You are on page 1of 13

Available online at www.sciencedirect.

com

ScienceDirect
Materials Today: Proceedings 9 (2019) 468–480 www.materialstoday.com/proceedings

ICNAN 2016

Controlled Hydrothermal Synthesis of Bismuth


Vanadate Nano-articulate Structures: Photooxidation
of Methicillin Resistant Staphylococcus aureus and
Organic Dyes
C. S. Vicas1,2, K. Namratha2, M. B. Nayan3 and K. Byrappa1,2*
1
Department of Materials Science, Mangalore University, Mangalagangotri, Mangalore, India.
2
Center for Materials Science and Technology, Vijnana Bhavan, University of Mysore, Mysore, India.
3
Department of studies in Environmental Science, University of Mysore, Mysore, India.

Abstract

Scheelite monoclinic bismuth vanadate nano-articulate crystals have been synthesized by using 1% sodium
dodecyl sulphate as surfactant and ammonia as pH controlling agent at 180⁰C for 4 hr (at pH = 0, 1, 3, 6 and 8)
under facile hydrothermal conditions. The products were characterized by X-ray diffraction (XRD), Fourier
transform infrared spectroscopy (FTIR), UV-Vis spectroscopy, field emission scanning electron microscopy
(FE-SEM) and energy dispersive X-ray spectroscopy (EDS). XRD and FE-SEM confirm the formation of
monoclinic structures at all pH ranges and a preferential crystal growth (crystallite size 20-30 nm) has been
observed along {040} facet with increase in NH4OH concentration. This selective growth resulted in structure
enhanced performance of photocatalytic degradation of organic dyes and bacterial inactivation. Photocatalytic
inactivation of methicillin resistant Staphylococcus aureus (MRSA) and comparison of the results to the
oxidation of Procion Red MX-5B, Alizarin Red S and Cibacron Brilliant Yellow 3G-P dyes using bismuth
vanadates has been reported originally in this paper.

© 2019 Elsevier Ltd. All rights reserved.


Selection and/or Peer-review under responsibility of International Conference on Nanoscience and
Nanotechnology (ICNAN’16).

Keywords: Nano-articulate crystals; MRSA inactivation; Preferential growth; Catalytic properties.

1. Introduction

Photocatalytic inactivation of methicillin resistant Staphylococcus aureus (MRSA) using


bismuth vanadate nanostructures and comparison of the results to the oxidation of dyes has been
reported for the first time in this paper. Photocatalytic obliteration of pathogens has apprehended
the attention of researchers for several environmental applications as the current disinfection
* Corresponding author. Tel.: +91 72596 67666
E-mail address: kbyrappa@gmail.com

2214-7853 © 2019 Elsevier Ltd. All rights reserved.


Selection and/or Peer-review under responsibility of International Conference on Nanoscience and Nanotechnology
(ICNAN’16).
C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480 469

processes involve the production of by-products that are harmful and in cases can cause cancer.1
Titanium dioxide (TiO2) is widely studied and is most popular among photocatalysts that are being
used. It is also known to be nontoxic with high oxidation capacity and chemical stability.2
Henceforth it is extensively used in creams and cosmetics.3 A major drawback with respect to TiO2
is its wide bandgap, which facilitates it to respond only in ultraviolet light, i.e., about 4% of total
solar energy.4,5 Thus, visible light photocatalysts like bismuth vanadate (BiVO4) have come to
limelight. BiVO4 is widely studied due to their lower bandgap of 2.4 eV.6 Photocatalytic property
of BiVO4 is identified to vary with their crystal structure. They exist in tetragonal zircon,
monoclinic and tetragonal scheelite structures, among which the most active phase under visible
light is known to be monoclinic scheelite structure.7-9
Preparation of BiVO4 of certain morphology and expression of particular facets, for
enhanced photocatalytic performance has been a challenging endeavour to researchers.10-12 It is well
understood from the previous researches that the expression of {040} facet of bismuth vanadate
increases its photocatalytic action. A few reports have depicted the usage of TiCl3, ethanolamine
and NaHCO3 so on, for increase in {040} facet.13,14 Usage of green shape-promoting agents were
also reported.15 In the present work, the preparation, characterization and photocatalytic activity of
BiVO4 photocatalysts synthesized by hydrothermal method at pH 0, 1, 3, 6 and 8 without addition
of any such structure enhancing agents has been reported.
The synthetic routes available for the synthesis of BiVO4 include majorly sol–gel method,
hydrolysis of metal alkoxides, hydrothermal method, etc.16 Among these, hydrothermal route is
known to be most effective for processing advanced functional materials as it facilitates control
over size and morphology of the nanoparticles and does not involve in post synthesis treatments.17
Methicillin-resistant Staphylococcus aureus (MRSA, MTCC - 1430) is a gram-positive cocci that is
a major pathogen that causes dental caries, skin and soft tissue infections (SSTIs), lung infection
and so on. It is an opportunistic pathogen and statistically two thirds of Staphylococcus in domestic
water is S. Aureus.18,19 Photocatalytic activity of BiVO4 crystals against S. aureus has been
thoroughly investigated and compared with the degradation of textile dyes like Procion Red MX-
5B and Cibacron Brilliant Yellow 3G-P and Alizarin Red S, a histochemical stain.

