You are on page 1of 33

Author’s Accepted Manuscript

Microwave Assisted Synthesis of Perovskite


Structured BaTiO3 Nanospheres via Peroxo Route
for photocatalytic Applications

M. Thamima, Y. Andou, S. Karuppuchamy

www.elsevier.com/locate/ceri

PII: S0272-8842(16)31720-5
DOI: http://dx.doi.org/10.1016/j.ceramint.2016.09.194
Reference: CERI13854
To appear in: Ceramics International
Received date: 3 August 2016
Revised date: 26 September 2016
Accepted date: 27 September 2016
Cite this article as: M. Thamima, Y. Andou and S. Karuppuchamy, Microwave
Assisted Synthesis of Perovskite Structured BaTiO3 Nanospheres via Peroxo
Route for photocatalytic Applications, Ceramics International,
http://dx.doi.org/10.1016/j.ceramint.2016.09.194
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Microwave Assisted Synthesis of Perovskite Structured BaTiO3 Nanospheres via Peroxo

Route for photocatalytic Applications

M. Thamimaa, Y. Andoub, S. Karuppuchamya*

a
Department of Energy Science, Alagappa University, Karaikudi-630 003,Tamilnadu, India
b
Graduate School of Life Science and Systems Engineering, Kyushu Institute of Technology, 2-4
Hibikino, Kitakyushu, Fukuoka 808-0196, Japan

skchamy@gmail.com

skchamy@alagappauniversity.ac.in
*
Corresponding author. Department of Energy Science, Alagappa University, Karaikudi-630 003,

Tamilnadu, India

Abstract

A novel perovskite structured barium titanate (BaTiO3) nanopowder was successfully

synthesized at low temperature by a facile and an inexpensive microwave method. The

synthesized nanosphere shaped BaTiO3 was analyzed using advanced characterization

techniques such as X-ray diffraction analysis, Fourier transform infrared, Micro Raman,

Scanning electron microscope with energy dispersive analysis and ultraviolet-visible diffuse

reflectance spectroscopic techniques. Photocatalytic dye degradation study was carried out

using newly synthesized BaTiO3 catalyst. The synthesized sphere shaped barium titanate

nanopowders effectively degrade the various dyes such as Methylene Blue (MB), Malachite

Green (MG) and Alizarin Red S (ARS) under UV light irradiation. The band gap energy of

synthesized BaTiO3 was calculated by UV-DRS spectra using kubelka-Munk equation. The

higher dye decomposition efficiency of nearly 100% was successfully achieved by BaTiO3

(BT10) nanosphere which was prepared using 10 min microwave irradiation.


Keywords

Alizarin Red S; Barium Titanate; Methylene Blue; Malachite Green; Perovskite.

1. Introduction

Photocatalysis is a promising oxidation method which has the prospective to

decompose dye effluents and eradicate microorganisms. Titanium dioxide nanomaterials have

been extensively used in photocatalytic degradation process. Hence, TiO2 nanomaterials are

broadly called as first-generation photocatalyst [1-11]. Semiconductor oxide photocatalytic

materials show very little catalytic activity under visible light, therefore, it is necessary to

develop alternate efficient photocatalyst. The barium titanate (BaTiO3) material has attracted

much attention because of their potential applications in various fields. The established

BaTiO3 photocatalyst was used to perform a dye decomposition study and it works

effectively due to its empty ‘d’ orbitals [12-17].

Barium titanate is a electro-ceramic material because it has some good physical

properties such as high permittivity etc. [18]. Hence, BaTiO3 nanomaterial has shown

interesting attention owing to its unique ferroelectric, magnetic, high dielectric and

piezoelectric properties. Moreover, it also used in wide variety of applications such as

capacitors, sensors, electronic applications, actuators [19], coatings for hospitals, thermistors,

laboratories, tiles fabrications and photocatalytic applications. BaTiO3 materials have

perovskite structure and the general formula of perovskite is ABX3. In this perovskite

structure, ‘A’ indicates the larger cation occupied by the sites of A, and ‘B’ indicates smaller

cation occupied by the sites of B. Perovskite families exhibit excellent physical properties

such as dielectric, ferroelectric, piezoelectric, pyroelectric behaviour and also it has plentiful

industrial applications [20]. Some of the examples of perovskite compounds are BaTiO3,
CaTiO3, PbZrO3, LiTiO3, PbTiO3, ZnTiO3, NiTiO3, PbLaZrTiO3 etc., which are mainly used

as piezoelectric compounds [21-23].

