You are on page 1of 11

Superlattices and Microstructures 55 (2013) 180–190

Contents lists available at SciVerse ScienceDirect

Superlattices and Microstructures


journal homepage: www.elsevier.com/locate/superlattices

Influence of Sn doping level on antibacterial


activity and certain physical properties of ZnO
films deposited using a simplified spray pyrolysis
technique
M. Vasanthi a, K. Ravichandran a,⇑, N. Jabena Begum a, G. Muruganantham a,
S. Snega a, A. Panneerselvam b, P. Kavitha b
a
P.G. & Research Department of Physics, AVVM Sri Pushpam College (Autonomous), Poondi, Thanjavur 613 503, Tamil Nadu, India
b
P.G. & Research Department of Botany and Microbiology, AVVM Sri Pushpam College (Autonomous), Poondi, Thanjavur 613
503, Tamil Nadu, India

a r t i c l e i n f o a b s t r a c t

Article history: Nanocrystalline tin–doped zinc oxide (ZnO:Sn) films with different
Received 10 November 2012 Sn doping levels (0, 2, 4, . . . , 10 at.%) were fabricated using a simpli-
Received in revised form 7 December 2012 fied spray pyrolysis technique. All the deposited films were charac-
Accepted 14 December 2012
terized in order to explore the influence of Sn doping level on
Available online 28 December 2012
antibacterial and certain physical properties. The XRD studies
revealed that all the films exhibited preferential orientation along
Keywords:
ZnO thin films
the (0 0 2) plane irrespective of the Sn doping level. The electrical
Spray pyrolysis sheet resistance (Rsh) sharply decreases with the increase in the
Electrical properties Sn doping level and attain a minimum value (3.88  102 X/h) at
Antibacterial activity 6 at.% and then increases for further doping. The reason for this
Opto-electronic devices observed variation in the Rsh value is discussed in detail. The opti-
SEM cal studies showed that all the films exhibited good transparency
(85%) in the visible region. The obtained photoluminescence
(PL) spectra endorsed the good crystalline quality of the films
and enhancement of the optical band gap (Eg) caused by Sn doping.
From the SEM images, it is inferred that the incorporation of Sn has
the tendency of repairing the porous structure of ZnO films. The
antibacterial activity of ZnO:Sn films was found to be enhanced
with the increase in Sn incorporation into the ZnO lattice.
Ó 2013 Elsevier Ltd. All rights reserved.

⇑ Corresponding author. Tel.: +91 4362 278602, mobile: +91 9443524180; fax: +91 4374 239328.
E-mail address: kkr1365@yahoo.com (K. Ravichandran).

0749-6036/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.spmi.2012.12.011
M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190 181

1. Introduction

Transparent conductive oxide (TCO) films such as tin oxide (SnO2), zinc oxide (ZnO), titanium diox-
ide (TiO2) and indium oxide (In2O3) have significant and potential role in the fabrication of next gen-
eration of optoelectronic devices in the UV region and display devices. Among these TCOs, ZnO based
thin films have recently gained much attention owing to its unique advantages over other oxide thin
films. It is a direct, wide band gap (Eg) semiconductor (Eg  3.37 eV) having hexagonal wurtzite struc-
ture with a large exciton binding energy (60 meV) [1]. Due to its high optical transparency, high elec-
trical conductivity and high stability in the presence of hydrogen plasma, it has fascinated for many
electronic and optoelectronic applications including sensors [2], piezo field emitters [3], transparent
electrodes [4], field emission devices [5], light emitting diodes [6] and electroluminescent devices
[7]. Furthermore, doped ZnO films significantly improve the efficiency of the above said devices.
The introduction of novel powerful antimicrobial agents is of great importance for the control of
pathogenic bacteria. Currently, the antibacterial agents can be classified into two categories; organic
and inorganic reagents. Inorganic antibacterial agents are more stable at high temperatures and pres-
sures compared with the organic materials [8]. ZnO is one of the significant inorganic metal oxides
exhibiting excellent antibacterial activities [9]. The efficient antibacterial property of ZnO nanostruc-
ture make it a suitable candidate for its possible application in the food preservation and packaging
systems, variety of medical and skin coatings, water purification, bio-imaging and drug delivery. Re-
cently, it is well established that the doping process improves the antibacterial efficacy of ZnO. Espe-
cially, the dopants like Sn, Mn, Mg, Sb, Co and N play a vital role in enhancing the antimicrobial
activities of zinc oxide [10–14].
Therefore, in the present study, tin–doped zinc oxide (ZnO:Sn) thin films with different Sn doping
levels (0, 2, 4, . . . , 10 at.%) have been fabricated using spray pyrolysis technique and their certain phys-
ical properties along with the antibacterial activity have been explored in order to understand the
influence of tin doping level. The reason for choosing spray pyrolysis technique is that, it is simple,
inexpensive and has flexible process modifications for large area TCO coatings. Moreover, it has many
advantages over other preparation techniques such as solvent casting method [15], pulsed laser depo-
sition [16], chemical bath deposition (CBD) [17], sol–gel [18], salvo-thermal process [19] and rf-mag-
netron sputtering [20].