2. Materials and method

2.1. Synthesis of BiVO4 crystals

Thalluri et al. (2014) synthesized BiVO4 at different pH ranges using ammonium


carbonate as promoting agent for preferential growth along {040} facet.20 In the current research no
shape promoting agents have been used in the preparation process. Synthesis of scheelite
monoclinic bismuth vanadates was carried out using 4M bismuth nitrate (Bi(NO3)3·5H2O, Rankem)
and 4M ammonium vanadate (NH4VO3, Rankem), dissolved in 50 ml of 2M nitric acid (HNO3,
Rankem) and 2M ammonia (NH4OH, Rankem) respectively as shown in Fig. 1. These solutions
were then mixed together under constant stirring for 30 min at room temperature to form a clear
yellow precipitate.21 It was then transferred into five Teflon liners (20 ml in each) in general
purpose autoclaves (SS 140) and their pH was adjusted to 0, 1, 3, 6 and 8 respectively, by adding
drops of conc. NH4OH. An amount of 20µL of 1% sodium dodecyl sulphate (SDS) aqueous
solution was added as a surfactant into each liner, prior to the treatment. Autoclaves were placed at
180⁰C (autogenous pressure) for 4 hr (Thermotek® Hot-air Oven, India). The resultant products
were washed repeatedly using ultrapure water (Purelab® Option Q7), filtered and dried overnight.
470 C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480

Fig. 1. Synthesis flow chart

2.2. Materials characterization

A series of characterization techniques were performed, such as X-ray diffraction


(Rigaku® Mini-Flex II, Japan) for analysing the phase purity of the materials, Furier transformation
infrared spectroscopy (Jasco®, Japan) for identifying the functional groups, UV-Vis spectroscopy
(Elico®, India) for obtaining absorption wavelength, scanning electron microscopy (Hitachi® S3400
N, Japan) and field emission scanning electron microscopy (Quanta 200 FEG) for morphological
studies and energy-dispersive X-ray spectroscopy (Thermo-scientific® UltraDry EDS detector,
USA) for elemental analysis.

2.3. Photocatalytic experiments

2.3.a. Dye degradation tests

Dye degradation experiments were performed in two sets, under visible light (solar
radiation) for 3 hr.22 Degradation ability of the synthesized photocatalysts was tested on Procion
Red MX-5B (PRM, λmax=538 nm), Alizarin Red S (ARS, λmax=450 nm) and Cibacron Brilliant
Yellow 3G-P (CBY, λmax=404 nm) dyes that were procured from Sigma Aldrich. One set of
experiments were performed by keeping the concentration of the dyes constant, i.e., 8 ppm and by
varying the catalysts concentration as 0.02, 0.04, 0.06, 0.08 and 0.1 g/L. Another set of experiments
were performed keeping the catalysts concentration constant, i.e., 0.1 g/L and varying the dyes
concentration as 8,16, 24 and 32 ppm. Percentage transmittance for the dyes were taken at every
half-an-hour and at the end of 3 hr, chemical oxygen demand (incubated in Hach® COD reactor)
was estimated and graphs were plotted.

