<<

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem Outcome of Incorporation in the Photocatalytic Activity of Nanoparticles

D. J. Jeejamol1, K. Jayakumari2, A. Moses Ezhil Raj1* 1Department of Physics & Research Centre, Scott Christian College (Autonomous), Nagercoil – 629 003, India 2Department of Physics, Sree Ayyappa College for Women, Chunkankadai, India

Abstract: Cadmium oxide (CdO) nanoparticles doped with different wt. % of copper have been prepared by means of chemically controlled co-precipitation method. The powder X-ray diffraction patterns of the prepared samples revealed the formation of single phase polycrystalline face centered cubic (Fm 3 m) CdO structure even after doping with Cu2+. Initially the crystallinity of CdO was improved on Cu doping, however, with the increase of dopant concentration, a change in lattice constant was noticed. Amorphization and/or composite formation were not observed for the present level of doping. The metal oxide phase formation was again established from its Cd-O stretching band at about 460 cm−1 and by the broad band with poor resolved shoulders at about 580 and 1000 cm−1 in the infrared spectra. Inclusion of Cu2+ in the CdO lattice was further ascertained from slight shift in peak positions and changes in its width. Raman spectral bands in pure and intentionally doped CdO around 70 and 260 cm-1 can be attributed to the LA-TA or LO-TO and TA+TO modes at the L-point of Brillouin zone respectively. The shift in the prominent bands and change in the Raman profile evidently proved the replacement of Cu2+ in the place of Cd2+. The photocatalytic activity of pure and doped CdO were tested for the degradation of methylene blue (MB) in aqueous medium under visible light. Copper doped CdO exhibited substantially higher visible-light-driven photocatalytic activity and the percentage of color abatement in the aqueous solution for 5 hours was good for the samples calcined at 400oC (71% for pure CdO, 79% for 1% Cu doped CdO and 72% for 2% Cu doped CdO). The degradation rate of methylene blue followed the pseudo-first order kinetics and the rate constant obtained (0.3 to 0.7 hr-1), proved the pure and Cu doped CdO nanoparticles as the best photo catalyst for the removal of color from textile waste water.

Keywords - FT-IR, FT-Raman, Nanoparticles, Photocatalytic, XRD ------*Corresponding author e-mail: [email protected] (A. Moses Ezhil Raj) | Phone: +91 4652 232888 | Fax: +91 4652 229800

I Introduction Transparent conducting (TCO) have attracted growing attention over the last decades as significant components of technological applications. Among these oxide, CdO is an important TCO with a variety of optoelectronic applications like, solar cells, optical communications, flat panel display [1–3]. CdO is an n-type semiconductor with optical of ~2.5 eV [4]. The n-type conduction of undoped CdO is attributed to its native vacancies and cadmium interstitials. Therefore, the conductivity of CdO nanoparticles can be controlled by those native defects [5] or by doping with metallic like: In [6], Sn [5], Al [7], Sc [8], and Y [9]. Generally, it was observed that doping with ions of radius less (to a limited extend) than that of Cd2+ slightly shrinks the CdO lattice parameters, increases the apparent energy gap, increases the carrier mobility and concentration, and hence increases the conductivity. For solar cell and other optoelectronic applications, it is welcomed to dope CdO with some ions in order to increase both the conductivity as well as the transmittance.

Published by SLGP. Selection and/or peer-review under responsibility of ICISEM 2016. Open access under CC BY-NC-ND license.