2. Experimental details

2.1. Preparation of ZnO:Sn films

Sn doped zinc oxide films were deposited onto the glass substrates with different Sn doping levels
(0, 2, 4, . . . , 10 at.%) by employing spray pyrolysis technique using perfume atomizer. In this perfume
atomizer, the atomization was achieved based on hydraulic pressure without using any carrier gas.
The starting solution was prepared by dissolving zinc acetate dihydrate (Zn(CH3COO)22H2O)
(0.1 M) salt in a mixture of deionized water, methanol and acetic acid taken in the ratio of 6:3:1
respectively. Tin doping was achieved by adding different proportions (2, 4, . . . , 10 at.%) of tin (II) chlo-
ride dihydrate (SnCl22H2O) with the starting solution. Temperature of the substrate was maintained
at 340 ± 5 °C using a temperature controller with chromel–alumel thermocouple. The intermittent
spray deposition followed in this study involves a spray and 10 s interval. The substrates were cleaned
ultrasonically with organic solvents before starting the deposition process in order to improve the
adhesive nature of the film.

2.2. Film characterization

The thickness of the samples was measured using stylus technique (Profilometer: Surf Test SJ-301)
and the obtained values are presented in Table 1. The crystal structure of the films was analyzed using
X-ray powder diffraction (PANalytical-PW 340/60 X0 pert PRO) technique with CuKa (k  1.5406 Å) as
182 M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190

Table 1
Structural, electrical and optical parameters of Sn-doped ZnO films.

Sn doping Thickness Crystallite Lattice Texture co-efficient Resistivity Optical band


level (at.%) (nm) size, D (nm) parameters *
of (0 0 2) plane q  102 (X cm) gap, Eg (eV)
(Å)
a c
0 610 33.97 3.208 5.203 2.810 11.895 3.18
2 620 42.29 3.257 5.199 1.883 8.692 3.24
4 600 21.64 3.354 5.198 1.229 6.576 3.25
6 620 16.91 3.265 5.195 1.186 2.399 3.28
8 590 21.15 3.302 5.199 1.650 3.345 3.26
10 580 26.68 3.276 5.201 1.412 4.240 3.23
*
Standard values a = 3.249 Å; c = 5.205 Å (JCPDS card No. 36-1451).

X-ray source. The electrical studies of the films were carried out using four point probe method. The
optical transmittance spectra were recorded using Perkin Elmer double-beam spectrophotometer
(LAMBDA 35 model) in the UV–Vis NIR region (300–1100 nm). The surface morphological and the
photoluminescence studies were made using scanning electron microscope (SEM) (JEOL-JSM 6390
with attachment INCA-Penta FETX3 OXFORD) and the spectro-flurometer (Jobin Yvon_FLUROLOG-
FL3-11) with Xenon Lamp (450 W), respectively.

2.3. Evaluation of antibacterial activity

Antibacterial activity of ZnO:Sn films against Escherichia coli (E. coli) was investigated using liquid
culture test. For this test, to prepare the medium, 20 mL of nutrient broth was inoculated with 200 lL
of E. coli. Then the samples (square in shape with the dimension of 1 cm  1 cm) were immersed in the
medium and then placed in orbital shaker at 200 rpm kept at a temperature of 37 °C. The same pro-
cedure was followed for the control plain glass plate. The optical density at 600 nm (OD600 nm) of this
medium (the bacteria and sample) was measured using Digital Spectrophotometer 169.