2.3.b. Bacterial inactivation test

Fresh overnight culture of Staphylococcus aureus (gram-positive facultative anaerobe,


MTCC 6908) were prepared in Luria-Bertani broth (HIMEDIA) and the cell suspension was diluted
and adjusted to McFarland’s standard 0.5. This cell suspension was collected into 5 sterilized flat
bottom culture tubes (30 ml each, Borosil®) aseptically in laminar air flow chamber and the
C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480 471

prescribed concentrations (0.1 g/L) of the photocatalysts were added. Two additional tubes were
taken along with these. One for positive control, in which bacterial culture was taken without
adding inhibitory agents and the other for negative control, in which bacterial culture was taken and
0.5 ml of Pen-Strep (ThermoFisher Scientific) antibiotic was added.
Culture tubes were capped and placed in the sunlight for ten hours in a shaker. For every
hour, absorbance readings were taken at 580-600 nm and inhibition percentage was calculated
using the following formula Eqs. (1).23,24


= 100

Where,
Pg – percentage growth,
S – absorbance of the sample with inoculum,
Sb – absorbance of the sample blank,
Cp – absorbance of the positive control,
Cn – absorbance of the negative control.

Percentage inhibition (Pi) can be calculated from Eqs. (2):

Pi 100 Pg

3. Results and discussion

3.1. Characterization

XRD spectra of the synthesized bismuth vanadates (Fig. 2) show the formation of well
crystalline scheelite monoclinic structures, at all the pH values of the reaction mixture. The graphs
obtained are in good agreement with JCPDS card No. 14-0688 as shown in Fig. 2, which
corresponds to I2/a space group of monoclinic bismuth vanadates with a= 5.195, b=11.70, c=5.092
and α=γ≠90⁰, β=90⁰. Although, a notable difference was observed in {040} peaks in spectra with a
change in the pH of the solution. It was evident that the synthesized crystals exhibited a preferential
growth along {040} planes compared to {121} peaks with increase in pH due to the addition of
NH4OH. This gives a control over nucleation kinetics without any additional structural changing
agents.25,26
The values of 2θ, Muller’s indices, full width half maximum, calculated crystallite size
were tabulated in Table. 1.
Crystallite size of the particles was calculated using the following formula (Eqs. 2).
d - spacing was calculated using Bragg’s law (Eqs. 3).

n λ=2dhkl sinθ
Where,
D = crystallite size
= wavelength of source (Cu = 1.544)
β = full width half maximum (FWHM)
θ = Bragg’s angle
n = order of the plane
d = d – space
hkl = Muller’s indices
472 C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480

Fig. 2. XRD spectra of the synthesized Bismuth vanadate


FE-SEM and SEM micrographs (Fig. 3a-e) show distinct morphological changes in
evidence to the XRD spectra. Samples exhibited perfect monoclinic shape at pH 0 (Fig. 3a1-a2). At
pH 1 the sample showed a slight distortion, yet maintained the monoclinic structure (Fig. 3b1-b2).
As the pH increased to 3, formation of granular agglomerates was observed (Fig. 3c1-c2). With the
increase in pH to 6, the synthesized material showed irregular shapes like that of cheese-puffs (Fig.
3d1-d2). At pH 8 the sample exhibited preferential growth in {040} plane forming beautiful lobate-
leaf like structures (Fig. 3e1-e4). Increase in pH during the synthesis using NH4OH has shown a
profound effect on the shape of the particles, by enhancing the probability of increase in V4+ and O-
H species, which was evident from EDS spectra shown in Fig. 4., thus exhibiting a preferential
growth of {040} facet.27,28 Based on these results, we come to an understanding that an increase in
pH expedites the expression of {040} facet, even without the addition of any structure enhancing
agents. This in turn helps in photodegradation as O-H groups act as precursors of hydroxyl radicals
and also facilitate the adsorption of the dyes on the surface of the catalyst more effectively.29 Apart
from atomic percentages, the purity of the synthesized materials was also confirmed by EDS

(Fig.Fig. 3. FE-SEM and SEM analysis of synthesized BiVO4- (a1-2) at pH 0; (b1-b2) at pH 1; (c1-c2) at pH 3; (d1-d2) at
pH 6; (e1-e4) at pH 8
C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480 473

4). The spectra show that the synthesized materials were without any impurities.

Fig. 4 (a, b) Elemental analysis – EDS spectra showing variation in atomic percentages in BiVO4 synthesized at
pH 0 and pH 8

FTIR spectra (Fig. 4) of the synthesised BiVO4 samples show a band around 730 – 1000 cm-1 was
attributed to the stretching mode of VO3- 4 and its branch at 613 cm-1 attributes to Bi-O.30 These
correspond to characteristic peaks of monoclinic bismuth vanadate, according to Liu et al. (2003)
and Gotic et al. (2005).31,32 With the increase in pH, Bi-O branch was observed to be diminishing
gradually and V-O becoming more prominent. This is in correspondence with XRD spectra which
shows an increase in {040} facet with pH and an increase atomic percentages of V and O as
observed from EDS graph. The peak observed at 1386 cm-1 was assigned to stretching mode of C-H
group and the one at 1760cm-1 attributes to C=O stretch from SDS which was used as the
surfactant.