356

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem Due to the unique optical and electrical properties of CdO, it has a wide range of applications such as phototransistors, solar cells, catalyst, chemical sensors, electroplating baths, liquid crystal displays, IR detectors, transparent electrodes and gas sensors [10, 11]. Most of the semiconducting oxides are photocatalysits. Among the metal oxide semiconductors, cadmium oxide is found to be a promising candidate for this application, as it can be activated with visible light (<590 nm). Different soft chemical methods such as co-precipitation, sol-gel and hydrothermal synthesis have been attempted to obtain pure CdO and doped CdO nanoparticles [12]. Here in the preparation of cadmium oxide nanoparticles via chemically controlled co-precipitation method followed by calcination and its usefulness as photocatalyst to degrade methylene blue, which is a basic aniline dye, C16H18N3SCl utilized in coloring paper, temporary hair coloring, dyeing cotton and wools, and coloring of paper stocks has been investigated. Structural characterization of pure and Cu2+ doped CdO nanoparticles were done to test their appropriateness as photocatalysts.

II Experimental In the co-precipitation method, hydroxide was added drop-wise to 0.5M dihydrate (Cd(CH3COO)2·2H2O) under constant stirring until the pH value of the solution attains 8. Resulted precipitate was then filtered and washed repeatedly after 18 to 20 hours using double distilled water to remove impurities. The hydroxide thus formed was calcined at different temperatures 400, 600 and 800°C for 2h to produce nanocrystalline powders. For doped samples the Cu source, 0.5M copper chloride was dissolved in required amounts along with the 0.5M cadmium acetate.Structural information was obtained from the x-ray diffraction patterns recorded in PANalytical X’pert-pro instrumentation. Fourier transform infrared (FT-IR) spectra were recorded with a Thermo Nicolet, Avatar 370 spectrometer in the range 400-4000 cm-1 using KBr pellet technique. Raman spectra were recorded with a BRUKER RFS 27: Stand alone FT-Raman Spectrometer. To investigate the photocatalytic property, 162 mg of the photocatalysts was dispersed in the 50 ml methylene blue aqueous solution with an initial −5 concentration of 10 mol/L. The photodegradation of methylene blue (C16H18N3S) was then studied from the change in optical property of methylene blue (MB) aqueous solution in presence of pure and Cu doped CdO nanoparticles before exposure of light and after exposure of light for 1h difference using the Varian Cary-5000 UV spectrophotometer.

III Results and discussion

3000 2CuCdO 600 3000 2100 2CuCdO 400 2000 2CuCdO 800

2000

1400 1000

1000 700 30000 0 1CuCdO 600 30000 2100 2000 1CuCdO 800

1CuCdO 400 2000

1400 1000

1000

700 30000 0.5CuCdO 600 30000 0 2000 0.5CuCdO 800

2100 0.5CuCdO 400 2000

1400 1000

1000 700 0 3000 0 0 3000

Intensity(a.u.) 2000 2100 CdO 600 Intensity(a.u.) 2000

Intensity(a.u.) CdO 800

1000

1400 CdO 400 1000 700 0 0

0 90 JCPDS:75-0593 90 JCPDS:75-0593 (1 1 1) 1 (1

90 JCPDS:75-0593 60 1) 1 (1

(2 0 0) 0 (2 60

(2 0 0) 0 (2

(2 2 0) 2 (2

(1 1 1) 1 (1

(3 1 1) 1 (3 30 0) 2 (2 60 2) 2 (2

30 1) 1 (3

(2 2 2) 2 (2

(2 0 0) 0 (2

(2 2 0) 2 (2

30 1) 1 (3 0 (2 2 2) 2 (2 0 0 30 40 50 60 70 30 40 50 60 70 2(degrees) 2(degrees) 30 40 50 60 70 2(degrees)

Fig. 1 XRD patterns of pure and Cu doped CdO nanoparticles calcined at 4000C, 6000C and 8000C

