2 Appl - Surf. Sci

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Applied Surface Science 458 (2018) 333–343

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

A critical study of the optical and electrical properties of transparent and T


conductive Mo-doped ZnO films by adjustment of Mo concentration
Sreenivasulu Reddy Tirumalareddygaria,1, Phaneendra Reddy Guddetib,1,

K.T. Ramakrishna Reddyb,
a
Department of Physics, Sir Vishveshwaraiah Institute of Science & Technology, Madanapalle 517325, India
b
Solar Energy Laboratory, Department of Physics, Sri Venkateswara University, Tirupati 517502, India

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

Keywords: Mo-doped zinc oxide (ZnO: Mo) transparent conductive thin films were prepared on the glass substrates using
Spray pyrolysis simple chemical spray pyrolysis technique by varying the Mo doping concentration in the range, 0–5 at.% at a
Mo-doped ZnO films constant substrate temperature, 400 °C. The effect of Mo-doping concentration on the physical behavior of ZnO
XPS analysis films was investigated. The X-ray Photoelectron Spectroscopy (XPS) analysis confirmed the presence of Zn, O,
Optical transmittance
and Mo in the layers with Mo in +6 state. The X-ray diffraction (XRD) patterns exhibited hexagonal wurtzite
Electrical parameters
crystal structure without any secondary phases. The microstructural analysis revealed the spherical nut-shaped
grains over the substrate surface. The optical studies revealed that the films with Mo-doping concentration of
2 at.% showed high optical transmittance and a wide band gap than pure and highly Mo-doped ZnO films. From
the optical transmittance versus wavelength data, the refractive index, extinction coefficient, dispersion con-
stants were evaluated. In addition, other optical parameters such as the optical conductivity, dielectric constants,
dissipation factor, electron energy loss functions and Haze were also calculated. The FTIR studies revealed the
presence of modes related to ZnO. Finally the electrical parameters such as resistivity, charge carrier mobility
and density of ZnO: Mo films were also analyzed and presented.

1. Introduction displays, light emitting diodes and solar cells [12–16]. ZnO films were
also used for chemical sensor application because of its high surface
Transparent conducting oxide (TCO) films have been used for dif- sensitivity to gases [17].
ferent applications, particularly for optoelectronic devices [1] and for Depending on the necessity and application, many researchers ex-
many decades detailed investigation has been carried out on these tensively studied doping of ZnO films using different transition metals
materials. The conventional TCOs are SnO2 and InO2 that possessed low such as Titanium (Ti) [18], Vanadium (V) [19], Chromium (Cr) [20],
metallic forms and also exhibit low stability when exposed to a hy- Manganese (Mn) [21], Iron (Fe) [22], Cobalt (Co) [23], Nickel (Ni)
drogen plasma [2,3], but zinc oxide films showed high stability in [24], Copper (Cu) [25], Zirconium (Zr) [26], Niobium (Nb) [27], Mo-
presence of hydrogen plasma [4]. Coming to the photovoltaic sector, lybdenum (Mo) [28], Gold (Au) [29] and Halogens such as Fluorine (F)
the most utilized TCO is tin-doped indium oxide (ITO) due to its low [30] and Chlorine (Cl) [31] also some metalloids such as Boron (B) [32]
resistivity, high transmittance and high work function [5,6]. However, and Germanium (Ge) [33] and also some pure metals such as Alumi-
ITO has some disadvantages such as a high energy barrier for injection nium (Al) [34], Tin (Sn) [35], Gallium (Ga) [36], Indium (In) [37] and
of holes at the ITO/hole transport layer and scarcity of metal indium. In Bismuth (Bi) [38] (see Fig. 1). Among all these dopant materials, Mo is
this scenario, doped ZnO films become promising candidates to replace particularly interesting due to its larger valency (5+ and 6+) with
ITO because they are nontoxic, inexpensive, earth-abundant and exhibit respect to that of Zn2+ ions, which suggests that each Mo atom can
comparable optical and electrical properties to ITO films. Due to the contribute 3 or 4 free electrons depending on valency to the ZnO lattice
low thermal expansion coefficient, wide energy band gap (3.3 eV) and that can alter the electrical conductivity significantly and reduce the
high exciton binding energy (60 meV) at room temperature [7–11], impurity ion scattering effect [39,40]. Moreover, its ionic radius
ZnO can be used as a transparent conductive material in flat panel (0.046 nm and 0.041 nm for Mo5+ and Mo6+ respectively) is lower


Corresponding author.
E-mail address: [email protected] (K.T. Ramakrishna Reddy).
1
These authors equally contributed to this work.

https://doi.org/10.1016/j.apsusc.2018.07.093
Received 9 February 2018; Received in revised form 13 March 2018; Accepted 12 July 2018
0169-4332/ © 2018 Published by Elsevier B.V.
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

drops of methanol. The substrate temperature was maintained constant


at 400 °C, while the Mo-doping concentration was varied in the range,
0–5 at.%, Solution flow rate of 6 ml/min was maintained during the
film deposition. Compressed air with a flow rate of 8 l/min was used as
the carrier gas. The precursor solution was sprayed on the substrate,
which was kept at a distance of 25 cm from the nozzle at an interval of
1 min.

