies higher than the band gap. These photogenerated
electrons may then participate in the reduction reaction.
Complementary electrodeposition experiments were performed 100, and 120 °C. As expected, the crystal quality of the SnS
under constant illumination (W = 100 W m−2) and showed crystals increases with temperature.
enhanced deposition current densities by a factor of 2 (j ≈ 0.4 3.4. SnS Phase Purity and Film Contamination. The
mA cm−2). unambiguous detection of minor secondary tin sulfide phases is
3.3. Effect of Deposition Temperature. The influence of not straightforward by XRD due to the overlapping of several
the deposition temperature has been studied from S8-saturated Bragg peaks of SnS, SnS2, and Sn2S3.19 Raman spectroscopy has
4387 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

been performed in order to assess the phase purity and lateral stoichiometric SnS film from S8-saturated [C4mim][DCA]
uniformity of the as-deposited SnS films. Figure 4 shows the containing 20 mM SnCl2 at −0.7 V at 100 °C. The binding
energies measured in the XPS analysis are standardized using
the C 1s signal as reference peak at 284.9 eV. The Sn 3d peaks
measured at 486.6 and 495.0 eV correspond to the binding
energy of Sn 3d5/2 and Sn 3d3/2, and the peak at a binding
energy of 161.1 eV corresponds to S 2p3/2 of tin monosulfide
(high-resolution spectrum in Figure 5Ab). These results are in
good agreement with reported values for pure SnS layers.12,23,29
No other XPS peaks at higher binding energies of tin oxide or
other tin sulfide phases were detected. The low binding energy
shoulders at 485.4 and 493.8 eV are attributed to metallic
Sn023,29 and are due to the partial loss of volatile sulfur under
ultrahigh vacuum in the XPS chamber. SnS(s) is in a pressure-
dependent equilibrium with Sn(s) and S2(g);5 sulfur evaporates
from the SnS surface in UHV, leaving metallic Sn0 on the film
surface.
XPS analysis has been performed on the as-deposited SnS
films in order to check that no organic contamination
originating from the ionic liquid has occurred. Figure 5Ba
shows representative XPS depth-profiling survey spectra of a
stoichiometric SnS thin film deposited on Mo from [C4mim]-
[DCA]−sat. S8−20 mM SnCl2 at E = −0.7 V and T = 100 °C.
The corresponding high-resolution scans of the C 1s and O 1s
signals are also represented in Figure 5Bb. O 1s and C 1s peaks
have only been detected at the surface of the SnS films due to
exposure to ambient air. After 800 s of Ar sputtering, no further
Figure 4. Raman spectroscopy of a SnS film electrodeposited from impurities were found in the bulk of the SnS layer (detection
[C4mim][DCA]−sat. S8−20 mM SnCl2 on Mo at −0.7 V, T = 100 limit of 0.5 at. %). The same trend was observed for nitrogen
°C. The bar plots represent the measured relative peak intensities of contaminations. No additional signals originating from Cl, Ag
SnS, Sn2S3, and SnS2 single-crystals for comparison. (The intense or Pt have been detected.
mode at 312 cm−1 of the SnS2 single crystal has been reduced by a 3.5. Electrochemical Growth Mechanism of SnS Thin
factor of 5 for clarity.) Films. In light of the above-mentioned studies, a mechanism
for the electrochemical formation of the SnS absorbers from
characteristic Raman peaks at 97, 160, 191, 218, and 288 cm−1, [C4mim][DCA]−S8 sat.−SnCl2 is proposed.
in agreement with single-phase orthorhombic SnS films The polysulfide S42− is formed cathodically at E ≤ −0.6 V,
prepared by ALD12 and atmospheric pressure chemical vapor and the dissociation reactions by disproportionation are slow in
deposition (APCVD).19 Raman line scans on the sample [C4mim][DCA] at room temperature.40 The disproportiona-
surface did not show any lateral variations, confirming the tion of elemental sulfur are thermally activated reactions52 (T >
lateral homogeneity of the SnS thin film. Additional Raman 80 °C). Thermal dissociation of sulfur by hemolytic scission
analysis has been performed on tin sulfide single crystals (SnS, occurs at high temperatures.53 These findings are coherent with
Sn2S3, and SnS2), and the relative peak intensities of these pure our observations in the ionic liquid [C4mim][DCA] at elevated
single-crystalline materials are shown in Figure 4 for temperatures. On the basis of our experiments, we propose that
comparison. The Raman peaks of the as-deposited SnS film the formation of the electroactive S2−, required for SnS
correspond very well with all the observed Raman modes deposition, occurs via the electrochemical formation of S42−
measured on the SnS single crystal. followed by the thermally activated disproportionation by the
No Raman peaks were observed at 307 and 312 cm−1, which following mechanism:
are the strongest modes for Sn2S3 and SnS2, respectively. For
clarity, the intense mode of SnS2 single crystal is reduced by a 2S4 2 − ⇔ S6 2 − + S2 2 −
factor of 5 in Figure 4. The relative high intensity of these
modes compared to those of SnS allows to conclude that no 3S2 2 − ⇔ 2S2 − + S4 2 −
secondary SnS2 and Sn2S3 phases are present and show that the
compact as-deposited SnS thin films prepared on Mo by the The growth and the composition of the film are ultimately
direct electrochemical synthesis from S8-saturated [C4mim]- controlled by the relative rates of the S2− formation and the
[DCA] containing 20 mM SnCl2 at −0.7 V at 100 °C are pure, reduction of Sn2+ at the electrochemical interface. As presented
single-phase orthorhombic SnS layers. This result shows in section 3.3, the formation of S2− by the coupled
significant improvements over previously reports, where electrochemical reduction−dissociation reactions is only fast
secondary tin sulfide and oxide phases have been present in enough at T ≥ 100 °C. At this temperature, stoichiometric
the as-deposited films formed by electrodeposition from single-phase SnS films are grown without the codeposition of
aqueous media.23,34 metallic Sn. At lower temperatures, the reduction rate of Sn2+ to
X-ray photoelectron spectroscopy (XPS) analysis has been Sn0 is too high compared to the S2− formation rate and mixed
performed to study the surface and bulk phase of the as- SnS/Sn films are deposited. By increasing the temperature
deposited films. Figure 5Aa shows the survey spectrum of a above 120 °C, the dissociation rate of polysulfide in the ionic
4388 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