3. Results and discussion

3.1. Structural studies

Fig. 1 shows the X-ray diffraction (XRD) patterns of the pristine and Sn doped ZnO films. The XRD
profiles revealed that the growth of the films is mainly along the c axis with hexagonal wurtzite struc-
ture. All the doped ZnO films have preferential orientation along the same (0 0 2) plane as that of the
pristine film irrespective of the Sn doping level, indicating that the incorporation of Sn into the Zn sites
does not alter the preferential growth. Other weak peaks present in the patterns are associated with
the (1 0 0), (1 0 1) and (2 0 1) planes of ZnO structure. No peaks related to tin oxide are observed in the
patterns, indicating the purity of the ZnO film.
Eventhough the preferential orientation is remained as (0 0 2) plane for doped films also, the degree
of preferential growth gradually declined with the increase in the Sn doping level as observed from the
intensity of the (0 0 2) plane.
The texture co-efficient (TC) of the undoped and doped ZnO films corresponding to the (0 0 2) plane
is calculated using the formula [21].

Ið0 0 2Þ=I0 ð0 0 2Þ
TC ð0 0 2Þ ¼ 1 PN
ð1Þ
N Iðh k lÞ=I0 ðh k lÞ

where I(h k l) and I0(h k l) are the measured relative intensity and standard intensity (JCPDS: 36-1451)
of the (h k l) plane respectively and I(0 0 2) and I0(0 0 2) are the corresponding values of the (0 0 2) plane
and N is the number of diffraction peaks. The calculated TC values of (0 0 2) plane are tabulated in Ta-
ble 1 and the variation in the texture co-efficient as a function of Sn doping level is presented in Fig. 2.
M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190 183

Fig. 1. X-ray diffraction patterns of Sn-doped ZnO films.

As seen from Fig. 2, the texture co-efficient falls gradually as the doping level increases, suggesting a
monotonical degradation in the crystalline quality due to Sn doping. The X-ray diffraction patterns
also show that the (0 0 2) peak gets widened along with gradual decrease in intensity, indicating a deg-
radation in crystalline quality. The crystallite size (D) is calculated from the Scherrer’s formula [22]

0:9 k
D¼ ð2Þ
b cos h

where k is the wavelength of the X-ray used (1.5406 Å), b is the full width at half maximum and h is
the Bragg’s angle. The calculated crystallite size of the films is found to be in the nanorange and their
corresponding lattice parameters (‘a’ and ‘c’) are presented in Table 1.
A slight right shift in the 2h value (Fig. 1) and decrement in the lattice parameter ‘c’ value (Table 1)
for the doped ZnO up to 6 at.% may be attributed to the incorporation of Sn4+ ions into the Zn2+ sites.

Fig. 2. Variation in the texture co-efficient (TC) of (0 0 2) plane of ZnO:Sn films as a function of Sn doping level.
184 M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190

This inference is arrived, because the ionic radius of Sn4+ (0.069 nm) is lesser than that of Zn2+
(0.074 nm) and hence the substitutional incorporation of Sn4+ into the Zn2+ sites could cause a slight
decrease in the interplanar distance (d). Beyond 6 at.% of Sn doping, a slight left shift in the 2h value
and increment in the ‘c’ value (Table 1) are obtained which may be due to the higher ionic radius of
Sn2+ ions (0.093 nm) generated at this critical doping level as discussed in detail in the forthcoming
electrical studies (Section 3.2).