In the XRD patterns of pure and Cu2+ doped CdO nanoparticles (Fig. 1), the observed “d” spacings and the respective prominent peaks correspond to reflections of (111), (200), (220), (311) and (222) planes and are in good agreement with the standard data (JCPDS card no.: 75-0593). Peak positions of the crystallized pure and doped CdO thus corresponds to the face centered cubic structure. To some extent the volume of unit cell of host CdO crystalline structure decreased with Cu doping due to smaller size of Cu2+ dopant ions in contrast to that of Cd2+. The lodging of Cu2+ in the position of Cd2+ to the shrinkage of lattice constants after doping. However, the substitution of copper does not affect the structure of CdO. [13].The decrease of intensity of the reflections with the increase of Cu% doping level, results in the deterioration of of CdO. Compression in the unit cell on doping was evidenced from the slight shift in diffraction peaks in the range 0.016–0.115° towards higher angles in Cu2+ doped CdO nanopowder [14]. This may be due to a small variation in the particles size [15]. The variations in the structural parameters with dopants and with calcination temperature are listed in Table 1.

357

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem

Table 1 Variation in lattice parameter, volume and crystallite size

Scherrer

Å

3

Sample Details Å Formula

(a) Lattice Lattice Volume D (nm) Parameter CdO 400 4.6937 103.40 39.32 CdO 600 4.6925 103.33 60.09 CdO 800 4.6852 102.84 107.25 0.5CuCdO 400 4.6978 103.67 73.83 0.5CuCdO 600 4.6984 103.71 85.03 0.5CuCdO 800 4.6878 103.01 139.88

1CuCdO 400 4.6908 103.22 46.57 1CuCdO 600 4.6952 103.5 93.99 1CuCdO 800 4.6921 103.3 138.32 2CuCdO 400 4.6929 103.35 47.96 2CuCdO 600 4.6936 103.4 74.17 2CuCdO 800 4.6914 103.26 110.69

The FTIR spectra (Fig. 2) of undoped and Cu2+ doped CdO nanopowder exhibited various vibrational bands related to metal oxides. The bands observed at 645–728 cm-1 are assigned to Cd-O metal oxide bonding [16]. CdO phase formation was further confirmed from its Cd-O stretching band at about 460 cm−1 and by the broad IR band with poor resolved shoulders at about 580 and 1000 cm−1 in the FTIR spectra [17]. Incorporation of Cu in the CdO lattice was ascertained from the slight shift in vibrational band positions as well as from the additional band, obtained on doping, around 550cm-1 which was due to Cu-O stretching.

2CuCdO 600

2CuCdO 400 2CuCdO 800

2370 1643 994 3486 2353 600

690

448 860 600 437 3496 1444 888 509 1406 922 500 1455 506

1CuCdO 400 1CuCdO 600 1CuCdO 800

2370

2362 1633 860 464 3424 940 850 596 1408 490 1639 527 1436 3445 1406 930 3449 0.5CuCdO 600 515 0.5CuCdO 800 0.5CuCdO 400

1606 1175

2350 1426

640 3468 845 2374 920 470

1595 1039 842 600 1422

% % Transmittance % % Transmittance

1395 456 % Transmittance Pure CdO 600 Pure CdO 400 501 Pure CdO 800

448 1434 1058 567 558

1066 710 1629 436 1629

1629 1427 588 3436 407 3424 3441 3500 3000 2500 2000 1500 1000 500 3500 3000 2500 2000 1500 1000 500 -1 3500 3000 2500 2000 -11500 1000 500 Wavenumber (cm-1) Wavenumber (cm ) Wavenumber (cm )

Fig. 2 FTIR Spectra of pure and Cu doped CdO nanoparticles

The deconvoluted (Gaussian multiple peak fit) Raman scattering (RS) peaks for pure and Cu doped CdO shown in Fig. 3 can be attributed to its weak second-order RS processes [18]. In the 50–80 cm-1 range the difference, two-phonon of states exhibits a maximum which can be attributed to LA-TA or LO- TO difference modes. In the 90–120 cm-1 region the one-phonon density of states exhibits sharp features arising from TA modes. The distinct peak at ~260 cm-1 was identified as a TA + TO mode at the L-point of Brillouin zone (BZ) according to two- phonon density of states (PDOS) calculations [19]. The absorption bands in the region 200–300 cm–1 correspond to Cd–O vibrational modes [20].