2.3. Characterization of MZO thin films

The elemental composition of the films was studied using VG


Microtech ESCA 2000 X-ray photoelectron spectrometer. The structural
properties were studied using a Siefert X-ray diffractometer with a Cu
Kα radiation source (λ = 1.542 Å). The surface morphology was ana-
lyzed using an FSI Serion scanning electron microscope. The optical
Fig. 1. Various dopants used to dope ZnO in the literature. properties were studied using Perkin Elmer Lambda – 950 UV–Vis-NIR
spectrophotometer in the wavelength range, 300–900 nm. The Fourier
than that of Zn2+ (0.060 nm) so that Mo substitutes easily Zn ions in the transform infrared (FTIR) spectrum of the films was recorded in the
host lattice and occupies less interstitial positions [41]. wave number range, 500–4000 cm−1 using a Thermo Nicolet IR-200
Several deposition techniques were used by researchers to grow FTIR spectrophotometer. Electrical resistivity (ρ), carrier mobility (μ)
ZnO: Mo thin films such as spray pyrolysis [42–44], RF and DC mag- and carrier concentration (n) of the films were measured using the Hall
netron sputtering [45,46] and ion beam sputtering [47]. Among these effect measurement setup (model: HMS-3000).
techniques, spray pyrolysis is a simple and low-cost technique to de-
posit ZnO films as it is easy to dope the layers with appropriate dopants 3. Results and discussion
and it does not require vacuum for film deposition. In our laboratory,
investigations were already made earlier on doped ZnO films using The visual appearance of the as-grown layers revealed pale whitish
different dopants like Ga, Mg, Mn, Co, Ni by spray pyrolysis technique colour, which was pinhole free and strongly adherent to the substrate
[48–53]. The present work is a continuation of this research where Mo surface.
is used as the dopant in ZnO layers. Further, a complete optical analysis
of ZnO: Mo films are important for the design of heterojunction solar 3.1. Composition
cells because only optical properties give the information to understand
the optical transmittance/reflectance, photon loses and band structures. The XPS spectra were recorded in the binding energy range,
Many researchers studied the optical properties of ZnO: Mo films, 0–1100 eV in order to evaluate the chemical composition and the va-
however, they analyzed only the transmittance, absorbance and re- lence state of Mo ion in ZnO lattice. Fig. 3 shows the wide scan XPS
flectance properties including the determination of optical band gap, spectrum of Mo-doped ZnO films deposited at a substrate temperature
refractive index and extinction coefficient values. This analysis was of 400 °C with the molybdenum doping concentration of 2 at.%. The
extended further to find the dielectric constants, optical conductivity, spectrum showed two strong peaks related to Zn2p3/2 and Zn2p1/2 at
dissipation factor, dispersion parameters, optical Haze and electron 1021.24 eV and 1044.42 eV respectively. The presence of both the
energy loss functions using optical data. Generally, the chemical po- peaks at these binding energy values is different from those of ele-
tential of doping ions can tune the optical transmittance, which mini- mental Zn, indicating that Zn is present in +2 chemical state in ZnO
mizes the structural imperfection and optical losses in the films. Lit- lattice. The observed binding energy values of Zn were closely related
erature survey revealed that there were no reports on the effect of Mo to the reported values in the literature [54]. The peak observed at
concentration on optical and electrical properties of MZO films de- 530.26 eV is due to the lattice oxygen, corresponding to O 1s, bonded
posited by chemical spray pyrolysis technique. A key to achieving with Zn and Mo atoms. Since the activation energy for diffusion of
better optical and electrical properties lies in a good understanding of oxygen is much higher than that required for diffusion of interstitial Zn
the role of Mo concentration in MZO films. Therefore in this paper, we atoms and Mo6+ ions, sufficient oxygen atoms from the atmospheric air
study the effect of doping concentration on compositional, structural, can diffuse into ZnO lattice to fill up the new oxygen vacancies formed
microstructural, optical and electrical properties of Mo-doped ZnO thin due to the increase of Zn2+ and Mo6+ ions during the growth of layers.
films. Fig. 4 shows the variation of molybdenum peaks intensity in the
films with molybdenum content in the precursor solution. Mo3d5/2
peak has been observed at the binding energy of 229.1 eV, while
2. Experimental and characterization details
Mo3d3/2 peak at 232.8 eV. These peak positions mainly depend on the
local structure of Mo atoms in the ZnO lattice, which provides in-
2.1. Reagents
formation on its chemical state. The observed binding energy values of
Mo are different from the elemental binding energy values and also
Zinc chloride (ZnCl2) is taken as a precursor for zinc and mo-
those of Mo-O related compounds. This indicates that molybdenum
lybdenum chloride (MoCl5)2 as the precursor for Mo (both chemicals of
cations are perfectly substituted in the zinc sites of the wurtzite struc-
99% purity obtained from Sigma Aldrich) to prepare Mo-doped ZnO
ture. Further, it is observed that the Mo peak intensity increases with
layers. Methanol was used as the complexing agent and distilled water
increase of Mo concentration in the precursor solution, which indicates
as the solvent in order to prepare a clear final solution.
proper doping of ZnO with Mo in the layers.

2.2. Formation of ZnO: Mo (MZO) thin films 3.2. Structural properties

ZnO: Mo films were deposited by chemical spray pyrolysis (see The X-ray diffraction patterns of undoped and Mo-doped ZnO films
Fig. 2) on Corning 7059 glass substrates by using an aqueous solution of recorded in the 2θ range of 20–70° are shown in Fig. 5. It revealed that
350 ml containing zinc chloride and molybdenum chloride with few all the grown films were polycrystalline in nature, showing multiple

334
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Fig. 2. Experimental procedure used for deposition of MZO thin films via chemical spray pyrolysis technique.

Fig. 4. XPS narrow scan spectra of Mo 3d5/2 and Mo 3d3/2 of ZnO: Mo thin
Fig. 3. The full-range XPS spectrum of ZnO: Mo films grown at 400 °C with Mo- films.
content of 2 at.%.

develops the grain growth along the (0 0 2) plane, in order to minimize


peaks corresponding to (1 0 0), (0 0 2), (1 0 1), (1 0 3) and (2 0 0) planes the overall surface free energy [13]. Hence, it can be concluded that the
of hexagonal wurtzite crystal structure. All the observed planes are very incorporation of Mo into the ZnO lattice modifies the free energy dis-
well matched with the JCPDS card no: 75–0576. No peaks corre- tribution during the film growth process [55]. A similar change in the
sponding to zinc, molybdenum or its oxides were observed in the XRD preferred orientation was reported by Pawar et al. [56] in Boron-doped
pattern, which indicates that there is no additional phase present in the ZnO films grown by spray pyrolysis. The authors reported that the
films. Irrespective of doping concentration, all the as-deposited films substitutional incorporation of dopant atoms at regular lattice va-
showed peaks similar to pure ZnO, which indicates that there is no cancies as well as interstitial sites of the respective ion, will not give any
structural deformation occurred in host ZnO lattice upon Mo-doping. change in the direction of preferred orientations. But when the dopant
This confirms the successful replacement of Mo ions in Zn lattice sites in atom occupies additional interstitial sites, a change in preferred growth
the ZnO matrix. Added to this, Mo-doping strongly affects the peak was observed.
intensity. When the Mo concentration is 2 at.%, the strongest diffraction The lattice parameters were calculated by using the appropriate
intensity is observed along the (1 0 0) plane, which indicates that films formulae mentioned in relations (1)–(3) given below as available in the
with 2 at.% Mo-content had a better preferential growth. Further, it can literature [8,57,58].
be noticed that the intensity of (1 0 0) plane is reduced with increasing
Mo content and the preferred orientation changed to (0 0 2) plane. This λ
a=
indicates that more incorporation of Mo (> 2 at.%) into the ZnO lattice 3 sin θ (1)