In recent years, the barium titanate has considered as an interesting material for

photocatalytic applications. The present work focuses on the development of BaTiO3

nanopowder at low temperature by a single-step wet chemical peroxo route. The developed

barium titanate nanomaterial was subjected to microwave irradiation for about 10, 20 and 30

min and denoted as BT10, BT20 and BT30, respectively. The synthesized materials were

analysed using advanced characterization techniques such as XRD, HR-SEM, FT-IR, Raman

and UV-DRS. The photocatalytic performance of BaTiO3 was studied using methylene blue

(MB), malachite green (MG) and alizarin red S (ARS) dyes. The decomposed dye solutions

were characterized by UV-visible spectroscopy.

2. Materials and methods

2.1. Synthesis of BaTiO3

The BaTiO3 nanomaterial was prepared by microwave assisted peroxo route. In a

typical reaction, the stoichiometric ratio of Ba(C2H3O2)2 and TiOSO4 salts were grounded

separately using mortar and pestle for 30 minutes. First, Ba(C2H3O2)2 was diluted using 50

ml of double distilled (DD) water under constant stirring for 1h. 0.1 M H2O2 solution was

added into TiOSO4 powder and it was diluted using 50 ml of DD water under constant

stirring until it forms a peroxo complex. After 1 h stirring, the two separate solutions were

mixed together to form a red-orange coloured solution and 5 ml of NaOH solution was added

for avoiding the formation of BaCO3. The colloidal pale yellow coloured precipitate was

obtained at the bottom of the beaker after completion of the reaction and the resultant

precipitate was heated at 95 oC for 30 min. The collected precipitate was washed with double

distilled water and ethanol and finally it was subjected to microwave irradiation at 400W for

10 min, 20 min and 30 min. The obtained powders are denoted as BT10, BT20 and BT30.
The photocatalytic dye degradation study was carried out using BaTiO3 powder and three

different dyes such as Methylene Blue (MB), Malachite Green (MG) and Alizarin Red S

(ARS).

2.2. Characterization of BaTiO3

The following analytical techniques were used to confirm the formation of barium

titanate. The advanced characterization techniques such as X-ray diffraction (XRD),

Scanning electron microscopy (SEM) with energy dispersive analysis (EDX), Fourier

transform infrared spectroscopy (FT-IR), Micro-Raman, UV-vis and UV-DRS were used.

Philips X’Pert PRO SUPER X-ray diffractometer equipped with monochromatized CuKα

radiation was used to analyse the crystallinity, lattice structure and phase of the synthesized

powders. Surface morphology of the samples was observed using X-650 scanning electron

micro-analyzer (JEOL-6300F, 15 kV) [24-27].

2.3. Photocatalytic Investigations

The photocatalytic performance of the BaTiO3 was investigated using various dyes

such as methylene blue, malachite green and alizarin red S under UV light irradiation. The

photocatalytic experiment was carried out using the following procedure. 100 mg of the

BaTiO3 powder (BT10) was added into 10 ppm dye solution. In order to investigate the

influence of catalytic behaviour of BaTiO3, various dyes and 100 mg of BaTiO3 powder

were added in 100 ml of 10 mg/L MB, MG and ARS dye solution, respectively. Prior to the

experiment, the dye and catalyst mixture was sonicated for 10 min and kept in the dark for 1

h to obtain the adsorption/desorption stability between dye and BaTiO3 nanocrystals. Then,

the mixture (catalyst with dye solution) was transferred into a 250 ml Borosil cylindrical jar.

Finally, the UV light (254 nm) was irradiated using 6W Hg lamp. The dye solution was

removed from the cylindrical jar after every 15 min of irradiation time and then the
absorption spectrum was measured by Ultraviolet–Visible spectrophotometer. The percentage

of dye degradation efficiency can be calculated using the following formula as:

Efficiency (%) = ln (Co−Cf)/Co ------ (1)

where, Co is the initial concentration considering (MB or MG or ARS) dye, and Cf is

the final concentration considering (MB or MG or ARS) dye absorption after completion the

photo-illumination reaction [28-29].