358

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem The observed shift in the wave number values of the different modes and the broad nature of the Raman bands indicate that the doping of CdO produces stress in CdO nanoparticles which is in support of the XRD findings. According to the molecular theory of vibration, Raman modes are specifically sensitive to local geometric disorientation to the neighboring disorder especially from sub-lattices or electric defects resulting from substitutions or vacancies. This point to the fact that the broadening of the characteristic Raman modes of crystalline CdO may originate from both the local disorientation and neighboring disorder introduced by Cu incorporation to the CdO host lattice [21]. Profile of the Raman peaks depends on the crystallization, residual stress, structural disorder and crystal defect in a sample and as a result the profile of the Raman peaks changed on doping and on different calcinations temperatures. The additional bands in the doped samples that correspond to the silent modes of CdO arise from the breakdown of the translational crystal symmetry induced by the non-native ions and defects that can result in the change of Raman activation and Raman peak shift.

2CuCdO600 346

159 159 302 302

74 74 2CuCdO400

106 106

272 272

223 223

364 364

85 85

173 173

319 319

262 262

221 221 73 73

1CuCdO600

284 284 198 198

115 115 1CuCdO400

254 254

333

202 202

289

75 75

237 237

75 75

75 75

256 256 266 266

0.5CuCdO400 99 0.5CuCdO600

298 298

378 378

169 169

Raman Intensity

Raman Intensity

78

184 184

258 258

113 113 104 104

CdO400 CdO600 325

261 261

76 76

100 200 -1 300 400 100 200 -1 300 400 Wavenumber cm Wavenumber cm

71 71

250 250 2CuCdO800

103 103

293 293

159 159

355 355

87 87 284

240 240 1CuCdO800

348 348

158 158

383 383

246 289 289

72 72 0.5CuCdO4800

139

360 360

Raman Intensity 76 76

242 242 CdO800

139 139

359 359

296 296

100 200 -1 300 400 Wavenumber cm

Fig. 3 FT-Raman Spectra of pure and Cu doped CdO nanoparticles for different calcination temperatures

In nature, semiconductors displays high absorbing behaviors based on their band-gap energy. The exceptional nanostructure of doped morphology allows them to absorb the solar-lights and promote photo- catalysis with higher efficiency [22, 23]. Considering the solar-light absorption of the nanostructures, it is intended simulated block-structures to act as photo-catalysts as well as solid phase adsorbent with higher capacity and sensitivity. In the present study, the prospect of using pure and Cu doped CdO semiconductor nanomaterial for the photocatalytic degradation of methylene blue dye (MB) was explored in the presence of visible irradiation. Solution with 10-5 M dye and 0.025 M photocatalyst was magnetically stirred for 5h period under sunlight. Decolorizing or removal efficiencies were calculated based on the absorption intensity of the methylene blue (MB) solutions at different times [24]. The optical absorption peak of MB at 665 nm was chosen to monitor the photocatalytic degradation process. With time increasing from 0 to 1 hr, the intensity of the characteristic absorption band at 665 nm decreased gradually, suggesting that the methylene blue was gradually photodegradated by the pure and doped CdO photocatalysts (Fig. 4). Starting from these absorption spectra, the dye concentration ‘[MB]’ at a given time‘t’, normalized as a function of time (h) is shown in Fig. 5 As −kt can be seen, the ‘[MB]’ . time has an exponential decay, characteristic of a first order reaction: [MB] = [MB]0e , where, ‘k’ is the rate constant and ‘[MB]0’ is the initial MB concentration. The linear fit between ln ([MB]0/[MB]) and irradiation time can be approximated as pseudo-first-order kinetics [25] (Fig. 6), the slope of which gives the rate constant ‘k’. ‘k’ value was found to be in the range 0.29755 to 0.73861 hr-1,which proves the best photocatalytic activity of pure and Cu doped CdO nanoparticles that make it a viable alternative for the removal of color from textile waste water.