Fig. 5. FTIR spectra of bismuth vanadate synthesized at different pH ranges


474 C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480

UV-Vis absorption spectrometry (Fig. 6) was studied for BiVO4 crystals that were
synthesized at different pH ranges. The obtained materials showed a characteristic absorption edge
at around 550 nm, which correlates with n to π* transition that correspond to the excitation of an
electron from one of the unshared pair to the π* orbital.33 Thus, calculated band gap value in the
range of 2.3-2.4 eV which is said to be clearly inside the visible range of the spectrum.
Calculation of band gap energy is as shown below (Eqs. 3) and the obtained band gap energies
were tabulated (Table. 1).

Band Gap Energy (E) = h*C/λ

Where,
h – Plank’s constant = 6.626 x 10-34 Joules sec
C – Velocity of light = 3.0 x 108 meter/sec
λ – Wavelength
Conversion factor 1eV = 1.6 X 10-19 Joules

Table. 1 – Shows 2θ values, {hkl} indices, full width half maximum, calculated crystallite size &
bandgap energies and morphology of the synthesize materials

Catalyst 2θ FWHM d - spacing Crystallite Morphology Band gap


synthesized (hkl) size (nm) (eV)

at pH 0 28.6 0.3874 3.1244 22.55 Monoclinic 2.32


{121} crystals

at pH 1 28.6 0.4190 3.1210 20.85 Distorted 2.34


{121} monoclinic

at pH 3 28.6 0.3569 3.1240 22.58 Granular 2.35


{121} agglomerates

at pH 6 28.6 0.379 3.1211 22.64 Cheese puff 2.38


{121} shapes

at pH 8 28.6 0.3678 3.1245 23.32 Lobate leaf-like 2.39


{121}
C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480 475

Fig. 6. UV-Vis spectra of bismuth vanadate synthesized at different pH ranges

3.2. Photocatalysis experiments

3.2.a. Dye degradation

Photocatalytic degradation for the selected dyes was performed for a duration of 3 hr in visible light
(sun light) initially by keeping the dyes concentration as constant (8 ppm) and varying the catalyst
concentration as 0.02, 0.04, 0.06, 0.08 and 0.1 g/L. Second set of experiments was performed
keeping the catalyst concentration as constant (0.1 g/L) and varying the dyes concentration as 8, 16,
24, 32 ppm. Percentage transmittance values were plotted (Fig. 8-10) and COD was estimated and
the obtained results were tabulated (Table. 2 and 3).
The results obtained confirm that all the as-prepared BiVO4 samples showed good
degradation capability against the selected dyes of 8, 16 and 24 ppm concentrations. When the dye
concentration was raise to 32 ppm, the degradation ability diminished. It is inferred that a
maximum fraction of degradation occurred with ARS dye and the least in case of PRM. A clear
enhancement of photodegradation property has been observed, when synthesized at pH 8. Lowest
COD was estimated as 19.45 mg/L against 8 ppm of ARS dye when 0.1 g/L of BiVO4 synthesized
at pH 8 was used, which corresponds to the percentage transmittance of 93.34%.
It was evident from EDS spectra that excess addition of NH4OH enhances the presence of
O-H species on the surface of the particles, which not only help in adsorption of dyes on the surface
but also act as precursors for hydroxyl radicals. Moreover, the preferential growth of {040} facet
provides a wider surface for the reaction mechanism to take place. Mechanism of photodegradation
is well understood with respect to this semiconductor. Adsorption of dyes on the surface of the
photocatalyst occurs due to the presence of excess O-H groups.34,35 These are later converted to O-
H radicals, which play a key role in degradation. Further O-H radicals react with water molecules
and produce additional radicals. Although bandgap energies of the synthesized materials are almost
similar, due to the growth of {040} facet and formation of lobate leaf-like structures at higher pH,
clarifies a structure enhanced activity when compared with the ones synthesized at lower pH [18].
Mechanism involved is illustrated in Fig. 7 (b).
476 C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480

Table. 2 – Effect of catalysts concentration on degradation of selected dyes


Dye concentration (8 ppm)