335
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

volume present in the as-deposited films was estimated from the fol-
lowing relation [60],
1
δ=
D2 (5)

The calculated values of dislocation density (δ) with Mo-content are


listed in Table1. It is observed that δ value decreased up to 2 at.% Mo-
content in the films and then increased afterwards, which indicates that
films prepared with 2 at.% Mo was of high quality with less defects
present. Cagler et al. reported a similar behavior in Sn-doped ZnO films
formed by spray pyrolysis [61]. The lattice mismatch occurred between
the polycrystalline film and glass substrate is mostly due to the varia-
tion of lattice strain (ε), which is obtained by using the following re-
lation,

β cos θ
ε=
4 (6)

where θ is diffraction angle and β is the full width at half maximum


(FWHM) in radians. The minimum value of ε obtained for Mo = 2 at.%
indicates fewer defects between the substrate and the as-deposited
polycrystalline MZO film. However for higher doping concentrations,
the lattice strain increases with increase of Mo-content which can be
attributed to a change in the nucleation mechanism of ZnO. As Mo-
content in the layers increase, the number of nucleation sites in ZnO
also increases and this results in a smaller grain size as reported in
literature [58]. The preferential orientation or random growth of
polycrystalline films can be analyzed by calculating the texture coeffi-
cient Tc (hkl) for preferred planes, which is calculated from the fol-
Fig. 5. XRD pattern of MZO films prepared at various concentrations of Mo (at. lowing equation [62],
%).
I(hkl)
⎛ I ⎞
0(hkl)
λ Tc = ⎜ I(hkl) ⎟
c= 1
⎜ ∑ ⎟
sin θ (2) (7)
⎝N I0(hkl)

1 4 h2 + hk + k 2 ⎤ l2 where I(hkl) is the measured relative intensity of preferred orientation
= ⎡ + 2
d (2hkl) 3⎢
⎣ a2 ⎥
⎦ c (3) plane (hkl), N is the number of reflections observed in the XRD pattern.
I0(hkl) is the intensity of the standard diffraction pattern (JCPDS: 75-
where θ is the diffraction angle, λ is the wavelength of CuKα radiation 0576). From the calculated values, it is found that MZO films prepared
(0.154 nm) and (h k l) are miller indices. with 2 at.% Mo-doping showed a higher texture coefficient and fewer
From the evaluated lattice parameters, it is found that the lattice lattice defects than other ZnO: Mo films. This indicates that a large
constants and interplanar spacing (d) were decreased with increase of number of crystallites are oriented along the (1 0 0) plane parallel to the
Mo-content present in the films. From the evaluated values, it is clear substrate surface for 2 at.% Mo-doped ZnO layers. The peak intensity
that ZnO films doped with 2 at.% Mo showed very less variation in the was strongly influenced by the Lorentz factor, which mainly depends on
structural parameters with a = 0.326 nm, c = 0.563 nm and the integrated intensity (i.e. measured area under the diffracted peak).
d = 0.282 nm compared to pure ZnO films that had a = 0.316 nm, The integrated intensity of a reflection at any particular Brags angle
c = 0.547 nm and d = 0.274 nm. In both cases, the lattice constants ‘a’ depends on the number of particles oriented at or near the angle θ. The
and ‘c’ are found to be larger than that of bulk ZnO values (0.321 nm Lorentz factor was calculated by using the relation [63],
and c = 0.519 nm according to JCPDS: 75-0576). A similar behavior
was reported by Swapna et. al [55] in Mo-doped ZnO films grown by 1
L=
spray pyrolysis method. The crystalline size (D) of the films was cal- (4 sin2 θ cos θ) (8)
culated using the Debye-Scherrer formula [8],
where θ is the diffraction angle. The calculated values of L are listed in
Cλ table1. Finally, the stacking faults (SF), which are planar defects that
D=
β cos θ (4) characterize the disordering of crystallographic planes, were also de-
termined for the experimental films by using the relation given below
where β is the full width at half maximum (FWHM) in radians, θ is [64],
diffraction angle, λ is the wavelength of X-rays and C is the correction
factor, which is taken as 0.9. The calculated crystallite size of MZO films 2π 2
SF = ⎡ ⎤
⎢ 45(3 tan θ) 12 ⎥ β
is listed in Table 1. The crystallite size increased up to 2 at.% and then (9)
⎣ ⎦
decreased with further increase of Mo-concentration. A similar beha-
vior was observed by Lin et al. [59] for ZnO: Mo films prepared by dc where θ is the diffraction angle and β is the full width at half maximum
magnetron sputtering method. These results indicate that increase of (FWHM) in radians. The calculated values of SF are listed in Table 1.
Mo-doping concentration could degrade the crystallinity of MZO films The films prepared with 2 at.% Mo-doping showed low SF value than
because of the formation of stress due to the difference in the ionic size the other films, which indicates that very low planar defects were
between Mo and Zn, and also due to the segregation of dopant atoms at present in ZnO: Mo films prepared with 2 at.% Mo. For large band gap
grain boundaries for high doping concentrations [57]. Furthur the materials, it constitutes an energy barrier in the band structure of the
dislocation density (δ), which is the number of dislocation lines per unit crystal that can affect the current transport in semiconductor devices.

336
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Table 1
Structural parameters of ZnO: Mo films prepared using different Mo-contents.
Mo concen-tration (at. Crystal plane (hkl) Crystallite size, d (nm) Dislocation density (×1015) Lattice strain, ɛ Texture coefficient (Tc) Lorentz factor Stacking fault
%)

0 (1 0 0) 8.47 0.139 0.0042 5.31 3.28 0.0083


1 (1 0 0) 52.85 35.7 0.0006 5.67 3.48 0.0013
2 (1 0 0) 84.47 14.0 0.0004 6.51 3.47 0.0008
3 (0 0 2) 64.49 24.0 0.0005 3.70 2.98 0.0010
4 (0 0 2) 53.31 35.1 0.0006 2.96 2.96 0.0012
5 (0 0 2) 8.73 0.136 0.0042 2.60 2.97 0.0078

Fig. 6. SEM pictures of MZO films prepared using different Mo-concentrations and statistical distribution of grain size of MZO films prepared at Mo concentration of 0
and 2 at.%.