359

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem The outcome of kinetic analyses show that a decrease in the MB concentration results in an increase in the apparent photocatalytic activity of the CdO nanoparticles. The entire photocatalysis process depends on the properties of the catalyst, and the effective light intensity in the system. Small-sized nanoparticles with great surface area are proficient substrates for absorption of light [26] which is essential for the enhancement of photocatalytic activity. Accordingly, the percentage of degradation of methylene blue in the aqueous solution for 5 hours is good for the samples calcined at low temperature (400oC): 71% for pure CdO (Crystallite size = 39 nm), 79% for 1% Cu doped CdO (Crystallite size = 46 nm) and 72% for 2% Cu doped CdO (Crystallite size = 48 nm).

0.7 (a) (b) (d) 0.6 0.6 (c) 0.6 0h 0.6 0h 0h 0h 1h 1h 0.5 1h 0.5 1h 0.5 2h 0.5 2h 2h 2h 3h 3h 3h 3h 0.4 4h 4h 0.4 4h 0.4 4h 0.4

5h 5h 5h 5h

0.3 0.3 0.3 0.3

Absorbance (a.u.) 0.2 0.2 0.2

Absorbance (a.u.) Absorbance (a.u.) Absorbance (a.u.) 0.2

0.1 0.1 0.1 0.1

0.0 0.0 0.0 0.0 550 600 650 700 550 600 650 700 Wavelength (nm) 550 600 650 700 550 600 650 700 Wavelength (nm) Wavelength (nm) Wavelength (nm)

0.6 0.6 (e) (f) 0.6 (g) 0h 0.40 (h) 0h 0.5 0h 0.5 1h 1h 1h 0h 2h 0.5 0.35 2h 2h 1h 3h 0.4 3h 0.4 3h 0.30 2h 4h 3h 4h 0.4 4h 5h 4h 5h 5h 0.25

0.3 0.3 5h

0.3

0.20

0.2 0.2 0.15

Absorbance (a.u.) 0.2

Absorbance (a.u.)

Absorbance (a.u.) Absorbance (a.u.) 0.10 0.1 0.1 0.1 0.05

0.0 0.0 0.0 0.00 550 600 650 700 550 600 650 700 550 600 650 700 550 600 650 700 Wavelength (nm) Wavelength (nm) Wavelength (nm) Wavelength (nm)

0.50 0.6 0.7 (l) (i) (k) 0.6 (j) 0.45 0h 0.6 0h 0.5 0h 0h 1h 0.40 1h 1h 1h 2h 0.5 2h 2h 0.5 2h 0.35 0.4 3h 3h 3h 3h 4h 0.30 4h 0.4 4h 4h 5h 0.4 5h

0.3 5h 5h 0.25

0.3 0.3 0.20

0.2 Absorbance (a.u.)

Absorbance (a.u.) 0.15 Absorbance (a.u.) Absorbance (a.u.) 0.2 0.2 0.10 0.1 0.1 0.1 0.05

0.0 0.00 0.0 0.0 550 600 650 700 540 560 580 Wavelength600 620 640 (nm)660 680 700 720 Wavelength (nm) 550 600 650 700 550 600 650 700 Wavelength (nm) Wavelength (nm)

Fig. 4 Variation of intensity of the absorption bands with irradiation time (a) CdO400 (b) CdO600 (c) CdO800 (d) 0.5CuCdO400 (e) 0.5CuCdO600 (f) 0.5CuCdO800 (g) 1CuCdO400 (h) 1CuCdO600 (i)1CuCdO800 (j) 2CuCdO400 (k) 2CuCdO600 (l) 2CuCdO800

1.1 1.1 1.1 1.1 1.0 CdO400 0.5CuCdO400 1.0 1CuCdO400 1.0 2CuCdO400 CdO600 0.5CuCdO600 1.0 1CuCdO600 2CuCdO600 0.9 CdO800 0.9 0.5CuCdO800 0.9 1CuCdO800 0.9 2CuCdO800 0.8 0.8 0.8 0.8 0.7 0.7

