Al-SnS VIRTCON 2020

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Investigation on the diffusion of aluminium into SILAR grown SnS thin film using

Al/SnS heterostructure

Pushan Banerjee
Assistant Professor, Department of Physics,
Vidyasagar Metropolitan College,
39, Sankar Ghosh Lane & 8A Shibnarayan Das Lane,
Kolkata – 700006, INDIA
Phone: +913322419508
Email: pushan2007@gmail.com

Abstract
Tin (II) sulfide (SnS) is a potentially promising II-VI semiconductor material to be used as an absorber
layer in inorganic thin film heterojunction solar cells, replacing the materials containing hazardous cadmium or
scarce indium and gallium. Chemical methods of growth of tin sulfide thin films make the overall process of
solar cell fabrication relatively inexpensive. In this regard, successive ionic layer adsorption and reaction
(SILAR) method is a comparatively new chemical way to grow SnS thin films uniformly over any substrate. In
the present investigation, SnS thin films have been grown over aluminum coated glass substrates following
SILAR method. The resulting films have been vacuum annealed at 150 oC and 250 oC to promote the diffusion
of aluminium into SnS and structural improvement of the films. X-ray diffraction and FESEM have been used to
determine the shift in lattice constant and observe the surface morphology. Bandgap lowering has been noticed
as a result of annealing and is the indication of dominance of holes following negative Burstein-Moss shift.
Photoconductivity studies have revealed the enhancement of diffusion-assisted photocurrent at higher annealing
temperature.

Keywords: Tin (II) sulfide, SILAR, thin film, photoconductivity, Burstein-Moss shift.

1. Introduction
With the growing demand of electrical power worldwide, use of solar photovoltaic
power has become extremely important since the last few decades. Among different ways of
generation of solar photovoltaic power, silicon solar cell is the main shareholder. However,
the higher cost of crystalline silicon cells and lower efficiency of amorphous silicon based
ones have forced the research community to look towards thin film heterojunction
alternatives. The conventional inorganic thin film alternatives like gallium arsenide, cadmium
sulfide, cadmium telluride or copper indium diselenide based solar cells contain either toxic
cadmium and tellurium; or scarce gallium and indium. In this respect, tin (II) sulfide (SnS)
has emerged as a potential candidate for the absorber layer of thin film heterojunction solar
cells. SnS has some favourable features like high absorption coefficient in the visible range of
electromagnetic spectrum, non-toxic and abundant nature of tin and sulfur, and tunable direct
bandgap depending upon the method of fabrication.
Both physical and chemical methods of growth are available for synthesis of SnS thin
films, e,g, thermal evaporation of SnS at low vacuum [1], vacuum evaporation of high purity
SnS powder [2 – 4], cathodic electrodeposition [5], electrochemical deposition [6 - 8],
chemical bath deposition [9, 10], spray pyrolysis [11], etc. The characteristics of SnS thus
obtained depend very much on the process of synthesis. Physical means of growth results in
higher crystalline quality and purity while chemical means offer ease of fabrication and
possibility of doping. SnS has p-type conductivity suitable for application absorber layer in
thin film heterojunction solar cells, and the conductivity can be enhanced by doping with
metals like copper and aluminium. [12].
The crystal structure of the material is orthorhombic involving weak van der Waals
force binding the layers of tin and sulphur atoms. This layered structure is favourable for
synthesis of SnS films through successive ionic layer adsorption and reaction (SILAR)
process. Being a modification of chemical bath deposition method, the SILAR process
resembles the atomic layer deposition method of thin films. The cost effectiveness of SILAR
process is a result of the use of dilute aqueous precursor solutions for growth of film at room
temperature and normal pressure, leaving no toxic by-product at the end. Ristov et al [13]
reported the first fabrication of SILAR grown SnS thin films using a cold (18 – 25 oC)
anionic precursor solution and a hot cationic precursor (80 – 90 oC) solution. In the present
work, room temperature SILAR process has been employed for growth of SnS thin films over
aluminium coated glass substrates, followed by thermal diffusion of aluminium into SILAR-
grown SnS matrix to enhance the doping of the semiconductor.