3.3. Morphological properties 450 nm compared with un-doped MZO film, where 60% of grains ex-
hibited the size between 260 and 280 nm, which could be useful to
Fig. 6 shows the SEM micrographs of ZnO: Mo layers with different enhance the performance of solar cells.
Mo-doping concentrations and statistical distribution of grains over the
substrate surface. All the layers showed ‘nut’ shaped grains that were 3.4. Optical properties
distributed over the substrate surface. It is well known from the particle
synthesis that the dopant atoms have a strong influence on the size of Optical properties in the semiconductor have their genesis in both
the resulting crystals. In analogy with this, the size of the grains in the intrinsic and extrinsic effects. Intrinsic optical transitions take place
films increased with Mo-concentration up to 2 at.%. An increase of between the electrons in the conduction band and holes in the valance
doping concentration decreases the grain size of respective layers. band including excitonic effects caused by the Coulomb interaction.
Further, it is observed that for Mo concentration of 2 at.%, the grains Extrinsic properties are related to dopants/impurities or point defects
grew bigger in hexagon shape compared to the rest of the concentra- and complexes, which usually create electronic states in the band gap
tions, which might be due to the effect of Mo-doping on the growth and therefore influence both optical absorption and emission processes,
process of the films [65]. The grain size from the SEM pictures was which results in the dissipation and variation of optical parameters.
estimated by using Fiji software. Fig. 6 shows the statistical distribution The optical transmittance of ZnO: Mo films grown on glass sub-
of the grains over the substrate surface. The un-doped MZO film showed strates using different Mo-dopant concentrations was studied in the
variation of grain size from 70 nm to 250 nm. However about 75% of wavelength range, 300–900 nm. Fig. 7 shows the optical transmission
grains demonstrated a size between 125 and 150 nm. But, the film versus wavelength spectra of ZnO: Mo films, which shows a sudden fall
grown at Mo-concentration of 2 at.% showed high grain size up to in optical transmittance near the fundamental absorption edge,

337
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Fig. 8. (αhυ)2 versus hυ plots for Mo-doped ZnO films.


Fig. 7. Optical transmission versus wavelength spectra of MZO films prepared
using different Mo-doping concentrations.
semiconductor and its relationship with the wavelength of incident
indicating the presence of direct band gap in these films. The layers light is known as dispersion. The optical constants are generally re-
doped with 2 at.% Mo-doping showed an optical transmittance of 90% presented by the refractive index (n) and extinction coefficient (k),
in the visible region that decreased at higher doping levels (> 2 at.%). which are the real and imaginary components of the complex refractive
This might be attributed to the increased scattering of light photons by index N = n + ik, respectively. The imaginary part then handles the
the defects that resulted due to doping [66]. From Fig. 7, it can be attenuation, while the real part accounts for refraction. The extinction
observed that the fundamental absorption edge is shifted towards coefficient is an amount of the fraction of light loss due to scattering
longer wavelength side with the increase of dopant concentration. The and absorption per unit distance.
structural analysis also revealed that the crystalline quality is degraded The refractive index and extinction coefficient of sprayed Mo-doped
with the increase of Mo concentration in the MZO films, which results ZnO thin films were calculated by using the relations.
in a decrease of transmittance. A similar observation was made by Lee 1+ R
et al. [67] for ZnO: Sn films prepared by sol-gel method. n=
1− R (12)

3.4.1. Absorption coefficient and optical band gap αλ


k=
The absorption coefficient (α) of prepared MZO films was evaluated 4π (13)
by using the following relation.
where R is the reflectance (%), α is the absorption coefficient and λ is
−ln(T ) the wavelength of the incident light. Fig. 9(a) and (b) shows the plots of
α=
t (10) ‘n’ and ‘k’ as a function of wavelength for ZnO: Mo layers prepared at
Here, t is the thickness of the layer and T is the transmittance of the different Mo-concentrations varying in the range, 0–5 at.%. From
layer. In the present work, the film thickness was taken as 300 nm. The Fig. 9(a), it is found that the value of ‘n’ increased with wavelength up
optical energy band gap (Eg) of the films was estimated using the em- to 650 nm and then decreased. Films prepared at 2 at.% Mo-con-
pirical relation, centration showed the highest refractive index than films with other
doping concentrations. The magnitude of ‘n’ varied from 1.19 to 2.60
αhν = A (hν−Eg )n (11) with the change of Mo-concentration in the films. From Fig. 9(b), it was
where A is a constant, hυ is the photon energy and n = 1/2, 2, 3/2 observed that for all layers, the extinction coefficient sharply decreased
and 3 for direct allowed, indirect allowed, direct forbidden and indirect first and slightly increased afterwards. In particular, the ‘k’ value de-
forbidden transitions respectively. Among these four types, the ex- creases to a minimum for wavelength up to 400 nm, indicating less
perimental data is best fitted for n = 1/2, indicating the presence of a absorption in this region. The magnitude of ‘k’ in the visible range
direct optical transition in these films. varied from 0.003 to 0.007 with the change of Mo concentration from
Fig. 8 shows (αhυ)2 versus (hυ) plots of MZO films. The energy band 0 at.% to 5 at.%. Wanes et al. [69] reported a similar change of ex-
gap was calculated by extrapolating the linear part of the plots in Fig. 8 tinction coefficient in sol-gel spin coated ZnO films.
onto the photon energy axis. The evaluated band gap initially increased
from 3.19 eV to 3.68 eV with the increase of Mo- concentration up to 3.4.3. Dispersion parameters
2 at.%. However, further increase of Mo-concentration, decreased the The energy dispersion parameters of the refractive index for ZnO:
band gap to 3.14 eV. This is due to the increased structural disorder at Mo films were investigated on the basis of single effective oscillator
higher Mo-doping that caused the generation of localized states in the model proposed by Wemple-DiDomenico. The refractive index ‘n’ at
band gap region, which in turn decreased the energy band gap. More- photon energies ‘hυ’ below the fundamental absorption edge is re-
over, the narrowing of energy band gap may be attributed to the strong presented by the relation [70].
s-d and p-d spin exchange interactions between the band electrons and E0 Ed
localized d electrons of Mo6+ ions substituting Zn2+ ions. Also, the n2 = 1 +
E02−(hυ)2 (14)
change in optical band gap can be explained in terms of Burstein-Moss
band gap narrowing and band gap widening due to the electro- where ‘E0’ is the single oscillator energy and ‘Ed’ is the dispersion
n–electron and electron-impurity scattering [68]. energy. These dispersion parameters are evaluated by fitting a linear
function in (n2 − 1)−1 against (hυ)2 plot. The values of E0 and Ed can be
3.4.2. Refractive index and extinction coefficient determined directly from the intercept and the slope on the vertical
Refractive index is the most complex optical constant of a axis. Fig. 10 shows the plot of (n2 − 1)−1 versus (hυ)2 for ZnO: Mo