Figure 5. (A) (a) XPS survey spectrum of a SnS film electrodeposited from [C4mim][DCA]−sat. S8−20 mM SnCl2 on Mo at −0.7 V and T = 100
°C. (A) (b) Corresponding high-resolution spectra of the Sn 3d peak and S 2p peak regions. (B) (a) XPS depth-profile spectra of the same SnS film
as a function of Ar sputtering. (B) (b) Corresponding high-resolution XPS spectra of the C 1s and O 1s peak regions.

liquid is very high and results in the formation of amorphous °C). Copper(II) sulfide was formed by the chemical reaction of
SnSx precipitates, depleting the ionic liquid in sulfur. the first deposited Cu0 and the molecular sulfur dissolved in the
The reduction potential of Sn2+ cations and the potential of ionic liquid. These differences are in agreement with previous
2−
S formation lie close together. Thus, the formation of tin reports on the sulfurization reaction of metals, which showed
monosulfide could occur alternatively via two different that copper sulfide formation is 104 times faster than the
mechanisms: (1) The chalcogen can react with the first reaction of tin with sulfur.49 Accordingly, we have demon-
deposited layer of metallic Sn0 and convert into SnS on the strated that the formation of SnS is not possible via this
surface. (2) The generated chalcogen reacts with Sn2+ cations at reaction mechanism. Izgorodin et al.42 have concluded on a
the electrochemical interface to form SnS on the electrode similar reaction scheme for the electrodeposition of CdS thin
surface. films on FTO-coated glass from the ionic liquid methyltribu-
To better understand and discriminate between the two tylphosphonium tosylate containing sodium thiosulfate as sulfur
mechanisms, we have investigated the growth of SnS on Sn precursor and CdCl2 at 130−150 °C. Analogically to our case,
electrodes in a S8-saturated [C4mim][DCA] at 100 °C. The Sn
the reduction potentials of Cd2+/Cd0 and S2O32−/S2− are close
electrode was immersed in the liquid at OCP, and no reaction
together, and the formation reaction of CdS proceeded via the
with elemental sulfur was observed after 30 min. The same
reaction between S2− and Cd2+ at the electrode interface.
observations have been made after applying a potential of −0.7
In summary, we can discard the SnS formation via path 1 and
V to the tin electrode for 30 min; negligible amounts of SnS
conclude that the overall formation reaction of orthorhombic,
have been formed in the S8-saturated [C4mim][DCA] at 100
°C. In contrast, when the same experiments were realized on a single-phase SnS in the S8-saturated [C4mim][DCA] ionic
Cu electrode instead of Sn, instantaneous formation of black liquid containing SnCl2 at high temperature is controlled by the
CuxS deposits have been observed at OCP in the S8-saturated electrochemical reduction of S8 to S42− followed by the
[C4mim][DCA] at 100 °C, confirming the results of Chen et thermally activated dissociation producing S2− anions, which
al.,43 who have very recently studied the electrodeposition of subsequently react with Sn2+ cations at the electrochemical
CuS in the ionic liquid 1-ethyl-3-methylimidazolium bis- interface (int):
(trifluoromethanesulfonyl imide) containing Cu(Tf2N)2 and
elemental sulfur on Pt electrodes at high temperature (T ≥ 120 Sn int 2 + + Sint 2 − → SnSsurf