2.4. Plausible Photocatalytic Reaction Mechanism

The prepared dye solution with catalyst was subjected to UV light (6W Hg lamp)

irradiation and consequently electron-hole pairs may be produced in the photocatalyst. At the

surface of the BaTiO3, the hydroxyl ions (OH-) and super oxides (O2-) could be generated

.
from air (containing oxygen and moisture). Then, it may generate hydroxyl radicals (OH )

.
and oxide radicals (O2 ) and ultimately these radicals to form CO2 and H2O by an efficient

interaction with the harmful organic substances [30-31]. Plausible photocatalytic reaction

mechanism is shown in scheme 1.

3. Results and Discussion

3.1. XRD
Fig.1. shows the XRD patterns of the as-synthesized microwave irradiated well

crystallized products. The XRD show peaks correspond to hexagonal BaTiO3 nanoparticles

and secondary phases (Ba2TiO4). All the diffraction peaks are indexed well with JCPDS card

no. 82-1175 which is corresponding to the P63/mmc space group. The X-ray diffraction

patterns confirmed the presence of BaTiO3 crystallites (Fig.1). The successful formation of

BaTiO3 nanoparticles is further confirmed by EDX spectrum owing to the presence of

chemical composition of Ba and Ti elements.


3.2. FT-IR

The FT-IR spectra of synthesized BaTiO3 nanocrystallines was recorded using FT-IR

spectrophotometer in the wave number range of 4000-400 cm-1. Fig. 2 shows FT-IR spectra

of BaTiO3 nanomaterials prepared by microwave irradiation of 10 min, 20 min and 30 min.

The FT-IR spectra display the bands randomly at 642, 813, 985, 1071, 1197, 1275, 1394,

1428, 1566, 1645, 1764 and 1856 cm-1. There is no broad band appears in the range of 3000-

3500 cm-1, which clearly shows that there is no moisture content in the synthesized sample.

The peaks at 1645 cm-1 may be assigned to the small amount of humidity or water. The band

appears between the ranges of 509-680 cm-1 indicates the presence of Ti-O stretching bond

[32]. The two characteristic bands appear at 1566 and 1428 cm-1 range, which are

corresponding to COO- asymmetric and symmetric stretching vibrations, respectively [33].

Peaks at 1071 cm-1 may be assigned to Ti-O-Ba bond. The absorption peak appears in the

range of 1733 cm-1 in the literature [34] which is here shifted to 1764 cm-1, and it indicates

the C=O stretching vibration of carboxylic group. The absorption bands at 1394, 1071, and

985 cm-1 can be considered as C-H bending modes of vibrations. The absorption peaks at 800

cm-1 is related to M-O bands, therefore the band at 813 cm-1 may be assigned to Ba-O or Ti-O

bond [35].

3.3. Raman spectroscopy

Fig. 3 shows the Raman spectra of BaTiO3 nanocrystals grown under microwave

irradiation. Besides, all phonons of the hexagonal P63/mmc symmetry (confirmed by XRD)

represented as three active Raman shifts, which are 2A1(TO) + E(LO) modes. The obtained

Raman peaks are in the range of 163, 516, and 702 cm-1. All the three spectra (Fig. 3) have

similar shapes, but the peak in 3(c), shows (BT30) high intensity at 163 cm-1 and the peaks at

516, 163 cm-1 corresponds to TO modes of A1 symmetry of BaTiO3 nanocrystal. The low

intensity peak at 516 cm-1 should be assigned to the coupling of TO modes associated with
hexagonal phase. The peak at 184 cm-1 should be assigned to the decoupling of TO modes

and it was reported [36] in our previous study [30-31]. Raman spectrum exhibited band at

163 cm-1 suggests that decoupling of TO modes i.e. A1(TO) phonons. The Raman shift at 720

cm-1 suggests that E(LO) modes [37] but in our study, the peak appears in the range of 702

cm-1, which may be E(LO) modes.

3.4. Scanning Electron Microscope with EDX Studies

Shape, size, and morphological distribution of BaTiO3 nanoparticles were examined

by SEM. Figure 4(a) and 4(b) shows the typical SEM morphological features of nanosized

BaTiO3 materials. Fig. 4(a) represents the 10 min microwave irradiated BaTiO3 (BT10)

sample and Fig. 4(b) represents the 30 min microwave irradiated BaTiO3 nanomaterials

(BT30). The smooth surface morphology of the BaTiO3 was observed in the synthesized

BaTiO3 samples. The SEM photograph depicts the hexagonal shape like morphology of

BaTiO3 and were clearly matches with XRD pattern. The average particle size of BaTiO3

(BT10) was calculated by SEM which is about 400 nm.