338
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Fig. 9. (a) Change of refractive index and (b) extinction coefficient of MZO films with wavelength.

energy varies in the same manner as that of dispersion energy. It is clear


that the variation of oscillator energy values with Mo-doping con-
centration can be attributed to the variation in the optical transmit-
tance. On other hand, the variation of dispersion energy can be oc-
curred due to the change in crystalline quality with Mo-doping
concentration in the ZnO: Mo films. The moments of optical dispersion
spectra, M−1 and M−3 were also calculated using the following rela-
tions,

M−1 M −31
E02 = , Ed2 =
M−3 M−3 (15)

The plots of moments of optical dispersion against Mo-concentra-


tion were also represented in Fig. 11. The obtained results showed that
the moments of optical dispersion spectra increased with Mo-doping up
to a concentration of 2 at.% and then decreased to a minimum value at
higher doping concentrations.
Fig. 10. (n2 − 1)−1 versus (hυ)2 plots for with ZnO layers doped different Mo-
concentration.
3.4.4. Optical conductivity
The optical response of ZnO: Mo films was analyzed in terms of
layers prepared using various Mo-doping concentrations. The evaluated optical conductivity (σopt), where dimensions are same as the frequency
values of oscillator energy and dispersion energy for the prepared ZnO: and is calculated using the relation [71],
Mo films were plotted against Mo concentration and are shown in
Fig. 11. From the figure, it can be clearly observed that the oscillator αnc
σopt =
4π (16)

Fig. 11. Variation of E0, Ed, M−1 and M−3 with Mo-doping concentration in ZnO films.

339
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Fig. 13. Plot of dissipation factor, tanδ versus hυ of MZO films prepared at
Fig. 12. Variation of optical conductivity (σopt) with hυ for MZO films prepared
different Mo concentrations.
using different Mo-concentrations.

3.4.6. Electron energy loss functions


where α is the absorption coefficient, ‘c’ is the velocity of light and ‘n’ is
The loss of energy of a fast moving electron in a medium is given in
the refractive index. The variation of optical conductivity as a function
terms of surface energy loss function (SELF) and volume energy loss
of photon energy (hυ) for ZnO: Mo films is shown in Fig. 12. The value
function (VELF). SELF and VELF define the loss of energy of fast moving
of optical conductivity is approximately saturated up to 3.25 eV and for
electron via characteristic plasma excitation when travelling through
hυ > 3.25 eV a sudden increase in σopt was observed, because at wide
surface and bulk of the material respectively. The electron energy loss
band gap region the MZO films have certain charge carrier strength to
functions were calculated from the real and imaginary parts of di-
perform optical conductivity.
electric constants using the fallowing relations [72].

3.4.5. Dielectric constant and dissipation factor εi


SELF =
The fundamental electron excitation spectrum of the film is re- [(εr + 1)2 + εi2] (20)
presented by the complex dielectric constant, which is frequency de-
εi
pendent. The dielectric constant is defined as the combination of real VELF =
(εr2 + εi2) (21)
and imaginary parts as, ε(ω) = εr(ω) + εi(ω). The real part (εr) shows
how much it will slow down the speed of light in the material, whereas Fig. 14(a) and (b) represent the variation of surface and volume
the imaginary part (εi) gave a clear description about how a dielectric energy loss function values versus wavelength for MZO films respec-
absorbs energy from an electric field due to dipole motion. The di- tively. In the visible region, the MZO films prepared with 2 at.% Mo-
electric constants were calculated using optical constants ‘n’ and ‘k’ by doping showed very less value (nearly zero) of electron energy loss.
[71], Compared to VELF, low value of SELF is needed to avoid scattering
Er = n2−k2 (17) losses in solar cells.

Ei = 2nk (18) 3.4.7. Haze


The real part of dielectric constant of prepared ZnO: Mo layers The Haze parameter determines the level of light scattering and is
varied in the range, 4.4–6.8 and the imaginary dielectric constant defined as the ratio between the diffused and the total (spec-
changes in the range, 0.07–0.13. Tuzemen et al. [70] reported a similar ular + diffused) light intensity for the case of reflected and transmitted
value of real part of dielectric constant in ZnO films prepared by pulsed light at the textured interface. The Haze value was evaluated from the
filtered cathodic vacuum arc deposition (PFCVAD) method. The var- diffused and total transmittance dada using the relation [73].
iation of real part of dielectric constant is ordered by the character of Tdiffused
variation in the refractive index, whereas, the imaginary part of di- HT =
Ttotal (22)
electric constant is by the character of variation in extinction coeffi-
cient. The evaluated Haze parameter, measured in air as a function of
The dissipation factor is a measure of energy loss rate of electrical wavelength is shown in Fig. 15. Generally, the scattering process de-
oscillation mode in a dissipative system, which represents the quality of pends on interface morphology, refractive index of the incident medium
oscillation present in the films after inserting light on it. Dissipation and the type of incident light [74]. In the present work, the evaluated
factor is a ratio of real and imaginary parts of the dielectric constant Haze values vary in the range, 0.14 – 0.64 for MZO films of different
and is also called as loss factor. It can be calculated using the relation Mo-doping values. It is most typical that random texturization is pre-
sent in thin film solar cells. From Fig. 15, it can be noticed that undoped
εi
tan δ = ZnO film shows less haze compared with all other films because of the
εr (19)
smooth surface. The Haze value increased with the increase of Mo-
The plot of dissipation factor against photon energy was represented concentration present in the films. This indicates that increase of Mo-
in Fig. 13. The dissipation factor shows very less quantity (0.015) up to dopant concentration causes the film to have rough surface so that the
the photon energy 3 eV. But, in the higher photon energy range, 3–4 eV, vertical and lateral features of the grains present on the surface leads to
an increased dissipation factor was observed. However, ZnO: Mo films a rather complex optical situation at the textured interface. Among all
prepared at a Mo-concentration as 2 at.% showed less value compared the Mo-doped films, films prepared with 2 at.% Mo-doping showed low
to other layers, which indicates the presence of less defects in these Haze, indicating less scattering due to nut shaped grains all over the
films. surface.

340
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Fig. 14. Change of (a) surface energy loss function, SELF and (b) volume energy loss function, VELF with wavelength for MZO films prepared using different Mo-
concentrations.