4389 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393


The Journal of Physical Chemistry C Article

As evidenced in Figure 1, the oxidation of Sn2+ to Sn4+ is not


observed on Mo electrodes in [C4mim][DCA], in contrast to
GC electrodes (cf. Supporting Information). Tin is stable in the
(2+) oxidation state in this ionic liquid, contrary to aqueous
electrolytes54,55 where Sn2+ can be readily oxidized to Sn4+.
Thus, the deposition of pure, single-phase tin monosulfide is
possible, avoiding the formation of the secondary phases SnS2
and Sn2S3.
Potentiostatic depositions with lower [SnCl2] = 10 mM in
S8-saturated [C4mim][DCA] have revealed that the deposition
current is independent of the metal cation concentration,
confirming the proposed formation mechanism. As long as Sn2+
cations are present at high enough concentration ([Sn2+]int ≥
[S2−]int) at the electrochemical interface, the generation of
reactive S2− species is the rate-determining step of the overall
SnS formation reaction.
As shown, stoichiometric SnS is only formed in a narrow Figure 6. Photoelectrochemical sweep voltammetry of a representative
voltage range between −0.7 and −0.8 V under potentiostatic SnS film in contact with a 0.2 M Eu3+ aqueous electrolyte under
control. The rate of S2− formation is directly proportional to chopped white illumination, vscan = 10 mV s−1, T = 25 °C. Inset shows
the current density at constant temperature. For large-scale the corresponding photocurrent response at E = −0.6 V (vs Ag|AgCl
production, galvanostatic electrodeposition on highly conduct- sat.).
ing substrates would allow the fine-tuning of the SnS growth by
controlling the rate of S2− generation. Current control and shows that a constant photocurrent is measured over several
careful deposition cell design would allow establishing a robust illumination periods. Sharp current−time transients during the
deposition system for uniform, stoichiometric SnS thin films dark-light switching are observed and are characteristic for good
from ionic liquids at high temperature on large area substrates. diode behaviors.
3.6. Optoelectronic Characterization. 3.6.1. Photocon- 3.6.2. Optical Absorption Properties. The absorption
ductivity. Biased photoelectrochemical measurements under behavior of the as-deposited SnS films was determined by
chopped white light illumination of the as-deposited SnS layers optical transmission measurements and modeling of the
have been performed. In contact with an appropriate redox experimental data in Diplot.44 In Figure 7A, the optical
electrolyte, a semiconductor film is expected to show a transmission spectrum of a ca. 300 nm thick SnS film is shown.
rectifying diode behavior in the polarization region of the The black curve is the experimental data, and the red curve
minority charge carrier current flow. In the case of a p-type shows the modeling results. The measured spectrum presents a
semiconductor (minority electron charge carriers), this Fabry−Perot oscillation with its maximum at 0.83 eV. This
behavior is expected at potentials negative of its flatband oscillation is best fitted with a refractive index of 2.35 and a film
potential. The direction of the photocurrent (anodic or thickness of 305 nm, which matches well the estimation of the
cathodic) allows the determination of the conductivity type. growth rate of the electrodeposition process in [C4mim]-
For a p-type semiconductor under illumination, the photo- [DCA]−sat. S8−20 mM SnCl2 at −0.7 V and T = 100 °C after
generated electrons manifest as a cathodic photocurrent. 15 min. The fitted curve shows the result of a model including
Figure 6 shows the photoelectrochemical behavior of an as- an indirect band gap Eg = 1.17 eV and Urbach energy of 50
deposited SnS thin film on Mo from [C4mim][DCA]−sat. S8− meV. The experimental data of the as-deposited SnS film were
20 mM SnCl2, E = −0.7 V and T = 100 °C. The sweep best fitted to this indirect transition model; i.e., the minimum of
voltammetry of the Mo/SnS electrode was performed in the conduction band is not at the Γ point.
contact with a 0.2 M Eu(NO3)3 aqueous electrolyte, scanning The extracted absorption coefficient is plotted as a function
from OCP = −0.07 V into the negative direction. The chopping of the photon energy in Figure 7B. Efficient absorption
period was asymmetric (longer illumination period). The dark coefficients α ≥ 104 cm−1 are observed at E ≥ 1.4 eV. These
current at OCP is low and increases gradually toward more values are consistent with the theoretical calculations performed
negative potentials, which can be related to the direct contact of by Vidal et al.,8 who concluded that tin monosulfide has a 1.07
uncovered Mo with the redox electrolyte and the onset of the eV indirect band gap with an effective absorption onset located
hydrogen evolution reaction in the Eu(NO3)3 aqueous 0.4 eV higher. From the plot of the absorption coefficient in
electrolyte. During the illumination periods, negative photo- Figure 7B, we can confirm very similar behavior of the as-
currents are clearly seen in the cathodic direction, demonstrat- deposited SnS films from [C4mim][DCA]−sat. S8−20 mM
ing the p-type photoconductivity of the as-deposited SnS layers. SnCl2. Effective absorption is achieved at E ≥ 1.5 eV, with
The photocurrent increases with the cathodic polarization, absorption coefficients of α ≥ 2 × 104 cm−1, which is sufficient
which is due to the increased band bending at the for complete light absorption in a typical 1 μm thick absorber
semiconductor−electrolyte junction and the filling up of layer for thin film solar cells. Makinistian et al.6 have previously
surface traps, both of which allow the minority carriers to be used ab initio DFT calculations to compute the electronic band
extracted more easily. The onset of the photocurrent response structure and also determined an indirect band gap of 1.16 eV
is located at E ≈ −0.2 V. The inset of Figure 6 presents the with anisotropic behavior along the three main crystallographic
chopped photoelectrochemical behavior of the as-deposited axes.
SnS film at a constant negative polarization of −0.6 V vs Ag| Experimentally, the indirect band gap of SnS has been
AgCl sat. The measurement confirms the p-type photoresponse observed by several groups.13,16,26,56 Hartman et al.16 have
of the as-deposited SnS thin films at negative polarization and recently deposited SnS thin films by RF sputtering under
4390 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