The elemental distributions were confirmed by the energy dispersive spectroscopic

analysis (EDX). The diagram of Ba, Ti and O elements are shown in Fig. 5(a) and 5(b) and it

confirms the formation of BaTiO3. There is no other elemental impurity found from EDX.

The atomic % and the weight % of elements are given in Table 1 and 2. The EDX spectra of

the BaTiO3 show the existence of Ba, Ti and O elements in the range of 11.95, 21.39 and

22.02 at%, respectively. Table 1 and Table 2 shows the existence of Ba, Ti and O elements at

around 7.42, 21.39 and 26.45 at%, respectively.

3.5. Ultraviolet - visible Diffuse Reflectance Spectroscopy


UV-DRS of the samples were measured using a UV-Shimadzu UV–Vis spectrometer.

The band gap of BaTiO3 material was calculated using Kubelka–Munk formula from UV–

DRS spectra. The Kubelka–Munk function can be expressed as follows:

F(R) = (1 − R)2 / 2·R ------ (2)

where, R is the reflection coefficient. Figure 6 shows the absorption spectrum of the barium

titanate nanocrystalline materials. The optical absorption spectrum was studied for the

microwave irradiated BaTiO3 samples in the wavelength range of 250 nm–2500 nm. The

earlier studies demonstrated that the band edge emission range of 400–500 nm is due to blue

shift emission and the emission bands in the range of 500–700 nm is due to red-shift emission

[38]. In the present study, the synthesized samples (BT10, BT20 and BT30) exhibited the

electron–hole recombination bands in the range of 520 nm (UV region) which is due to the

red shift emissions (Fig. 6).

The band gap of BaTiO3 samples was evaluated using UV-DRS spectra. Band gap

was plotted against energy Vs calculated Kubelka-Munk values. Fig. 7 (a), (b) and (c) shows

the band gap value for BT10, BT20 and BT30 samples, respectively. BT10 has a lower band

gap of 2.32 eV compared to BT20 and BT30 which are 2.71 and 3.01 eV, respectively. A

lower band gap energy material (BT10) is promising for higher photocatalytic activity.

4. Photocatalytic Applications

The photocatalytic activity of the synthesized BT10, BT20 and BT30 photocatalysts were

evaluated by photocatalytic decomposition of various dyes such as methylene blue (MB),

Malachite Green (MG) and Alizarin Red S (ARS) under UV light illumination. The

photocatalytic decomposition study was carried out in 10 mg/L of each MB, MG and ARS dye

solutions using 100 mg of already prepared photocatalyst of BT10, BT20 and BT30. The dark

absorption test was also carried out. The concentration of photocatalyst plays an essential role in
the photocatalytic decomposition study. The BT10 photocatalyst shows higher decomposition

efficiency among the three synthesized samples due to the presence of sphere shape BaTiO3

nanomaterials and lower band gap. The B10 powder may have many active sites and ultimately

the produced active hydroxyl radicals shown in Fig. 6, 7 and 8.

4.1. Methylene Blue (MB) Dye

In order to examine the photocatalytic activity of synthesized BaTiO3, MB dye was used

for this study. The MB dye was completely decomposed within 240 min under UV light

irradiation. The decomposition efficiencies of 85%, 79% and 68% was achieved for BT10, BT20

and BT30 photocatalyst, respectively, while MB dye was used as a medium (Fig. 8a). The

present study clearly demonstrates that the BT10 sample has a greater decomposition efficiency

than BT20 and BT30 (BT10 > BT20 > BT30).

4.2. Malachite Green (MG) Dye

Similarly, the photocatalytic activity was examined by the presence of synthesized

photocatalysts using MG dye under UV light illumination. The MG dye was completely

decomposed within 60 min. BT10 material shows 100% decomposition efficiency. BT20 and

BT30 exhibit 85% and 80% degradation efficiency, respectively (Fig. 8 b). BT30 sample shows

higher dye adsorption capacity, which was examined in the dark test. The MG dye solution with

BT30 sample was subjected to sonication for 10 min and then this solution was kept in dark

condition about 1 h. The green colour dye solution turned into a colourless solution i.e., MG dye

molecules entirely absorbed by the surface of the BT30 material. Again, the colourless solution

was subjected to UV light irradiation. Then MG dye colour was returned and after that, the

photocatalytic study was performed for 60 min under UV light. After 60 min of UV light

illumination, BT30 sample completely decomposed the MG dye. In this case, BT10 have a

higher decomposition efficiency than BT20 and BT30 (BT10 > BT20 > BT30).