Fig. 15. Haze parameter (measured in air) of ZnO:Mo films prepared using
different Mo-doping concentrations. The related SEM picture of film surface is
shown in the inset.
Fig. 16. FTIR spectra of MZO films prepared at Mo-concentrations of 0 at.% and
3.5. FTIR studies 2 at.%.

The FTIR spectra for undoped and Mo-doped ZnO films are shown in samples and the absorption in 2308 cm−1 is because of the presence of
Fig. 16. Usually, atoms/molecules exhibit great adsorption on the sur- CO2 molecules in air. Here we notice that although the frequency of
face of nanostructured materials as compared to their interior and certain vibrational modes remain the same for undoped and Mo-doped
surface chemistry of the nanomaterials is of immense interest [75–77]. ZnO nanostructures, the intensities of these modes vary due to struc-
In order to establish the presence or absence of various vibrational tural modifications occurred because of Mo-doping.
modes in ZnO: Mo films and subsequently to detect the possible al-
teration in the vibrational modes at the different stages of structural
3.6. Electrical properties
transformation, the FTIR spectroscopy of the prepared films was per-
formed. The infrared absorption peak observed at 654 cm−1 is common
The electrical properties of both undoped and ZnO: Mo films were
to doped and undoped ZnO films that correspond to the vibrational
studied by using Hall effect measurements. All the as-grown films
energies of ZneO [78]. The appearance of this mode re-affirms the
showed n-type electrical conductivity. Fig. 17 shows the variation of
formation of ZnO phase in all the films prepared, strongly supporting
electrical resistivity (ρ), carrier mobility (μ) and carrier concentration
the SEM and XRD results. Further to determine the influence of Mo-
(N) of ZnO: Mo films with various Mo-doping concentrations that vary
doping on the structural modification of ZnO films, the FTIR spectra of
in the range of 0–5 at.% deposited on glass substrates. The electrical
undoped and 2 at.% Mo-doped ZnO films were compared. The broad
resistivity first decreased with the increase of Mo-concentration in the
band at 775 cm−1 was assigned to CeH mode of vibration. The peak at
films that showed a minimum value for the doping concentration of
1009 cm−1 may be attributed to C]C stretching mode. The band pre-
2 at.%. However, when the Mo-doping concentration in the layers ex-
sent at 1376 cm−1 is attributed to the CeO stretching frequencies and
ceeds 2 at.%, the resistivity increased. The initial decrease of electrical
the peak at 1598 cm−1 to C]O. The absorption band at 3564 cm−1
resistivity was due to the increase of crystallite size with doping con-
indicates the existence of hydroxyl (eOH) groups on the surface of the
centration that results in a decrease of grain boundary area, followed by

341
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Fig. 17. Variation of electrical resistivity (ρ), carrier mobility (μ) and carrier concentration (N) with Mo concentration in MZO films.

a reduction in the scattering of the carriers at the grain boundaries transmittance approximately in the range 80–90% above the funda-
leading to a minimum resistivity [79]. This was also supported by the mental absorption edge. The energy band gap values varied in the range
XRD data obtained on the films grown with different Mo-doping con- 3.14–3.68 eV with the change of Mo-concentration. The films prepared
centrations. with Mo-concentration as 2 at.% showed high optical conductivity and
On further increase of Mo-concentration beyond 2 at.%, imperfec- low dissipation factor, indicating high optical response and low power
tions such as grain boundaries or dislocations are introduced in the loss. Also, this optimum doping value led to less electron energy loss
films due to the degradation of c-axis preferred orientation and crys- and low haze parameter. The Hall effect measurements showed that the
talline quality, which leads to an increase of defect concentration in the film prepared at this Mo concentration had the low resistivity of
layers. Many free carriers are then caught by the large number of traps 1.9 × 10-2 Ω cm, high mobility 41 cm2/V s and carrier concentration
formed by the defects, which results in a reduction in the carrier con- 7.8 × 1018 cm−3. Hence ZnO: Mo films prepared at 2 at.% of Mo exhibit
centration and an increase of electrical resistivity. Since the carrier better structural and opto-electrical properties that could be used as
mobility is affected mainly by the scattering process associated with TCO layers in heterojunction solar cells.
imperfections present in the films, the mobility of ZnO: Mo films de-
creases with the increase in Mo-concentration. The dependence of Acknowledgements
carrier density, resistivity, and mobility of the films on the Mo-con-
centration is very similar to the results reported by Gokulakrishnan The authors would like to thank Dr. N. Revathi and Prof. E.
et al. [42] for ZnO: Mo films formed by spray pyrolysis technique. Mellikov, Dept of Materials Science, Tallinn Technical University,
The electrical resistivity decreased for light Mo-doping concentra- Estonia for XRD and SEM analysis, and Prof. M. Topic, Department of
tion (≤2 at.%) and increased significantly for higher doping con- Electrical Engineering, University of Ljubljana, Slovenia for the optical
centrations (> 2 at.%). With the replacement of Zn2+ ions by Mo6+ measurements.
ions, the electron concentration increased considerably with the Mo-
content at lower doping i.e. up to 2 at.% resulting in lowering of film References
resistivity. At higher Mo-doping, the resistivity increased due to poor
crystal quality. The electron concentration was enhanced for lighter [1] Wu Zhang, Jie Gan, Lequn Li, Zhigao Hu, Liqun Shi, Ning Xu, Jian Sun, Jiada Wu,
doping due to the fact that Mo ions at the interstitial sites provided Mater. Sci. Semicond. Process. 74 (2018) 147–153.
[2] S. Major, S. Kumar, M. Bhatnagar, K.L. Chopra, Appl. Phys. Lett. 49 (1986)
more electrons. When doping was more, the electron concentration 394–396.
reduced significantly because of the formation of more Mo-related [3] D. Goyal, P. Solanki, B. Marathe, M. Takwale, V. Bhide, Jpn. J. Appl. Phys. 31
clusters and defects. The factors that influence the carrier mobility (1992) 361–364.
[4] D. Goyal, M.G. Takwale, V.G. Bhide, Proc. National Solar Energy Convention, Tata
might be ionizing impurity and lattice scattering. As Mo-doping content McGraw-Hill, New Delhi, 1990.
in the layers increased, these scatterings are strengthened due to [5] I. Hamberg, C.G. Granqvist, J. Appl. Phys. 60 (1986) 123–159.
emergence of Mo-related clusters and defects. [6] C.G. Granqvist, A. Hultake, Thin Solid Films 411 (2002) 1.
[7] A. Bougrine, A.E.I. Hichou, M. Addou, J. Ebothe, A. Kachouane, M. Troyan, Mater.
Chem. Phys. 80 (2003) 438–445.
[8] U. Ozgur, Y.I. Alivov, C. Liu, A. Teke, M.A. Reshchikov, S. Dogan, V. Avrutin,
4. Conclusions S.J. Cho, H. Morkoc, J. Appl. Phys. 98 (2005) 041301/1–041301/123.
[9] S. Pearton, D. Norton, K. Ip, Y. Heo, T. Steiner, Prog. Mater. Sci. 50 (2005) 293–340.
[10] S. Colis, H. Bieber, S. Begin-Colin, G. Schmerber, C. Leuvrey, A. Dinia, Chem. Phys.
Mo-doped ZnO thin films were successfully grown by chemical Lett. 422 (2006) 529.
spray pyrolysis technique by varying the Mo-dopant concentration in [11] A. Dinia, J. Ayoub, G. Schmerber, E. Beaurepaire, D. Muller, J. Grob, Phys. Lett. A
the range 0–5 at.% at constant substrate temperature of 400 °C. The XPS 333 (2004) 152–156.
[12] F. Maldonado, A. Stashans, J. Phys. Chem. Solids 71 (2010) 784–787.
analysis revealed proper Mo incorporation in ZnO lattice with the in- [13] J.B. Park, S.H. Park, P.K. Song, J. Phys. Chem. Solids 71 (2010) 669–672.
crease of Mo-content in the films and Mo presence in +6 state. The [14] Y.C. Lin, J.Y. Li, W.T. Yen, Appl. Surf. Sci. 254 (2008) 3262–3268.
structural studies indicated hexagonal wurtzite structure with a change [15] Y.C. Lin, S.J. Chang, Y.K. Su, T.Y. Tasi, C.S. Chang, S.C. Shei, C.W. Kuo, S.C. Chen,
Solid State Electron 47 (2003) 849–853.
in preferred orientation from (1 0 0) to (0 0 2) plane with the increase of
[16] M.G. Ambia, M.N. Islam, Sol. Energy Mater Sol. Cells 28 (1992) 103–111.
Mo-content in the layers. Films formed with 2 at.% Mo showed high [17] S.T. Shishiyanu, T.S. Shishiyanu, I. Oleg Lupen, Sens. Actuators B 107 (2005)
crystallite size and texture coefficient with minimum dislocation den- 379–386.
sity and lattice strain. All the films showed an average optical [18] J. Liu, S.Y. Ma, X.L. Huang, L.G. Ma, F.M. Li, F.C. Yang, Q. Zhao, X.L. Zhang,