2. Experimental Methodology
The process of sample preparation can be divided into three steps – (i) vacuum
evaporation of aluminium thin film onto glass substrate, (ii) growth of SnS thin films over
aluminium using SILAR process and (iii) thermal interdiffusion of aluminium into SnS to
achieve doping. In the first step, 5N pure aluminium wire was evaporated onto cleaned
sodalime glass substrates in a vacuum chamber at a base vacuum of 5×10-5 milibar. The film
thickness, about 25 nm, was measured using an in-situ digital thickness monitor. Next, SnS
thin films were grown over the aluminium coated glass substrates using aqueous solutions of
0.1M tin sulphate (SnSO4) as cationic precursor and 0.1M sodium sulfide (Na2S) as anionic
precursor at room temperature. After 150 cycles of alternate dipping of the substrates in the
above solutions followed by drying, the film thickness was determined gravimetrically and
was found to be around 200 nm. The details of the growth of SnS film employing SILAR
process can be found elsewhere [14, 15]. After the growth of SnS films, one set of films was
kept without heat treatment and was named “as-grown” film. Two other sets of the Al/SnS
heterostructure were annealed at 150 oC and 250 oC for one hour in a vacuum level of 2×10-5
milibar to promote interdiffusion of aluminium into SnS.
In the next part of the work, the three sets of films were analyzed for crystallinity
using normal X-ray diffraction process employing a Brukers D8 Advance X-ray
diffractometer with Cu Kα radiation (λ = 1.5406 Å) in the range of Bragg angles 10° ≤ 2θ ≤
70° at a scanning rate of 1° min− 1. The morphology of the films was studied using a Quanta
PEC scanning electron microscope. The optical bandgap of the samples were determined
using the optical transmission characteristics of the films, measured with a Filmetrics thin
film analyzer in the wavelength of 400 – 1100 nm. Finally, the photoconductivity of the
samples were studied by recording the current across two circular contact pads of silver,
deposited 8 mm apart over the films, applying a bias of 30V across them. At first the samples
were illuminated using light from a 1000W tungsten-halogen lamp for 120 second to observe
the growth and saturation of photocurrent. Then the light was kept switched off for 130
second to observe the decay of photocurrent. The current was recorded using a Keithley 238
high current source and Interactive Characterization Software.

3. Results and Discussion


3.1 Structural characterization
Figure 1shows the stacked X-ray diffraction pattern of the three sets of films. The
prominent peaks appeared in the pattern correspond to (021), (040) and (130) sets of
crystalline planes. This is in conformity with the earlier studies [15] regarding the growth of
SnS films using SILAR method. The peaks were matched with JCPDS card no. 83-1758 and
the full width at half maximum (FWHM) for the peaks were calculated. Using the peak
positions and the FWHM values of the three peaks, the average crystallite size and lattice
microstrain were calculated using the relation [16]:

(1)

where  is the FWHM in radian,  is the Bragg angle for the peak, λ is the wavelength of
radiation used, L is the average crystallite size and ϵ is the microstrain. The calculated values
of L and ϵ have been shown in Table-1.
Fig. 1: X-ray diffraction pattern of the films.

Table-1: Crystallite size and microstrain of films


Sample Average crystallite size (nm) Microstrain
As-grown 11.22 - 0.0285
Annealed at 150 oC 15.19 - 0.0165
Annealed at 250 oC 20.62 - 0.0112

It can be viewed from Table-1 that as a result of annealing, the crystallite size
gradually increased and the negative microstrain reduced, i.e. the compressive strain inside
the system was relaxed. This is a very obvious result of annealing, because the thermal
energy helped the atoms to reorient themselves.
Figures 2a and 2b show respectively the surface morphology of the as-grown film and
that annealed at 250 oC. The surface of the film appears to be free of pinholes. The as-grown
film had a comparatively rough surface, but annealing resulted to a smoother surface
morphology due to reorientation of the grains. The grain size was found to be a few tens of
nanometers, matching with those found from X-ray diffraction pattern.
Fig. 2a: Morphology of the as-grown film.

Fig. 2b: Morphology of the film annealed at 250 oC.