4.3. Alizarin Red S (ARS) Dye


Fig. 8(c) shows the photocatalytic activity of synthesized BaTiO3 photocatalyst using

ARS dye. BT10 catalyst shows higher decomposition efficiency of 100% within 60 min, BT20

shows about 84% of degradation efficiency in 120 min and BT30 shows 81% of degradation

efficiency in 180 min. The present study also shows that BT10 has a highest decomposition

efficiency than BT20 and BT30 (BT10 > BT20 > BT30).

Photocatalytic performance of the synthesized materials (BT10, BT20 and BT30) was

successfully examined using various dyes such as MB, MG and ARS. It has been found that

BT10 photocatalyst exhibits the excellent photocatalytic activity compared to BT20 and BT30,

due to the presence of sphere shaped BaTiO3 nanomaterials and lower band gap which was

calculated by UV-DRS spectra. The BT10 powders may have highest active site and ultimately

produced many numbers of active hydroxyl radicals.

4.4. Ultra Violet Visible Spectroscope

Figure 9(a), 9(b) and 9(c) show the UV–Vis spectra of MB dye in the presence of 100

mg of (a) BT10 (b) BT20 and (c) BT30 under the irradiation of UV light. The UV–Vis

spectra of MB dye was measured at different time intervals. The MB dye exhibited maximum

absorption bands at 665 and 290 nm wavelengths. The reaction time increases with the

decrease of the MB dye absorption. Similarly, MG dye decomposition was evaluated using

UV-Vis spectroscopy at various time intervals. The typical absorption peak of the MG dye

appears at 230, 320 and 620 nm wavelengths which are shown in Fig. 10(a). The

decomposition of the dye was increasing while increasing the UV light illumination time.

Similarly, the ARS dye decomposition was evaluated using UV-Vis spectroscopy at various

time intervals. The typical absorption peak of the ARS dye appears at 260, 330 and 480 nm

wavelengths which are shown in Fig. 10(b). The ARS dye absorption band intensity

decreases with increasing the reaction time intervals, i.e. the dye decomposition is directly

proportional to the UV light illumination time. The present investigation clearly evinces that
the developed barium titanate photocatalyst is capable of degrading the MB, MG and ARS

dye effectively.

4.5. Kinetic Studies

Kinetic investigations were carried out to study the photocatalytic decomposition of the

methylene blue (MB), malachite green (MG) and alizarin red S (ARS) dye. The photocatalytic

degradation of MB, MG and ARS were found to follow first-order kinetics. The kinetic equation

can be expressed as follows:

ln Ci=Ct ¼ kt ------ (3)

where, Ci is the initial concentration of the dye solution, Ct is the dye solution concentration at ‘t’

time, k is the rate constant (first-order). The kinetics of the MB, MG and ARS dye decomposition

with barium titanate sample (BT10, BT20, and BT30 samples) is shown in Fig. 11(a), 11(b) and

11(c). The higher photocatalytic degradation efficiency was found in BT10 sample: similarly,

higher rate constant values were obtained for BT10 sample. Figure 11(a) shows the kinetic values

for samples BT10, BT20, and BT30 for degradation of MB dye and the values are 0.006120 min-1,

0.006041 min-1 and 0.004247 min-1, respectively. Figure 11(b) shows the kinetic values for BT10,

BT20, and BT30 samples and the kinetic values for MG dye degradation are 0.01953 min-1,

0.01316 min-1 and 0.01105 min-1, respectively (Fig.11b). Similarly, Figure 11(c) shows the kinetic

values for BT10, BT20, and BT30 for ARS dye degradation and the values are about 0.005352

min-1, 0.004557 min-1 and 0.002156 min-1, respectively. It is clear that the rate constant of BT10

was greater than that of BT20 and BT30 which were calculated from kinetic study and this was

well matched with photocatalytic decomposition study.