Figure 7. (A) Optical transmission measurements (black) and Figure 8. (A) EQE spectrum of the best Mo/SnS/CdS/i-ZnO/
modeling (red) of a 300 nm thick, compact, single-phase SnS absorber Al:ZnO/Ni−Al solar cell. (B) J−V characteristic curve of the
layer. (B) Absorption coefficient extracted from the model including corresponding SnS device in the dark and under illumination. SnS
the indirect transition of the SnS layer. films electrodeposited from [C4mim][DCA]−sat. S8−20 mM SnCl2 at
−0.7 V and T = 100 °C.

various growth conditions. All the films showed indirect band


gaps in the range of 1.08−1.18 eV and α ≥ 2 × 104 cm−1 at E >
1.5 eV. Nonetheless, a large discrepancy still exists in the of the absorber layer. The EQE measurements confirm the
literature on the interpretation of the electronic band structure optical absorption measurements.
and the nature of the band gap in SnS absorbers, as several Figure 8B shows the current density−voltage (J−V)
reports used direct transition models to fit the absorption characteristics curve in the dark and under illumination of the
coefficient. Band gap values ranging from 1.3 to 1.8 eV are corresponding SnS device. The cell achieved a 0.17 ± 0.02%
reported.
total area power conversion efficiency with VOC = 155 ± 0.10
The optical absorption properties of the as-deposited SnS
mV, JSC = 3.42 ± 0.80 mA cm−2, and FF = 31.8 ± 0.3% on a
absorber films from [C4mim][DCA]−sat. S8−SnCl2 seem
total area of 0.20 ± 0.02 cm2. The errors are evaluated based on
satisfactory with high absorption coefficients above 1.5 eV,
compatible for the application of the SnS films as absorbers in the three best cells. Nevertheless a low shunt resistance Rp =
PV devices. 97.8 Ω cm2 around JSC is observed, lowering the power by
3.7. SnS Thin Film Solar Cells. Complete solar cell devices allowing alternative current paths between the cell’s front and
were realized in order to demonstrate the potential of the as- back. These losses can be due to absorber defects or/and metal
deposited SnS films formed by the direct synthesis from the impurities present in the absorber layer. The series resistance of
ionic liquid [C4mim][DCA]−sat. S8−20 mM SnCl2 at high the cell is RS = 1.44 Ω cm2.
temperature. Even though the performances are still low, the achieved
Figure 8A shows the external quantum efficiency (EQE) power conversion efficiencies are comparable with previously
spectrum of the best performing SnS solar cell. The spectrum reported values of 0.20% and 0.17% for SnS devices formed by
shows a maximum EQE of 30% at 520 nm and presents the electrodeposition30 and sputtering,31 respectively. So far,
typical CdS absorption in the wavelength region below 500 nm. absorber films synthesized by spray pyrolysis7 gave the highest
The current decrease at long wavelengths above 550 nm is reported efficiency for SnS based solar cells of 1.3%, with VOC =
characteristic of moderate absorption and transport properties 260 mV, JSC = 9.6 mA cm−2, and FF = 53%.
4391 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393
The Journal of Physical Chemistry C Article