3.2 Optical characterization


Figure 3a depicts the transmission characteristics of the three sets of films. The set
annealed at 250 oC has the highest transmission while the as-grown film had the lowest. This
was because of the undiffused aluminium layer in the later, preventing transmission through
itself. Diffusion of aluminium into SnS resulted in bond formation between the metal and the
semiconductor by reducing the dangling bonds and enhancing the transmission.
Figure 3b shows the Tauc plot for the films, from which the direct optical bandgap of
the resulting semiconductor has been estimated using the relation [17]
(h)2 = h – Eg (2)
and taking the intercept of the tangent to the plot over h axis.

Fig. 3a: Transmission spectra of the films.

Fig. 3b: Bandgap determination of the films.


It can be seen that the estimated optical bandgap of the films ranged from 1.70 eV for
as-grown film to 2.00 eV for the highest temperature annealed film. These values of optical
bandgap are higher than that (~ 1.4 eV) for crystalline SnS, presumably due to the
nanocrystalline nature of the grains. Also, the blue shift of optical bandgap is due to the
Burstein-Moss effect [18], where in p-type semiconductor, doping makes the material
degenerate and Fermi level is shifted into the valence band. As a result, the optical bandgap
of the material is increased. Thus, diffusion of aluminium has caused p-type doping into SnS.

3.3 Photoconductivity measurement


Figure 4 shows the plot of normalized photocurrent (i.e. photocurrent w.r.t. the dark
current in individual samples at the given bias) against time. As explained in section 3.2, the
sample having the highest doping i.e. annealed at 250 oC showed the highest normalized
photocurrent and the highest rate of decay of photocurrent with time.

Fig. 4: Photoresponse of the films.


These features can be attributed to the filling up of dangling bonds and defect states in
SnS by aluminium atoms as a result of thermal diffusion of the metal. Initially, the as-grown
SnS film contained numerous defect states, acting as recombination centres. So, the
photocurrent generated as a result of illumination was comparatively lower due to the
trapping of photogenerated electrons and holes. As aluminium atoms reduced the number of
trapping sites, the photocurrent was found to be enhanced.

4. Concluding Remarks
In this study, the diffusion of aluminium into SILAR grown SnS films and consequent
change in its p-type conductivity were studied. SILAR method has been employed here as a
comparatively new way of atomic layer deposition process of SnS films by chemical means.
With change in annealing temperature of the Al/SnS bilayer structure, the size of the
nanocrystals slightly enhanced combined with the reduction of overall compressive
microstrain. The crystallites gradually rearranged themselves at higher annealing
temperatures. The diffusion of aluminium into SnS resulted in the occupation of vacant SnS
lattice sites by the aluminium atoms, thus reducing the defect states and lowering the trapping
of charges. As a result, the photogenerated current reaching the electrical contacts got
enhanced in the structure. At the same time, aluminium being a group-III element, acted as a
supplier of holes in the IV-VI semiconductor and more holes from the top of the valence band
occupied the states below the semiconductor Fermi level, so the Fermi level moved more
towards the valance band. Now only the electrons from deeper levels in the valence band
could reach the conduction band, demanding higher optical excitation energies, resulting in a
blue shift in the optical bandgap. Thus, annealing the Al/SnS bilayer at higher temperature
i.e. 250 oC incorporated more holes into the structure through the diffusion of aluminium and
enhanced the p-type doping of SnS. Hall measurement to determine and compare the
individual carrier concentration in the samples has not been possible in the present case,
owing to the chance of short-circuiting the top contacts with the bottom metal layer. This
possibility of short-circuit did not affect the photoconductivity, because the metal layer would
not contribute to the photogenerated carriers.

Acknowledgements
The author would like to thank Dr. Subrata Das, School of Energy Studies, Jadavpur
University for his inspiration and constant mentoring in this work. He would thank Prof.
Biswajit Ghosh, School of Energy Studies, Jadavpur University for providing infrastructure
for this work. The help received from Dr. Mukul Gupta of UGC-DAE CSR, Indore towards
XRD measure measurements and Dr. J. B. M. Krishna of UGC-DAE CSR, Kolkata towards
UV-Visible spectrophotometry is also thankfully acknowledged.