3.2. Electrical properties

The variation in electrical sheet resistance (Rsh) of the ZnO:Sn films as a function of Sn doping level
in the starting solution is shown in Fig. 3 and the corresponding resistivity (q) values are presented in
Table 1. The Rsh decreases sharply as the doping level increases, attaining a minimum value
(3.88  102 O/h) at 6 at.% and then increases for further doping. This variation in Rsh can be explained
as follows: The initial decrease in Rsh is caused by the substitution of Sn4+ ions into the Zn2+ sites, be-
cause each of this substitution results in the creation of two free electrons. In other words, the substi-
tuted Sn4+ ions act as donors to the system thus reducing the Rsh. This trend is valid up to certain
doping level only (in this study, it is 6 at.%) and for doping levels beyond this critical value, the sce-
nario is different. Beyond the critical doping level, the Sn4+ ions are reduced to Sn2+ state by absorbing
a pair of electrons already available in the system, there by causes a reduction in the number of free
carriers which in turn increases the Rsh. This inference on the obtained electrical results is derived on
the basis of one of the important characteristics features of the p-block elements as discussed below:
the p-block elements of the periodic table having higher atomic number, have lesser tendency to use s
electrons for chemical bonding and hence they are stable only at the lower oxidation states. Therefore
Sn being an element of this kind is more stable in its +2 oxidation state rather than +4 states. Hence,
we strongly believed that the Sn4+ ions are reduced to the Sn2+ state by absorbing a pair of electrons in
the vicinity beyond a critical doping level. It is reasonable to assume that the reduction process can
start only when the number of free electrons available in the system is sufficiently high to initiate this
phenomena and this stage arrives at the doping level of 6 at.% (which we named as critical doping le-
vel) for the deposition conditions employed in the present study. The above discussion on the critical
doping level is corroborated well with the X-ray diffraction studies (Section 3.1) and optical studies
(Section 3.3).
Similar discussions are reported by Kojima et al. [23] for Sb doping in SnO2 films. They established
clearly that Sb5+ ions were reduced to Sb3+ ions acting as acceptors beyond the certain doping level.

Fig. 3. Variation in the electrical sheet resistance (Rsh) of ZnO:Sn films as a function of Sn doping level.
M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190 185

Bougrine et al. [24] reported a mechanism of Sn incorporation into ZnO lattice in which they ex-
plained that after the critical doping level, the Sn atoms cannot occupy lattice sites of Zn and they have
a tendency to occupy interstitial sites and formed as neutral defects becoming ineffective as dopant
impurity. Our argument is that if it was the case, then there should not be any increase in the Rsh when
the doping level increases beyond the critical doping level and the Rsh should have the same value as it
had at the critical doping level. But, they observed an increase in the Rsh beyond the critical doping
level. From the experimental results and the characteristics of Sn, it is clearly evident that the mech-
anism suggested by us satisfactorily explains the behaviour of the incorporated Sn in the ZnO lattice as
far as the electrical property is concerned.
Vaezi et al. [25] reported another possible mechanism for this raise in the Rsh. They explained that
at higher doping concentrations, more Sn atoms are incorporated into the Zn sites which cause resid-
ual stress and this stress field would reflect electric carriers resulting in an increase in Rsh. It appears to
be another reasonable mechanism related to the variation in the Rsh.

3.3. Optical studies

Transmittance spectra recorded for undoped and Sn-doped ZnO films in the wavelength range from
300 to 1100 nm are shown in Fig. 4. From the spectra, it can be noted that all the films exhibit good
transparency (85%) in the visible region along with sharp fundamental absorption edge suitable for
optoelectronic applications especially for solar cell applications. But, the ZnO films doped with 2–
6 at.% of Sn show a blue shift in the absorption edge whereas for 8 and 10 at.% of Sn doping, a red shift
is observed. The inset of Fig. 4 clearly depicts this shifting in the absorption edge. These shifts may be
attributed to the variation in the carrier concentration and thus these shifts confirmed the proposed
mechanism related to the variation in the Rsh as discussed in the electrical studies (Section 3.2). This
discussion is further supported by the variation in the optical band gap (Eg) values of the ZnO:Sn films.
The Eg values are estimated by plotting the graph between the first derivative of transmittance with
respect to the wavelength (dT/dk) and the average wavelength (kavg) (Fig. 5)[26]. The Eg value of the
pristine ZnO film is 3.18 eV (Table 1). The Eg value increases gradually as the Sn doping level increases
and attains a maximum value of 3.28 eV at the critical level (6 at.%) of Sn doping. Beyond this critical
doping level, the band gap again decreases as shown in Fig. 6. This variation in Eg is associated with the
Moss–Burstein effect [1].