4. CONCLUSIONS (2) Candelise, C.; Winskel, M.; Gross, R. Implications for CdTe and
CIGS technologies production costs of indium and tellurium scarcity.
In summary, we have introduced a new low-cost methodology Prog. Photovoltaics Res. Appl. 2012, 6, 816−831.
for the direct synthesis of single-phase orthorhombic, p-type (3) Todorov, T. K.; Reuter, K. B.; Mitzi, D. B. High-efficiency solar
SnS thin films from S8-saturated [C4mim][DCA] containing cell with earth-abundant liquid-processed absorber. Adv. Mater. 2010,
SnCl2 at high temperature. The electrochemical behavior was 22 (20), E156−E159.
studied, and a mechanism of the growth of tin monosulfide was (4) Redinger, A.; Berg, D. M.; Dale, P. J.; Siebentritt, S. The
proposed. The stoichiometry, composition, and morphology of consequences of kesterite equilibria for efficient solar cells. J. Am.
the thin films are controlled by the applied potential and the Chem. Soc. 2011, 133 (10), 3320−3323.
deposition temperature. We conclude that the SnS formation (5) Scragg, J. J.; Ericson, T.; Kubart, T.; Edoff, M.; Platzer-Björkman,
proceeds via the interfacial reaction between S2− and Sn2+. Tin C. Chemical insights into the instability of Cu2ZnSnS4 films during
annealing. Chem. Mater. 2011, 23 (20), 4625−4633.
is stable under its (2+) oxidation state in [C4mim][DCA] and
(6) Makinistian, L.; Albanesi, E. A. On the band gap location and
allows the synthesis of single-phase tin monosulfide at high core spectra of orthorhombic IV-VI compounds SnS and SnSe. Phys.
temperature on Mo electrodes. XRD, XPS, and Raman Status Solidi B 2009, 246 (1), 183−191.
spectroscopy clearly demonstrated the sole presence of the (7) Ramakrishna Reddy, K. T.; Koteswara Reddy, N.; Miles, R. W.
orthorhombic SnS without any secondary phases in the as- Photovoltaic properties of SnS based solar cells. Sol. Energy Mater. Sol.
deposited layers. XPS depth profiling confirmed that no carbon Cells 2006, 90 (18−19), 3041−3046.
and nitrogen impurities were incorporated in the deposits. The (8) Vidal, J.; Lany, S.; d’Avezac, M.; Zunger, A.; Zakutayev, A.;
as-deposited SnS films showed an indirect band gap at 1.17 eV Francis, J.; Tate, J. Band-structure, optical properties, and defect
and efficient absorption above 104 cm−1 at photon energy physics of the photovoltaic semiconductor SnS. Appl. Phys. Lett. 2012,
higher than 1.4 eV, in agreement with theoretical calculation of 100 (3), 4.
SnS band structure.8 Although the photovoltaic device studies (9) Ichimura, M.; Takeuchi, K.; Ono, Y.; Arai, E. Electrochemical
deposition of SnS thin films. Thin Solid Films 2000, 361, 98−101.
are still preliminary, we have successfully demonstrated that the
(10) Noguchi, H.; Setiyadi, A.; Tanamura, H.; Nagatomo, T.; Omoto,
as-deposited p-type tin monosulfide absorbers can be O. Characterization of vacuum-evaporated tin sulfide film for solar-cell
successfully finalized into working solar cells with performances materials. Sol. Energy Mater. Sol. Cells 1994, 35 (1−4), 325−331.
comparable to other deposition methods. (11) Nassary, M. M. Temperature dependence of the electrical
Ongoing research focuses on the further development of the conductivity, Hall effect and thermoelectric power of SnS single
direct electrochemical synthesis of SnS thin films from various crystals. J. Alloys Compd. 2005, 398 (1−2), 21−25.
ionic liquids allowing the growth at temperatures above 200 °C (12) Sinsermsuksakul, P.; Heo, J.; Noh, W.; Hock, A. S.; Gordon, R.
for photovoltaic applications. G. Atomic layer deposition of tin monosulfide thin films. Adv. Energy