0.7 0.7

0.6 0.6 0.6 0.6

-kt -kt 0.5 0.5 -kt 0.5 [MB]Normalized [MB] = [MB] e 0.5 [MB]Normalized [MB] = [MB] e

0 [MB]Normalized

0 [MB]Normalized [MB] = [MB] e 0 0.4 0.4 0.4 0.4 [MB] = [MB] e-kt 0 0.3 0.3 0.3 0.3

0.2 0.2 0.2 0.2 0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5 Irradiation Time (hr) Irradiation Time (hr) Irradiation Time (hr) Irradiation Time (hr)

Fig. 5 The dye concentration ‘[MB]’ at a given time ‘t’ normalized as a function of time

360

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem

3.0 1.6 CdO400 3.0 1CuCdO400 3.5 CdO600 -1 0.5CuCdO400 1.4 Slope = k = 0.3474 hr 1CuCdO600 2CuCdO400 0.5CuCdO600 -1 2.5 CdO800 2.5 Slope = k = 0.3304 hr 2CuCdO600 0.5CuCdO800 1CuCdO800 3.0 1.2 2CuCdO800

) -1

0

) 0 ) 2.0 2.0 Slope = k = 0.59814 hr 2.5

1.0 0 -1 )

0 Slope = k = 0.73861 hr 0.8 -1 1.5 Slope = k = 0.59045 hr -1 2.0

1.5

Slope = k = 0.53545 hr -1

-1

0.6 Slope = k = 0.58144 hr

[MB]/[MB] Slope = k = 0.30696 hr

[MB]/[MB]

( [MB]/[MB]

( 1.5 (

1.0 [MB]/[MB] (

-ln 1.0

0.4 -ln

-ln -ln -1 1.0 Slope = k = 0.29755 hr -1 0.2 0.5 Slope = k = 0.63685 hr -1 0.5 Slope = k = 0.53314 hr-1 0.0 Slope = k = 0.30072 hr 0.5 0.0 -0.2 0.0 0.0 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 Time (hr) Time (hr) Time (hr) Time (hr)

Fig. 6 The linear fit between ln ([MB]0/[MB]) and irradiation time

IV Conclusion The present study outlines synthesis of pure and Cu2+ doped cadmium oxide nanopowder using the chemical co-precipitation method followed by calcination. XRD patterns revealed cubic (fcc) structure for the prepared pure and Cu2+ doped CdO nanoparticles. The values of lattice parameter, unit cell volume and crystallite size are calculated based on the XRD data. There were some changes in the crystallite size of pure CdO with the addition of Cu-dopant. Accordingly, it is possible to control the nanostructure shape and size by the proper doping of parent. By FTIR spectroscopy, the functional groups present in the pure and Cu2+ doped CdO nanopowder were examined and the doping was confirmed by the band appeared around 550cm-1 which was due to the Cu-O stretching. In the photocatalytic activity, the color abatement was maximal for the pure and doped samples calcined at 4000C. Thus, it can be concluded that, pure and Cu2+ doped CdO mediated degradation of methylene blue promises to be a versatile, economic, environmentally benign and efficient method of waste water treatment, if all parameters are properly optimized. In the Raman spectra, shift has been observed in the prominent peak positions of the doped samples due to the effect of additive ions.