5. Conclusions
In summary, the whole synthetic procedure is carried out using a domestic microwave

oven, which opens up a new route to scale up the preparation of BaTiO3 nanoparticles. The

structural, optical and morphological characterization was carried out for the synthesized BaTiO 3

nanomaterials. The photocatalytic dye degradation studies and kinetic studies were carried using

developed BaTiO3 nanomaterials. The highest decomposition efficiency of 100 % was achieved

using 10 min microwave irradiated BaTiO3 (BT10) nanomaterials. There is no characteristic new

band was observed in the UV-visible spectroscopic study of MB, MG and ARS dyes, reveals that

the dyes were directly decayed into harmless CO2 and water. The band gap energy of synthesized

BT10, BT20 and BT30 photocatalyst were found to be 2.32 eV, 2.71 eV and 3.01 eV respectively,

which are calculated by UV-DRS spectra using Kubelka-Munk equation. BT10 photocatalyst have

lower band gap energy compared to BT20 and BT30; hence, it shows good photocatalytic activity

and highest first order rate constant value.

Acknowledgements

The authors are grateful to the Department of Physics and Chemistry, Alagappa

University for providing the characterization facilities.

References

[1] S. Karuppuchamy, N. Suzuki, S. Ito, T. Endo, A novel one-step electrochemical

method to obtain crystalline titanium dioxide films at low temperature, Curr. Appl. Phys.

9 (2009) 243-248.

[2] N. Okada, S. Karuppuchamy, M. Kurihara, An efficient dye-sensitized

photoelectrochemical solar cell made from CaCO3-coated TiO2 nanoporous film, Chem.

Lett. 34 (1) (2005) 16-17.


[3] H. Matsui, T. Okajima, S. Karuppuchamy, M. Yoshihara, The electronic behavior of

V2O3/TiO2/carbon clusters composite materials obtained by the calcination of a

V(acac)3/TiO(acac)2/polyacrylic acid complex, J. Alloy. Compd. 468 (2009) 27-32.

[4] H. Matsui, S. Nagano, S. Karuppuchamy, M. Yoshihara, Synthesis and

characterization of TiO2/MoO3/carbon clusters composite material, Curr. Appl. Phys. 9

(2009) 561-566.

[5] H. Matsui, Y. Saitou, S. Karuppuchamy, M.A. Hassan, M. Yoshihara, Photo-

electronic behavior of Cu2O-and/or CeO2-loaded TiO2/carbon cluster nanocomposite

materials, J. Alloy. Compd. 538 (2012) 177-182.

[6] N. Suzuki, S. Karuppuchamy, S. Ito, Uniform coating of a crystalline TiO2 film onto

steel plates by electrochemical deposition using staged pulse current, J. Appl.

Electrochem. 39 (2009) 141-146.

[7] H. Matsui, S. Yamamoto, Y. Izawa, S. Karuppuchamy, M. Yoshihara, Electron

transfer behavior of calcined material obtained from a samarium-O–phenylene-S–nickel-

S–phenylene-O hybrid copolymer, Mater. Chem. Phys. 103 (2007) 127-131.

[8] R.D. Kumar, S. Karuppuchamy, Facile synthesis of honeycomb structured SnO/SnO 2

nanocomposites by microwave irradiation method, J. Mater. Sci. Mater. Electron. 26

(2015) 6439-6443.

[9] S. Karuppuchamy, Y. Andou, T. Endo, Preparation of Nanostructured TiO2 for

Flexible Dye sensitized Solar Cell Applications, Appl. Nanosci. 3 (2013) 291-293.

[10] M. Nishikawa, S. Hiura, Y. Mitani, Y. Nosaka, Enhanced photocatalytic activity of

BiVO4 by co-grafting of metal ions and combining with CuBi2O4, J. Photochem.

Photobiol. A 262 (2013) 52– 56.


[11] C.P. Ireland, C. Ducati, Investigating the photo-oxidation of model indoor air

pollutants using field asymmetric ion mobility spectrometry, J. Photochem. Photobiol. A

312 (2015) 1–7.

[12] M.T. Buscaglia, M. Viviani, Z. Zhao, V. Buscaglia, P. Nanni, Synthesis of BaTiO3

Core-Shell Particles and Fabrication of Dielectric Ceramics with Local Graded Structure,

Chem. Mater. 18 (2006) 4002-4010.

[13] C. Jiang, K. Kiyofumi, Y. Wang, K. Koumoto, Synthesis of BaTiO3 Nanowires at

Low Temperature, Crystal. Growth. Design. 7 (2007) 2713-2715.

[14] J.J. Urban, W.S. Yun, Q. Gu, H. Park, Synthesis of Single-Crystalline Perovskite

Nanorods Composed of Barium Titanate and Strontium Titanate, J. Am. Chem. Soc. 124,

(2002) 1186-1187.