Mater. 2011, 1 (6), 1116−1125.
(13) Tanusevski, A.; Poelman, D. Optical and photoconductive
ASSOCIATED CONTENT properties of SnS thin films prepared by electron beam evaporation.
* Supporting Information
S Sol. Energy Mater. Sol. Cells 2003, 80 (3), 297−303.
(1) SEM top-view images of films deposited at different (14) Miles, R. W.; Ogah, O. E.; Zoppi, G.; Forbes, I. Thermally
potentials; (2) SEM cross-section image of the initial compact evaporated thin films of SnS for application in solar cell devices. Thin
Solid Films 2009, 517 (17), 4702−4705.
p-type SnS underlayer; (3) CVs of Mo, Pt, and GC electrodes
(15) Devika, M.; Reddy, K. T. R.; Reddy, N. K.; Ramesh, K.;
in [C4mim][DCA]−20 mM SnCl2 at 100 °C. This material is Ganesan, R.; Gopal, E. S. R.; Gunasekhar, K. R. Microstructure
available free of charge via the Internet at http://pubs.acs.org. dependent physical properties of evaporated tin sulfide films. J. Appl.

■ AUTHOR INFORMATION
Corresponding Author
Phys. 2006, 100 (2).
(16) Hartman, K.; Johnson, J. L.; Bertoni, M. I.; Recht, D.; Aziz, M.
J.; Scarpulla, M. A.; Buonassisi, T. SnS thin-films by RF sputtering at
room temperature. Thin Solid Films 2011, 519 (21), 7421−7424.
*E-mail: marc.steichen@uni.lu. (17) Sajeesh, T. H.; Warrier, A. R.; Kartha, C. S.; Vijayakumar, K. P.
Notes Optimization of parameters of chemical spray pyrolysis technique to
get n and p-type layers of SnS. Thin Solid Films 2010, 518 (15), 4370−
The authors declare no competing financial interest.


4374.
(18) Hibbert, T. G.; Mahon, M. F.; Molloy, K. C.; Price, L. S.; Parkin,
ACKNOWLEDGMENTS I. P. Deposition of tin sulfide thin films from novel, volatile
The present work was realized under the financial support of (fluoroalkythiolato)tin(IV) precursors. J. Mater. Chem. 2001, 11 (2),
469−473.
the Fonds National de la Recherche (F.N.R.) Luxembourg
(19) Price, L. S.; Parkin, I. P.; Hardy, A. M. E.; Clark, R. J. H.;
within the framework of the CORE Junior Track Project C11/ Hibbert, T. G.; Molloy, K. C. Atmospheric pressure chemical vapor
MS/1211521. The authors also thank the CRP Gabriel deposition of tin sulfides (SnS, Sn2S3, and SnS2) on glass. Chem. Mater.
Lippmann for use of SEM/EDX and XRD. Joffrey Didierjean 1999, 11 (7), 1792−1799.
is gratefully acknowledged for XPS measurements. Dr. Masato (20) Avellaneda, D.; Delgado, G.; Nair, M. T. S.; Nair, P. K.
Kurihara (TDK Corporation Japan) is acknowledged for solar Structural and chemical transformations in SnS thin films used in
cell finishing. Dr. Sebastian Fiechter (Helmholtz Zentrum chemically deposited photovoltaic cells. Thin Solid Films 2007, 515
Berlin) is gratefully acknowledged for providing tin sulfide (15), 5771−5776.
single-crystal materials. (21) Ghosh, B.; Das, M.; Banerjee, P.; Das, S. Fabrication and optical


properties of SnS thin films by SILAR method. Appl. Surf. Sci. 2008,
254 (20), 6436−6440.
REFERENCES (22) Sato, N.; Ichimura, M.; Arai, E.; Yamazaki, Y. Characterization
(1) Wadia, C.; Alivisatos, A. P.; Kammen, D. M. Materials availability of electrical properties and photosensitivity of SnS thin films prepared
expands the opportunity for large-scale photovoltaics deployment. by the electrochemical deposition method. Sol. Energy Mater. Sol. Cells
Environ. Sci. Technol. 2009, 43 (6), 2072−2077. 2005, 85 (2), 153−165.