References
1. Shama A. A., Zeyada H.M., 2003, ‘Electronic dielectric constants of thermally
evaporated SnS thin films’, Optical Materials, 24 (3), 555 – 561.
2. Johnson J.B., Jones H., Latham B.S., Parker J.D., Engelken R., Barber C., 1999,
‘Optimization of photoconductivity in vacuum-evaporated tin sulfide thin films’,
Semiconductor Science and Technology, 14 (6) 501 – 507.
3. Devika M., Reddy N. K., Ramesh K., Ganesan V., Gopal E.S.R., Reddy K.T. R.,
2006, ‘Influence of substrate temperature on surface structure and electrical resistivity
of evaporated tin sulfide films’, Applied Surface Science, 253 (3), 1673 – 1676.
4. Ghosh B., Bhattacharjee R., Banerjee P., Das S., 2011, ‘Structural and optoelectronic
properties of vacuum evaporated SnS thin films annealed in argon ambient’, Applied
Surface Science, 257 (8), 3670 – 3676.
5. Ghazali A., Zainal Z., Hussein M.Z., Kassim A., 1998, ‘Cathodic electrodeposition of
SnS in the presence of EDTA in aqueous media’, Solar Energy Materials and Solar
Cells, 55 (3), 237 – 249.
6. Ichimura M., Takeuchi L., Ono Y., Arai E., 2000, ‘Electrochemical deposition of SnS
thin films, Thin Solid Films, 361–362, 98 – 101.
7. Cheng S., Chen G., Chen Y., Huang C., 2006, ‘Effect of deposition potential and bath
temperature on the electrodeposition of SnS film’, Optical Materials, 29 (4), 439 –
444.
8. Ghosh B., Roy R., Chowdhury S., Banerjee P., Das S., 2010, ‘Synthesis of SnS thin
films via galvanostatic electrodeposition and fabrication of CdS/SnS heterostructure
for photovoltaic applications’, Applied Surface Science, 256 (13), 4328 – 4333.
9. Avellaneda D., Delgado G., Nair M.T.S, Nair P.K., 2007, ‘Structural and chemical
transformations in SnS thin films used in chemically deposited photovoltaic cells’,
Thin Solid Films, 515 (15), 5771 – 5776.
10. Tanusevski A., 2003, ‘Optical and photoelectric properties of SnS thin films prepared
by chemical bath deposition’, Semiconductor Science and Technology, 18 (6), 501 –
505.
11. Reddy N. K., Reddy K.T. R., 2007, ‘Preparation and characterisation of sprayed tin
sulfide films grown at different precursor concentrations’, Materials Chemistry and
Physics, 102 (1), 13 – 18.
12. Ristov M., Sinadinovski Gj., Grozdanov I., Mitreski M., 1989, ‘Chemical deposition
of thin (II) sulphide thin films’, Thin Solid Films, 173 (1), 53 – 58.
13. Ristova M., Ristov M., 1998, ‘Silver doped CdS films for PV applications’, Solar
Energy Materials and Solar Cells, 53 (1-2), 95 – 102.
14. Ghosh B., Das M., Banerjee P., Das S, 2008, ‘Fabrication and optical properties of
SnS thin films by SILAR method’, Applied Surface Science, 254 (20), 6436 – 6440.
15. Ghosh B., Das M., Banerjee P., Das S, 2008, ‘Fabrication of SnS thin films by the
successive ionic layer adsorption and reaction method’, Semiconductor Science and
Technology, 23 (12), 125013.
16. Quadri S. B., Skelton E. F., Dinsmore A. D., Yang J., Gray H. F., Ratna B. R., 1999,
‘Size induced transition temperature reduction in nanoparticles of ZnS’, Physical
Review B, 60 (13), 9191 – 9193.
17. Bhatt R., Bhaumik I., Ganesamoorthy S., Karnal A.K., Swami M.K., Patel H.S.,
Gupta P.K., 2012, ‘Urbach tail and bandgap analysis in near stoichiometric
LiNbO3 crystals’, Physica Status Solidi A, 209 (1), 176 – 180.
18. Hassan A., Jin Y., Irfan M., Jiang Y., 2018, ‘Acceptor modulated optical
enhancements and bandgap narrowing in ZnO thin films’, AIP Advances, 8, 035212.

You might also like