References

[1] Z. Zhao, D. L. Morel, C. S. Ferekides, Thin Solid Films 413 (2002) 203. [2] M. Yan, M. Lane, C. R. Kannewurf, R. P. H. Chang, Appl. Phys. Lett. 78 (2001) 02342. [3] B. Sahin, Y. Gulen, F. Bayansal, H. A. Cetinkara, H. S. Guder, Superlattices Microstruct. 65 (2014) 56–63. [4] D. M. Carballeda-Galicia, R. Castanedo-Perez, O. Jimenez-Sandoval, S. Jimenez-Sandoval, G. Torres- Delgado, C. I. Zuniga-Romero, Thin Solid Films 371 (2000) 105. [5] Z. Zhao, D. L. Morel, C. S. Ferekides, 2002. Thin Solid Films 413, 203. [6] A. J. Freeman, K. R. Poeppelmeier, T. O. Mason, R. P. H. Chang, T. J. Marks, 2000. Mater. Res. Soc. Bull. 25, 45. [7] R. Maity, K. K. Chattopadhyay, 2006. Sol. Energy Mater. Sol. Cells 90, 597. [8] S. Shu, Y. Yang, J. E. Medvedova, J. R. Ireland, A. W. Metz, J. Ni, C. R. Kannewurf, A. J. Freeman, T. J. Tobin, 2004. J. Am. Chem. Soc. 126, 13787. [9] Yang, Yu., S. J. Shu, J. E. Medvedeva, J. R. Ireland, A. W. Metz, Jun, Ni, M. C. Hersam, A. J. Freeman, T. J. Marks, 2005. J. Am. Chem. Soc. 127, 8796. [10] G. Singh, I.P.S. Kapoor, R. Dubey, P. Srivastava, Mater. Sci. Eng. B 176 (2011) 121. [11] A.S. Kamble, R.C. Pawar, N.L. Tarwal, L.D. More, P.S. Patil, Mater. Lett. 65 (2011) 488. [12] J. B. Raoof, Ojani, and A. Kiani, J. Electroanal. Chem. 515 (2000) 45. [13] Li Hui, Zhang Yongzhe, Pan Xiaojun, Zhang Hongliang, Wang Tao, Xie Erqing, Journal of Nanoparticle Research 11 (2009) 917–921 [14] M. Vigneshwaran, R. Chandiramouli, B.G. Jeyaprakash, D. Balamurugan, J. Appl. Sci. 12 (2012) 1754. [15] V. Ramasam, K. Praba, G. Murugadoss, Spectrochim. Acta, Part A 96 (2012) 963. [16] K. Gurumurugan, D. Mangalaraj, S.K. Narayandass, J. Electron. Matter. 25 (1996) 765. [17] M. Ristic, S. Popovic, S. Music, Mater. Lett. 58 (2004) 2494. [18] R. Cusco, J. lbanez, N. D. Amador, L. Artus, J. Z. Perez, and V. M. Sanjose, J. Appl. Phys. 107, 063519 (2010). [19] R. Oliva, J. Ibanez, L. Artus, R. Cusco, J. Zuniga-Perez, and V. Munoz-Sanjose, J. Appl. Phys. 113, 053514 (2013). 361

Journal of Applied Science and Engineering Methodologies Volume.2, No.3, (2016): Page.356-362 www.scientistlink.com/jasem [20] J. W. Robinson 1991 Practical handbook of spectroscopy (ed.) J W Robinson (USA: CRC Press Inc.) 1st ed., Ch. 5, p. 533 [21] G. Gouadec, P. Colomban, Prog. Cryst. Growth Charact. Mater. 53 (2007) 1. [22] F. Wang, D. Zhao, Z. Guo, L. Liu, Z. Zhang, D. Shen, Nanoscale 5 (2013) 2864–2869. [23] Y. Zheng, C. Chen, Y. Zhan, X. Lin, Q. Zheng, K. Wei, J. Zhu, Journal of Physical Chemistry C 112 (2008) 10773–10777. [24] A. I. Inamdar, A. C. Sonavanl, S. K. Sharma, H. Im, P. S. Patil, J.AlloysCompd 495 (2010) 76. [25] Alamgir, WasiKhan, Shabbir Ahmad, A. H. Naqvi, Mater.Lett. 133 (2014) 28. [26] L. Wang, L. Chang, B. Zhao, Z. Yuan, G. Shao, and W. Zheng, Inorg. Chem. 47 (2008) 1443.

362