4392 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393


The Journal of Physical Chemistry C Article

(23) Mathews, N. R.; Anaya, H. B. M.; Cortes-Jacome, M. A.; phosphonium ionic liquid. Phys. Chem. Chem. Phys. 2009, 11 (38),
Angeles-Chavez, C.; Toledo-Antonio, J. A. Tin sulfide thin films by 8532−8537.
pulse electrodeposition: structural, morphological, and optical proper- (43) Chen, Y. H.; Davoisne, C.; Tarascon, J. M.; Guery, C. Growth of
ties. J. Electrochem. Soc. 2010, 157 (3), H337−H341. single-crystal copper sulfide thin films via electrodeposition in ionic
(24) Kang, F.; Ichimura, M. Pulsed electrodeposition of oxygen-free liquid media for lithium ion batteries. J. Mater. Chem. 2012, 22 (12),
tin monosulfide thin films using lactic acid/sodium lactate buffered 5295−5299.
electrolytes. Thin Solid Films 2010, 519 (2), 725−728. (44) Lotter, E. www.diplot.de; Besigheim, Germany, 2009.
(25) Brownson, J. R. S.; Georges, C.; Levy-Clement, C. Synthesis of a (45) Guetay, L.; Redinger, A.; Djemour, R.; Siebentritt, S. Lone
delta-SnS polymorph by electrodeposition. Chem. Mater. 2006, 18 conduction band in Cu2ZnSnSe4. Appl. Phys. Lett. 2012, 100 (10), 4.
(26), 6397−6402. (46) Han, D. H.; Kim, B. S.; Choi, S. J.; Jung, Y. J.; Kwak, J.; Park, S.
(26) Mishra, K.; Rajeshwar, K.; Weiss, A.; Murley, M.; Engelken, R. M. Time-resolved in situ spectroelectrochemical study on reduction of
D.; Slayton, M.; McCloud, H. E. Electrodeposition and character- sulfur in N,N′-dimethylformamide. J. Electrochem. Soc. 2004, 151 (9),
ization of SnS thin-films. J. Electrochem. Soc. 1989, 136 (7), 1915− E283−E290.
1923. (47) Dhingra, S. S.; Kanatzidis, M. G. Syntheses and characterization
(27) Cheng, S.; Chen, G.; Chen, Y.; Huang, C. Effect of deposition of the 1st thallium polysulfide anions. Inorg. Chem. 1993, 32 (11),
potential and bath temperature on the electrodeposition of SnS film. 2298−2307.
Opt. Mater. 2006, 29 (4), 439−444. (48) Leong, T.-I.; Hsieh, Y.-T.; Sun, I. W. Electrochemistry of tin in
(28) Subramanian, B.; Sanjeeviraja, C.; Jayachandran, M.; the 1-ethyl-3-methylimidazolium dicyanamide room temperature ionic
Chockalingam, M. J. Electrochemical synthesis of SnS thin films for liquid. Electrochim. Acta 2011, 56 (11), 3941−3946.
photoelectrochemical cells. Des., Fabr. Charact. Photonic Dev. 1999, (49) Kammlott, G. W.; Franey, J. P.; Graedel, T. E. Atmospheric
3896, 474−482. sulfidation of copper-alloys. 2. Alloys with nickel and tin. J. Electrochem.
(29) Zainal, Z.; Hussein, M. Z.; Ghazali, A. Cathodic electro- Soc. 1984, 131 (3), 511−515.
deposition of SnS thin films from aqueous solution. Sol. Energy Mater. (50) Memming, R. Semiconductor Electrochemistry; Wiley-VCH Verlag
Sol. Cells 1996, 40 (4), 347−357. GmbH: Berlin, 2001.
(30) Gunasekaran, M.; Ichimura, M. Photovoltaic cells based on (51) Rodriguez, H.; Gurau, G.; Holbrey, J. D.; Rogers, R. D. Reaction
pulsed electrochemically deposited SnS and photochemically of elemental chalcogens with imidazolium acetates to yield imidazole-
deposited US and Cd1-xZnxS. Sol. Energy Mater. Sol. Cells 2007, 91 2-chalcogenones: direct evidence for ionic liquids as proto-carbenes.
(9), 774−778. Chem. Commun. 2011, 47 (11), 3222−3224.