PRM ARS CBY

Catalyst (g/L) COD (mg/L) %T COD (mg/L) %T COD (mg/L) %T


pH 0
0.02 g/L 238.52 7.43 173.18 32.65 169.17 32.34
0.04 g/L 156.98 21.45 139.29 36.17 143.43 41.54
0.06 g/L 112.36 39.07 98.45 43.56 112.45 45.67
0.08 g/L 88.44 58.34 65.87 59.77 74.34 52.12
0.1 g/L 52.76 73.12 36.45 86.45 41.65 79.54
pH 1
0.02 g/L 237.89 9.13 162.11 34.13 156.47 35.76
0.04 g/L 139.34 25.55 113.14 46.17 137.33 42.38
0.06 g/L 95.33 56.75 65.46 61.03 102.54 53.33
0.08 g/L 71.54 64.56 57.88 71.87 70.81 61.26
0.1 g/L 44.75 76.57 36.57 87.77 38.56 83.45
pH 3
0.02 g/L 225.46 12.56 145.55 38.65 138.87 34.45
0.04 g/L 120.17 42.19 92.49 58.34 125.45 37.23
0.06 g/L 72.55 63.13 54.72 67.16 97.13 45.77
0.08 g/L 49.43 81.65 43.87 79.36 64.83 64.22
0.1 g/L 38.92 82.66 32.54 88.34 32.14 85.35
pH 6
0.02 g/L 216.23 16.75 122.76 43.44 134.32 37.56
0.04 g/L 108.73 53.34 90.18 52.54 119.54 44.33
0.06 g/L 67.44 69.66 49.66 76.34 89.54 56.26
0.08 g/L 44.41 78.87 33.34 82.87 63.27 66.76
0.1 g/L 32.28 86.24 21.54 90.45 29.51 86.23
pH 8
0.02 g/L 178.13 31.54 103.46 47.76 123.77 42.45
0.04 g/L 97.87 58.06 82.16 65.54 105.67 49.66
0.06 g/L 50.56 72.23 47.33 78.12 60.33 67.18
0.08 g/L 36.64 81.17 29.78 86.54 41.54 73.13
0.1 g/L 27.12 87.68 19.45 93.34 24.87 89.43

Table. 3 – Effect of dyes concentration on degradation ability of the catalysts


Catalyst (0.1 g/L)

pH 0 pH 1 pH 3 pH 6 pH 8

Dye COD %T COD %T COD %T COD %T COD %T


(ppm) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L)
8 52.76 73.12 44.75 76.57 38.92 82.66 32.28 86.24 27.12 87.68
PRM 16 107.23 54.39 93.76 52.67 85.37 66.39 78.17 66.03 76.65 64.05
24 199.56 24.34 164.44 27.76 150.64 32.19 146.67 43.76 123.33 46.12
32 452.69 3.54 436.34 4.57 421.87 4.33 398.98 3.98 270.54 8.78
8 36.45 86.45 36.57 87.77 32.54 88.34 21.54 90.65 19.45 93.34
ARS 16 87.44 69.44 83.65 67.55 77.68 62.78 54.45 76.45 51.34 82.78
24 145.19 37.13 137.71 48.61 132.45 50.67 127.77 42.59 121.84 38.52
32 197.55 16.87 180.34 23.87 175.76 32.81 157.32 39.01 154.56 41.74
8 41.65 79.54 38.56 83.45 32.14 85.35 29.51 86.23 24.87 89.43
CBY 16 86.67 60.65 77.55 70.42 71.65 65.14 66.45 77.43 63.12 73.76
24 114.24 47.03 112.67 51.13 106.53 57.05 86.76 58.54 77.32 61.76
32 146.87 31.43 146.91 42.15 133.43 46.02 146.13 39.87 126.87 43.18
C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480 477