342
S.R. Tirumalareddygari et al. Applied Surface Science 458 (2018) 333–343

Superlattice Microstruct. 52 (2012) 765–773. [50] L. Raja Mohan Reddy, P. Prathap, K.T. Ramakrishna Reddy, Curr. Appl. Phys. 9
[19] M.E. Koleva, P.A. Atanasov, N.N. Nedialkov, H. Fukuoka, M. Obara, Appl. Surf. Sci. (2009) 667–672.
254 (2007) 1228–1231. [51] K.T. Ramakrishna Reddy, V. Supriya, Y. Murata, M. Sugiyama, Surf. Coat. Technol.
[20] S. Yilmaz, M. Parlak, S. Ozcan, M. Altunbas, E. McGlynn, E. Bacaksiz, Appl. Surf. 231 (2013) 149–152.
Sci. 257 (2011) 9293–9298. [52] M. Rajendraprasad Reddy, K.T. Mutsumi Sugiyama, Ramakrishna Reddy, Adv.
[21] V.R. Shinde, T.P. Gujar, C.D. Lokhande, R.S. Mane, Sung-Hwan Han, Mater. Chem. Mater. Res. 602 (2012) 1423–1426.
Phys. 96 (2006) 326–330. [53] A. Ei Hichou, M. Addou, J. Ebothe, M. Troyon, J. Luminescence 113 (2005)
[22] Fei Gao, Xiao Yan Liu, Li. Yun Zheng, Mei Xia Li, Yong Mei Bai, Juan Xie, J. Cryst. 183–190.
Growth 371 (2013) 126–129. [54] C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, Handbook of X-ray
[23] Dhruvashi, P.K. Shishodia, Thin Solid Films 612 (2016) 55–60. Photoelectron Spectroscopy, Perkin Elmer, Eden Prairie, 1979.
[24] Xing Zhao, R.V. Erjia Liu, Jingsheng Chen Ramanujan, Curr. Appl. Phys. 12 (2012) [55] P. Singh, A. Kumar, D. Kaur Deepak, J. Cryst. Growth 306 (2007) 303–310.
834–840. [56] B.N. Pawar, S.R. Jadkar, M.G. Takwale, J. Phys. Chem. Solids 66 (2005)
[25] D.L. Hou, X.J. Ye, H.J. Meng, H.J. Zhou, X.L. Li, C.M. Zhen, G.D. Tang, Appl. Phys. 1779–1782.
Lett. 90 (2007) 142502/1–142502/3. [57] C. Zegadi, A. Abderrahmane, A. Djelloul, S. Hamzaoui, M. Adnane, D. Chaumont,
[26] V. Gokulakrishnan, S. Parthiban, K. Jeganathan, K. Ramamurthi, Appl. Surf. Sci. K. Abdelkebir, Int. Rev. Phys. 9 (2015) 39–45.
257 (2011) 9068–9072. [58] Nanda Shakti, P.S. Gupta, Structural and optical properties of sol-gel prepared ZnO
[27] J.M. Lin, Y.Z. Zhang, Z.Z. Ye, Appl. Surf. Sci. 255 (2009) 6460–6463. thin film, Appl. Phys. Res. 2 (2010) 19–28.
[28] S. Shi, G. He, M. Zhang, X. Song, J. Li, X. Wang, J. Cui, X. Chen, Z. Sun, Sci. Adv. [59] Y.C. Lin, B.L. Wang, W.T. Yen, C.H. Shen, Thin Solid Films 519 (2011) 5571–5576.
Mater. 4 (2012) 193. [60] G.B. Williamson, R.C. Smallman, Phil. Mag. 1 (1956) 34–46.
[29] Narendar Gogurla, Sayan Bayan, Poulomi Chakrabarty, Samit Kumar Ray, J. Lumin. [61] Y. Caglar, S. Aksoy, S. Ilican, M. Caglar, Superlattices Microstruct. 46 (2009)
194 (2018) 15–21. 469–475.
[30] L. Cao, L.P. Zhu, W.F. Chen, Z.Z. Ye, Opt. Mater. 35 (2013) 1293–1296. [62] R. Romero, D. Leinen, E.A. Dalchiele, J.R. Ramos Barrado, F. Martin, Thin Solid
[31] E. Chikoidze, M. Nolan, M. Modreanu, V. Sallet, P. Galtier, Thin Solid Films 516 Films 515 (2006) 1942–1949.
(2008) 8146–8149. [63] B.D. Cullity, Elements of X-ray diffraction, Massachusetts, 1956.
[32] Q. Huang, Y. Wang, S. Wang, D. Zhang, Y. Zhao, X. Zhang, Thin Solid Films 520 [64] M. Jothibasa, C. Manoharanb, S. Johnson Jeyakumara, P. Praveenc, I. Kartharinal
(2012) 5960–5964. Punithavathya, J. Prince Richarda, Sol. Energy 159 (2018) 434–443.
[33] D.J. Kim, M.H. Lee, J.S. Park, D. Do, W. Lee, M.H. Kim, T.K. Song, H.I. Choi, [65] P. Prathap, N. Revathi, Y.P. VenkataSubbaiah, K.T. Ramakrishna Reddy,