[15] Y. Ma, E. Vileno, S.L. Suib, P.K. Dutta, Synthesis of Tetragonal BaTiO3 by

Microwave Heating and Conventional Heating, Chem. Mater. 9 (1997) 3023-3031.

[16] S.O. Brien, L. Brus, C.B. Murray, Synthesis of Monodisperse Nanoparticles of

Barium Titanate: Toward a Generalized Strategy of Oxide Nanoparticle Synthesis, J. Am.

Chem. Soc. 123 (2001) 12085-12086.

[17] Q. Feng, M. Hirasawa, K. Yanagisawa, Synthesis of Crystal-Axis-Oriented BaTiO3

and Anatase Platelike Particles by a Hydrothermal Soft Chemical Process, Chem. Mater.

13 (2001) 290-296.

[18] M.T. Buscaglia, V. Buscaglia, R. Alessio, Coating of BaCO3 Crystals with TiO2:

Versatile Approach to the Synthesis of BaTiO3 Tetragonal Nanoparticles, Chem. Mater.

19 (2007) 711-718.

[19] Z. Zhou, H. Tang, H.A. Sodano, Vertically Aligned Arrays of BaTiO3 Nanowires,

ACS Appl. Mater. Interface. 5 (2013) 11894−11899.


[20] Y. Mao, H. Zhou, S.S. Wong, Synthesis, Properties, and Applications of Perovskite-

Phase Metal Oxide Nanostructures, Mater. Matter. 5.2 (2010) 50-57.

[21] H. Matsui, T. Kuroda, K. Otsuki, K. Yokoyama, T. Kawahara, S. Karuppuchamy, M.

Yoshihara, Electronic behavior of calcined material from a tellurium-S-phenylene-O-

strontium-O-phenylene-S hybrid copolymer, Tanso. 222 (2006) 114-117.

[22] R.D. Kumar, S. Karuppuchamy, Microwave-assisted synthesis of copper tungstate

nanopowder for supercapacitor applications, Ceram. Int. 40 (2014) 12397-12402.

[23] H. Miyazaki, H. Matsui, Y. Kita, S. Karuppuchamy, S. Ito, M. Yoshihara, Electronic

behavior of visible light sensitive ZrO2/Cr2O3/carbon clusters composite materials, Curr.

Appl. Phys. 9 (2009) 155-160.

[24] M. Thamima, S. Karuppuchamy, Biosynthesis of Titanium Dioxide and Zinc Oxide

Nanoparticles from Natural Sources: A Review, Adv. Sci. Eng. Med. 7 (2015) 18-25.

[25] R.D. Kumar and S. Karuppuchamy, Synthesis and characterization of nanostructured

Zn-WO3 and ZnWO4 by simple solution growth technique, J. Mater. Sci. Mater.

Electron. 26 (2015) 3256-3261.

[26] S. Karuppuchamy, Y. Andou, T. Endo, Preparation of nanostructured TiO2

photoelectrode for flexible dye-sensitized solar cell applications, Appl. Nanosci. 3 (2013)

291-293.

[27] H. Matsui, N. Bandou, S. Karuppuchamy, M.A. Hassan, M. Yoshihara, Efficient

photocatalytic activity of MnO2-loaded ZrO2/carbon cluster nanocomposite materials

under visible light irradiation, Ceram. Int. 38 (2012) 1605-1610.


[28] K. Santhi, P. Manikandan, C. Rani, S. Karuppuchamy, Synthesis of nanocrystalline

titanium dioxide for photodegradation treatment of remazol brown dye, Appl. Nanosci. 5

(2015) 373-378.

[29] H. Miyazaki, H. Matsui, T. Kuwamoto, S. Ito, S. Karuppuchamy, M. Yoshihara,

Synthesis and photocatalytic activities of MnO2-loaded Nb2O5/carbon clusters composite

material, Microporous. Mesoporous Mater. 118 (2009) 518-522.

[30] M. Thamima, S. Karuppuchamy, Synthesis, characterization and photocatalytic

properties of rod-shaped titanium dioxide, J. Mater. Sci. Mater. Electron. DOI:

10.1007/s10854-015-3774.

[31] M. Thamima and S. Karuppuchamy, Microwave Assisted Synthesis of Zinc Oxide

Nanoparticles, Inter. J. ChemTech. Res. 8 (2015) 250-256.