(31) Malaquias, J.; Fernandes, P. A.; Salome, P. M. P.; da Cunha, A. (52) Belkin, S.; Wirsen, C. O.; Jannasch, H. W. Biological and
F. Assessment of the potential of tin sulphide thin films prepared by abiological sulfur reduction at high-temperatures. Appl. Environ.
sulphurization of metallic precursors as cell absorbers. Thin Solid Films Microbiol. 1985, 49 (5), 1057−1061.
2011, 519 (21), 7416−7420. (53) Meyer, B. Elemental sulfur. Chem. Rev. 1976, 76 (3), 367−388.
(32) Devika, M.; Reddy, N. K.; Gunasekhar, K. R. Structural, (54) Stirrup, B. N.; Hampson, N. A. Electrochemical reactions of tin
electrical, and optical properties of as-grown and heat treated ultra-thin in aqueous electrolytic solutions. Surf. Technol. 1977, 5 (6), 429−462.
SnS films. Thin Solid Films 2011, 520 (1), 628−632. (55) Ammar, I. A.; Darwish, S.; Khalil, M. W.; Eltaher, S. A review on
(33) Hibberd, C. J.; Chassaing, E.; Liu, W.; Mitzi, D. B.; Lincot, D.; the electrochemistry of tin. Mater. Chem. Phys. 1989, 21 (1), 1−47.
Tiwari, A. N. Non-vacuum methods for formation of Cu(In,Ga)(Se,S) (56) Johnson, J. B.; Jones, H.; Latham, B. S.; Parker, J. D.; Engelken,
(2) thin film photovoltaic absorbers. Prog. Photovoltaics 2010, 18 (6), R. D.; Barber, C. Optimization of photoconductivity in vacuum-
434−452. evaporated tin sulfide thin films. Semicond. Sci. Technol. 1999, 14 (6),
(34) Omoto, K.; Fathy, N.; Ichimura, M. Deposition of SnSxOy films 501−507.
by electrochemical deposition using three-step pulse and their
characterization. Jpn. J. Appl. Phys., Part 1 2006, 45 (3A), 1500−1505.
(35) Brownson, J. R. S.; Georges, C.; Larramona, G.; Jacob, A.;
Delatouche, B.; Levy-Clement, C. Chemistry of tin monosulfide
(delta-SnS) electrodeposition effects of pH and temperature with
tartaric acid. J. Electrochem. Soc. 2008, 155 (1), D40−D46.
(36) Steichen, M.; Dale, P. J. Synthesis of trigonal selenium nanorods
by electrodeposition from an ionic liquid at high temperature.
Electrochem. Commun. 2011, 13 (8), 865−868.
(37) Armand, M.; Endres, F.; MacFarlane, D. R.; Ohno, H.; Scrosati,
B. Ionic-liquid materials for the electrochemical challenges of the
future. Nat. Mater. 2009, 8 (8), 621−629.
(38) Endres, F.; Abbott, A. P.; MacFarlane, D. R. Electrodeposition
from Ionic Liquids; Wiley-VCH: Weinheim, 2008.
(39) Boros, E.; Earle, M. J.; Gilea, M. A.; Metlen, A.; Mudring, A.-V.;
Rieger, F.; Robertson, A. J.; Seddon, K. R.; Tomaszowska, A. A.;
Trusov, L.; Vyle, J. S. On the dissolution of non-metallic solid
elements (sulfur, selenium, tellurium and phosphorus) in ionic liquids.
Chem. Commun. 2010, 46 (5), 716−718.
(40) Manan, N. S. A.; Aldous, L.; Alias, Y.; Murray, P.; Yellowlees, L.
J.; Lagunas, M. C.; Hardacre, C. Electrochemistry of Sulfur and
Polysulfides in Ionic Liquids. J. Phys. Chem. B 2011, 115 (47), 13873−
13879.
(41) Dale, P. J.; Samantilleke, A. P.; Shivagan, D. D.; Peter, L. M.
Synthesis of cadmium and zinc semiconductor compounds from an
ionic liquid containing choline chloride and urea. Thin Solid Films
2007, 515 (15), 5751−5754.
(42) Izgorodin, A.; Winther-Jensen, O.; Winther-Jensen, B.;
MacFarlane, D. R. CdS thin-film electrodeposition from a

4393 dx.doi.org/10.1021/jp311552g | J. Phys. Chem. C 2013, 117, 4383−4393

You might also like