Fig. 4. Transmittance spectra of Sn-doped ZnO films and the widened portion of absorption edge (inset).
186 M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190

Fig. 5. The (dT/dk) vs kavg plots of ZnO:Sn films.

Fig. 6. Variation in the optical band gap of Sn-doped ZnO films as a function of Sn doping level.

3.4. Photoluminescence (PL) studies

Fig. 7 shows the room temperature photoluminescence (PL) spectra of the undoped and Sn-doped
ZnO thin films. All the samples exhibit two distinct emission peaks: a sharp UV emission band near
390 nm and another band at 575 nm attributed to green emission in the visible region. It is well
known that the UV band corresponds to the ZnO band edge transition and the green emission band
is related to defects density [27]. The obtained less intense green emission for all the samples indicates
the lesser defect concentration of these films. The shifts in the UV peak towards lower wavelength side
(blue shift) along with increased intensity up to 6 at.% of Sn doping level evince the enhancement of
radiative recombination centres due to the substitution of Sn4+ ions into Zn2+ sites as discussed in elec-
trical studies (Section 3.2). This blue shift in PL spectra corroborates well with the shift of absorption
edge observed in the optical transmittance spectra. At the same time, the reduction in UV intensity for
M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190 187

Fig. 7. Room temperature photoluminescence spectra of Sn-doped ZnO films.

Fig. 8. SEM images of ZnO films with Sn doping level of (a) 0 at.%, (b) 2 at.%, (c) 6 at.%, (d) 10 at.%.

the ZnO films doped with 8 and 10 at.% of Sn is attributed to the introduction of defects responsible for
non-radiative transition.

3.5. Surface morphology

The surface morphology of undoped and Sn-doped zinc oxide films contemplated with SEM is
shown in Fig. 8a–d. From Fig. 8, it can be clearly seen that, the undoped ZnO film (Fig. 8a) exhibits
188 M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190

Fig. 9. Growth analysis of E. coli against Sn doping level in ZnO films by monitoring the optical density (OD) at 600 nm.

the feature of hexagonal form grains with some micro-pores in the matrix. This micro-pore structure
is seemed to be repaired by Sn doping as shown in Fig. 8(b–d). This result is expected, as the introduc-
tion of compressed stresses due to the difference in ionic radii between Zn2+ and Sn4+/Sn2+ ions can
cause a reduction in the porosity. A similar behaviour was observed by Kuan et al. [28] and Bougrine
et al. [24] for Sn-doped ZnO films prepared by the sol–gel and spray pyrolysis methods respectively. In
their study, Kuan et al. [28] reported that, the metal dopants not only suppressed the crystallization,
but also decreases the porosity of the film. The inhibition of crystallization in the present study is con-
firmed by the XRD profiles of Sn-doped ZnO thin films (Section 3.1). In addition, due to this decrease in
porosity of the films, the microstructure of the film became denser with the increase in Sn doping. A
small but systematic variation in the size and shape of the grains has also been observed with the in-
crease in Sn concentration in the ZnO films.

3.6. Antibacterial activity

Antibacterial activity of ZnO:Sn films with different doping levels of Sn (0, 2, . . . , 10 at.%) was ex-
plored using liquid culture test. The optical density (OD) of each sample was measured at 600 nm
to analyse the antibacterial activity against E. coli and the obtained values are plotted against Sn dop-
ing levels (Fig. 9). From Fig. 9, it is inferred that, the antibacterial activity is enhanced with the increase
in Sn doping level.
According to the mechanism related to the antibacterial activity of ZnO material, it is reported that,
the released Zn2+ ions from ZnO lattice [29], on coming in contact with bacterial cells, can produce
reactive oxygen species (ROS) such as hydroxyl radicals, superoxides and H2O2 [30,31]. The H2O2 di-
rectly penetrate the cell membrane and kill the bacteria, whereas, the negatively charged hydroxyl
radicals and superoxides cannot penetrate into the cell membrane and remain in direct contact with
the outer surface of bacteria causing severe damage to proteins, lipids and DNA [32]. Hence, in the
present work, the enhancement in the antibacterial activity with the increase in Sn doping level
may be due to the increase in the liberation of Zn2+ from the film which plays a major role in the above
said mechanism of antibacterial activity. When the Sn concentration increases, more Zn2+ sites of ZnO
matrix are occupied by Sn4+/Sn2+ ions as discussed in electrical studies (Section 3.2) and hence it is
reasonable to expect an increased number of Zn2+ ions on the interstitial positions of ZnO matrix.
As Zn2+ ions in the interstitial positions may be easily released from the lattice rather than that in na-
tive positions, they can participate effectively in the antibacterial activity. An interesting result ob-
served in this study is the remarkable steep decrease in the optical density (Fig. 9) beyond the
M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190 189