[32] A.S. Dzunuzovic, M.M.V. Petrovic, B.S. Stojadinovic, N.I. Ilic, J.D. Bobic, C.R.

Foschini, M.A. Zaghete, B.D. Stojanovic, Multiferroic (NiZn) Fe2O4-BaTiO3 composites

prepared from nanopowders by auto-combustion method, Ceram. Int.

http://dx.doi.org/10.1016/j.ceramint.2015.07.096.

[33] P. Chen, Y. Zhang, F. Zhao, H. Gao, X. Chen, Z. An, Facile microwave synthesis

and photocatalytic activity of monodispersed BaTiO3 nanocuboids, Mater. Charact. 114

(2016) 243–253.

[34] J.S.J. Zhang, Comparison of rheological, mechanical, electrical properties of HDPE

filled with BaTiO3 with different polar surface tension, Appl. Sur. Sci.

http://dx.doi.org/doi:10.1016/j.apsusc.2015.10.156.
[35] R. Ashiri, Detailed FT-IR spectroscopy characterization and thermal analysis of

synthesis of barium titanate nanoscale particles through a newly developed process,

Vibrational. Spectros. 66 (2013) 24– 29.

[36] M. L. Moreira, G.P. Mambrini, D.P. Volanti, E.R. Leite, M.O. Orlandi, P.S. Pizani,

V.R. Mastelaro, C.O.P. Santos, E. Longo, J.A. Varela, Hydrothermal Microwave: A New

Route to Obtain Photoluminescent Crystalline BaTiO3 Nanoparticles, Chem. Mater. 20

(2008) 5381–5387.

[37] Z. Lazarevi, N. Romcevi, M. Vijatovi, N. Paunovi, M. Romcevi, B. Stojanovi, Z.

Mitrovi, Characterization of Barium Titanate Ceramic Powders by Raman Spectroscopy,

Acta physica. Polonica. A. 115 (2009) 808-810.

[38] Y. Yang, J. Shi, S. Dai, X. Zhao, X. Wang, High third-order non-resonant optical

non-linearities of surface modified CdS quantum dots embedded in BaTiO3 thin film,

Thin Solid Film. 437 (2003) 217–222.


Scheme 1. Schematic representation of photocatalytic reaction mechanism of BaTiO3

materials

Fig. 1 XRD patterns of Prepared BaTiO3 samples (a) BT10 (b) BT20 and (c) BT30

Fig. 2 FT-IR spectrum of synthesized BaTiO3 samples (a) BT10 (b) BT20 and (c) BT30

Fig. 3 Raman spectra of microwave irradiated BaTiO3 Nanomaterials (a) BT10 (b) BT20 and

(c) BT30

Fig. 4 SEM images of microwave irradiated for 10 min and 30 min BaTiO 3 nanomaterials (a)
BT10 at 10 m (b) BT10 at 1m, (c) BT30 at 10 m and (d) BT30 at 1m
Fig. 5 EDX spectra of microwave irradiated for 10 min and 30 min BaTiO3 nanomaterials (a)
BT10 (b) BT30
Fig. 6 UV-vis DRS Absorbance spectra of microwave irradiated for 10, 20 and 30 min
BaTiO3.
Fig. 7 UV–vis DRS spectra of the Kubelka–Munk plot with corresponding band gap energy
of BT10 (b) BT20 and (c) BT30 samples
Fig. 8 Photocatalytic degradation of (a) Methylene blue dye (b) Malachite green dye and (c)
Alizarin Red S dye
Fig. 9 UV-vis spectra of Methylene blue dye with 0.1 g of BaTiO3 (a) BT10 (b) BT20 and (c)
BT30
Fig. 10 UV-vis spectra of (a) MG dye and (b) ARS dye
Fig. 11 Kinetic studies of (a) MB dye (b) MG dye and (c) ARS dye.
Table1. EDX spectra of Barium titanate (BT10) nanomaterials

Elements (K) wt% At%

CK 11.86 21.39

OK 16.11 22.02

Na K 11.21 15.51

SK 09.21 09.13

Ba L 41.60 11.95

Ti L 10.03 21.39
Table2. EDX spectra of Barium titanate (BT20) nanomaterials

Elements wt% At%


(K)
CK 23.77 40.87

OK 17.84 26.45

Na K 04.27 04.58

SK 05.79 04.45

Ba L 41.37 07.42

Ti L 11.73 21.39
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Scheme 1

You might also like