K.W. Jang, W.J. Kim, J. Korean Phys. Soc. 61 (2012) 920–923. R.W. Miles, Solid State Sci. 11 (2009) 224–232.
[34] C.A. Tseng, J.C. Lin, Y.F. Chang, S.D. Chyou, K.C. Peng, Appl. Surf. Sci. 258 (2012) [66] P.K. Benney Joseph, V. Manoj, Bull. Mater. Sci. 28 (2005) 487–493.
5996–6002. [67] J.H. Lee, B.O. Park, Thin Solid Films 426 (2003) 94–99.
[35] K.C. Yung, H. Liem, H.S. Choy, J. Phys. D: Appl. Phys. 42 (2009) 185002. [68] E. Burstain, Phys. Rev. 93 (1954) 638.
[36] V. Bhosle, A. Tiwari, J. Narayan, J. Appl. Phys. 100 (3) (2006) 033713. [69] H. Ben Wannes, R. Benabderrahmane Zaghouani, R. Ouertani, A. Araujo,
[37] B.C. Jiao, X.D. Zhang, C.C. Wei, J. Sun, Q. Huang, Y. Zhao, Thin Solid Films 520 M.J. Mendes, H. Aguas, E. Fortunato, R. Martins, W. Dimassi, Mater. Sci. Semic.
(2011) 1323–1329. Process. 74 (2018) 80–87.
[38] F. Chouikh, Y. Beggah, M.S. Aida, J. Mater. Sci. Mater. El. 22 (2011) 499–505. [70] Ebru Senadım Tuzemen, Sıtkı Eker, Hamide Kavak, Ramazan Esen, Appl. Surf. Sci.
[39] X. Xiu, Y. Cao, Z.Y. Pang, S. Han, J. Mater. Sci. Technol. 25 (2009) 785–788. 255 (2009) 6195–6200.
[40] H.K. Kim, S.H. Huh, J.W. Park, J.W. Jeong, G.H. Lee, Chem. Phys. Lett. 354 (2002) [71] Sreedevi Gedi, Vasudeva Reddy, Minnam Reddy, Chinho Park, Jeon Chan-Wook,
165–172. K.T. Ramakrishna Reddy, Opt. Mater. 42 (2015) 468–475.
[41] I. Soumahoro, S. Colis, G. Schmerber, C. Leuvrey, S. Barre, C. Ulhaq-Bouillet, [72] D. Tiwari, Tapas K. Chaudhuri, T. Shripathi, U. Deshpande, V.G. Sathe, J Mater Sci:
D. Muller, M. Abd-lefdil, N. Hassanain, J. Petersen, A. Berrada, A. Slaoui, A. Dinia, Mater Electron 25 (2014) 3687–3694.
Thin Solid Films 566 (2014) 61–69. [73] O. Kluth, B. Rech, L. Houben, S. Wieder, G. Schope, C. Beneking, H. Wagner,
[42] V. Gokulakrishnan, S. Parthiban, K. Jaganathan, K. Ramamurthi, Ferroelectrics 423 A. Loffl, H.W. Schock, Thin Solid Films 351 (1999) 247–253.
(2011) 126–134. [74] J. Krc, B. Lipovsek, M. Bokalic, A. Campa, T. Oyama, M. Kambe, T. Matsui, H. Sai,
[43] A. Boukhachem, B. Ouni, M. Karyaoui, A. Madani, R. Chtourou, M. Amlouk, Mat. M. Kondo, M. Topic, Thin Solid Films 518 (2010) 3054–3058.
Sci. Semic. Processing. 15 (2012) 282–292. [75] N. Goswami, P. Sen, Solid State Commun. 132 (2004) 791–794.
[44] R. Swapna, M.C. Santhoshkumar, Ceram. Int. 38 (2012) 3875–3883. [76] Y. Veera Manohara Reddy, T. Sravani Bathinapatla, T. Luczak, M. Osinska,
[45] J.L. Shi, H. Ma, G.H. Ma, J. Shen, Appl. Phys. A 92 (2008) 357–360. H. Maseed, L. Ragavendra, L. SubramanyamSarma, V.V.S.SMadhavi Srikanth, New
[46] C.C. Kuo, C.C. Liu, S.C. He, J.T. Chang, J.L. He, Vacuum 85 (2011) 961–967. J. Chem. 42 (2018) 3137–3146.
[47] B.D. Cullity, Elements of X-ray Diffraction, Addison-Wesley Publication Co., 1967. [77] Y.V. Manohara Reddy, V. Prabhakara Rao, M. Lavanya, M. Venu, M. Lavanya,
[48] K.T. Ramakrishna Reddy, T.B.S. Reddy, I. Forbes, R.W. Miles, Surf. Coat. Technol. G. Madhavi, Mater. Sci. Eng. C-Mater. Biol. Appl. 57 (2015) 378–386.
151 (2002) 110–113. [78] R. Elilarassi, G. Chandrasekaran, Mater. Chem. Phys. 123 (2010) 450–455.
[49] K.T. Ramakrishna Reddy, P. Prathap, N. Revathi, A.S.N. Reddy, R.W. Miles, Thin [79] M. Ohyama, H. Kozuka, T. Yoko, J. Am. Ceram. Soc. 81 (1998) 1622–1632.
Solid Films 518 (2009) 1275–1278.

343

You might also like