doping level of 6 at.% indicates a phenomenal enhancement in the antibacterial activity. The reason for
this sudden increase in the antibacterial activity can be understood on the basis of the following facts:
As already discussed in the electrical studies (Section 3.2), beyond 6 at.% of Sn doping level, Sn4+ ions
are reduced into Sn2+ state. Since these Sn2+ ions have higher ionic radius (0.093 Å) than Sn4+ ions
(0.069 Å), there could be a reduction in the voids spaces of the interstitial positions which induces
the release of more number of Zn2+ ions from the ZnO lattice.
In the study on size dependent bacterial growth inhibition and mechanism of antibacterial activity
of zinc oxide nanoparticles, Raghupathi et al. [33] observed an increase in antibacterial activity with a
decrease in particle size. Hence, in our study, the decrease in particle size with the increase in Sn con-
centration as observed from SEM images (Fig. 8) may also be another reason for the improved antibac-
terial activity at higher Sn doping levels.

4. Conclusion

Good quality nanocrystalline undoped and Sn-doped ZnO films were deposited using spray pyro-
lysis technique and the effects of tin doping level on certain physical properties and antibacterial
activity were studied. The structural studies revealed that all the films exhibit hexagonal wurtzite
structure with preferential orientation along the (0 0 2) plane. The observed variation in the electrical
sheet resistance (Rsh) induced by different concentration of Sn, was clearly explained on the basis of
the important characteristic features of p-block elements. The blue shift in the PL spectra corroborated
well with the shift in the absorption edge of the optical transmittance spectra. From the SEM images, it
was realized that the concentration of Sn largely governed the surface morphology of the zinc oxide
films. From the above results it is concluded that the tin acts as an active dopant element to enhance
the opto-electrical characteristic of zinc oxide films which have potential application as dynamic
transparent oxide layer for solar cell applications. Further, it is found that the antibacterial activity
of ZnO:Sn film is well pronounced for higher Sn doping.

Acknowledgements

Financial support from the University Grants Commission of India through the Major Research Pro-
ject (F. No. 40-28/2011(SR)) is gratefully acknowledged.

References

[1] K. Saravanakumar, K. Ravichandran, J. Mater. Sci. Mater. Electron. 23 (2012) 1462–1469.


[2] Ian Y.Y. Bu, Chih-Chin Yang, Superlattices Microstruct. 51 (2012) 745–753.
[3] T. Prasada Rao, M.C. Santhosh Kumar, A. Safarulla, V. Ganesan, S.R. Barman, C. Sanjeeviraja, Physica B 405 (2010) 2226–
2231.
[4] V. Shelke, B.K. Sonawane, M.P. Bhole, D.S. Patil, J. Mater. Sci. Mater. Electron. 23 (2012) 451–456.
[5] Sen Yi, Li, Pang Lin, Chia Ying Lee, Tseung Yuen Tseng, Chorng Jye Huang, J. Phys. D: Appl. Phys. 37 (2004) 2274–2282.
[6] Yoon-Duk Ko, Ki-Chul Kim, Young-Sung Kim, Superlattices Microstruct. 51 (2012) 933–941.
[7] Y.L. Wu, A.I.Y. Tok, F.Y.C. Boey, X.T. Zeng, X.H. Zhang, Appl. Surf. Sci. 253 (2007) 5473–5479.
[8] J. Sawai, J. Microbiol. Methods 54 (2003) 177–182.
[9] L.-H. Li, J.-C. Deng, H.-R. Deng, Z.-L. Liu, X.-L. Li, Chem. Eng. J. 160 (2010) 378–382.
[10] X. Zhang, R. Zhou, P. Liu, L. Fu, X. Lan, G. Gong, Int. J. Appl. Ceram. Technol. 8 (5) (2011) 1087–1098.
[11] B.H. Soni, M.P. Deshpande, S.V. Bhatt, S.H. Chaki, H. Kaheria, Arch. Appl. Sci. Res. 3 (6) (2011) 173–179.
[12] P. Madahi, N. Shahtahmasebi, A. Kompany, M. Mashreghi, M.M. Bagheri-Mohagheghi, A. Hosseini, Phys. Scr. 84 (2011)
035801–035805.
[13] Min Zheng, Wu Jiaqing, Appl. Surf. Sci. 255 (2009) 5656–5661.
[14] M.G. Nair, M. Nirmala, K. Rekha, A. Anukaliani, Mater. Lett. 65 (2011) 1797–1800.
[15] I.S. Elashmawi, N.A. Hakeem, L.K. Marei, F.F. Hanna, Physica B 405 (2010) 4163–4169.
[16] B.D. Ngom, T. Mphane, N. Manyala, O. Nemraoui, U. Buttner, J.B. Kana, A.Y. Fasasi, M. Maaza, A.C. Beye, Appl. Surf. Sci. 255
(2009) 4153–4158.
[17] R. Chandramohan, T.A. Vijayan, S. Arumugam, H.B. Ramalingam, V. Dhanasekaran, K. Sundaram, T. Mahalingam, Mater. Sci.
Eng. B 176 (2011) 152–156.
[18] Seval Aksay, Yasemin Caglar, Saliha Ilican, Mujdat Caglar, Superlattices Microstruct. 50 (2011) 470–479.
[19] S. Banerjee, K. Rajendran, N. Gayathri, M. Sardar, S. Senthilkumar, V. Sengodan, J. Appl. Phys 104 (2008) 043913–043917.
[20] Zhiyun Zhang, Chonggao Bao, Wenjing Yao, Shengqiang Ma, Lili Zhang, Shuzeng Hou, Superlattices Microstruct. 49 (2011)
644–653.
190 M. Vasanthi et al. / Superlattices and Microstructures 55 (2013) 180–190

[21] K. Ravichandran, G. Muruganantham, K. Saravanakumar, S. Karnan, B. Kannan, R. Chandramohan, B. Sakthivel, Surf. Eng. 25
(2009) 82–87.
[22] R. Anandhi, R. Mohan, K. Swaminathan, K. Ravichandran, Superlattices Microstruct. 51 (2012) 680–689.
[23] M. Kojima, H. Kato, M. Gatto, Phil. Mag. B 68 (1993) 215–218.
[24] A. Bougrine, M. Addou, A. Kachouane, J.C. Bernede, M. Morsli, Mater. Chem. Phys. 91 (2005) 247–252.
[25] M.R. Vaezi, S.K. Sadrnezhaad, Mater. Sci. Eng. B 141 (2007) 23–27.
[26] G. Muruganantham, K. Ravichandran, S. Sriram, B. Sakthivel, J. Optoelectron. Adv. Mater. 14 (2012) 277–281.
[27] Mahmoud, E. Waleed, T. Al-Harbi, J. Cryst. Growth 327 (2011) 52–56.
[28] Kuan Jen Chen, Fei Yi Hung, Yen Ting Chen, Shoou Jinn Chang, Zhan Shuo Hu, Mater. Trans 51 (2010) 1340–1345.
[29] J. Panigrahi, D. Behera, I. Mohanty, U. Subudhi, B.B. Nayak, B.S. Acharya, Appl. Surf. Sci. 258 (2011) 304–311.
[30] R. Tankhiwale, S.K. Bajpai, Colloids Surf. B: Biointerf. 90 (2012) 16–20.
[31] R. Kohen, A. Nyska, Toxicol. Pathol. 30 (2002) 620–650.
[32] M. Fang, J.H. Chen, X.L. Xu, P.H. Yang, H.F. Hildebrand, Int. J. Antimicrob. Agents 27 (2006) 513–517.
[33] K.R. Raghupathi, R.T. Koodali, A.C. Manna, Langmuir 27 (2011) 4020–4028.

You might also like