3.2.b. Bacterial inactivation studies

Photoxidation inactivation of MRSA using the synthesized semiconductors was studied by


the proposed 10 hr experiment (Fig. 11). It was evident from the results obtained that all the
synthesized samples showed a good activity against the growth of bacteria. It was also observed
that due to insufficiency of direct interaction between the particles surface and the bacteria, the
results were not spontaneous as in the case of dye degradation.36 Thus prolonged exposure to solar
radiation facilitated the formation of more O-H radicals which interact with the bacterial cell wall,
causing its disruption. However, a structure based enhancement in the activity was also observed in
the present scenario. The mechanism involved is illustrated in Fig. 7 (a). As proposed by Pablos et
al., an electrostatic repulsion between the negatively charged bacterial surface and the photocatalyst
surface would occur unless the intervention of an intermediate counter ion layer of cations takes
place.37 Thus, it was assumed that hydroxyl radical formation was hindered and hole transfer could
be favoured under visible light.38
Oxidative stress shoulders a major role in antibacterial activity causing cytotoxic effects.
Unlike the enhancement of the activity with respect to dye degradation, bacterial inhibition
performed by all the synthesized materials remained more or less similar, although excellent
inhibition rate of 94% has been witnessed in case of the material synthesized at pH 8. This marks
for a superior property of BiVO4 in inhibiting the growth of pathogenic bacteria, which can be
harnessed in real time applications.

Fig. 7. (a) Photooxidation mechanism of bacterial inactivation; (b) Photodegradation of organic dyes
478 C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480

Fig. 8. Effect of time (duration of photocatalysis) on the degradation of Procion Red MX-5B

Fig. 9. Effect of time (duration of photocatalysis) on the degradation of Alizarin Red S


C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480 479

Fig. 10. Effect of time (duration of photocatalysis) on the degradation of Cibacron Brilliant Yellow 3G-P

Fig. 11. Microbial inactivation of S. Aureus (MRSA)

4. Conclusion

Preferential growth of {040} facet of bismuth vanadate was achieved under hydrothermal
conditions, by varying the pH from 0-8 using ammonia and SDS as surfactant, without additional
structure modifying agents. It was understood from EDS spectra that higher pH ranges facilitate the
presence of V+4 and O-H groups, which enable this preferential growth. This was confirmed by
XRD spectra and FE-SEM micrographs and the purity of the materials by FTIR and EDS spectra.
Photocatalytic degradation of Procion Red MX-5B, Alizarin Red S and Cibacron Brilliant Yellow
3G-P was studied under solar radiation. A superior performance has been displayed by BiVO4
nano-lobate leaf-like structures that were formed at pH 8, when compared to that of the other
synthesized materials. This is due to increase in O-H groups on the surface, that help in adsorption
of the dye to the semiconductor surface and also act as precursors for O-H radicals. Nano-lobate
480 C. S. Vicas et al. / Materials Today: Proceedings 9 (2019) 468–480

crystals also showed a 94% of bacterial growth inhibition and is original being reported in this
paper. Even though the bandgap energies of all the synthesized materials remained more or less
same in the range of 2.3-2.4 eV, a structure based enhancement in the photodegradation activity
was observed. This outstanding dye degradation property and simultaneous nullification of
pathogenic bacteria using 0.1 g/L (or less) BiVO4 nanocrystals stands as a windfall in real time
applications.

Acknowledgement

The authors gratefully acknowledge the financial support given by University with
Potential for Excellence project grants from UGC, India. The authors acknowledge the help
received from the research scholar of our group, Mr. Kashinath Lellela and Mr. Abdo Hezam.

References

[1] H. Zhang, G. Chen, D.W. Bahnemann, J. Mater. Chem. 19 (2009) 5089-5121.


[2] P.S.M. Dunlop, J.A. Byrne, N. Manga, B.R. Eggins, J. Photochem. Photobiol. A: Chem. 148 (2002) 355-363.
[3] E. Gkika, P. Kormali, S. Antonaraki, D. Dimoticali, E. Papaconstantinou, A. Hiskia, Int. J. Photoenergy. 6 (2004) 227-231.
[4] C. Belver, C. Adán and M. Fernández-García, Catal. Today. 143 (2009) 274-281.
[5] C. Belver, C. Adán, S. García-Rodríguez, M. Fernández-García, Chem. Eng. J. 224 (2013) 24-31.
[6] A. Kubacka, M. Fernández-García, G. Colón, Chem. Rev. 112 (2012) 1555-1614.
[7] W.Z. Yin, W.Z. Wang, L. Zhou, S.M. Sun, L. Zhang, J. Hazard. Mater. 173 (2010) 194-199.
[8] X. Zhang, L. Dua, H. Wang, X. Dong, X. Zhang, C. Ma, H. Ma, Microporous Mesoporous Mater. 173 (2013) 175–180.
[9] J. Yu, A. Kudo, Adv. Funct. Mater. 16 (2006) 2163-2169.
[10] D. Wang, H. Jiang, X. Zong, Q. Xu, Y. Ma, G. Li, C. Li, Chem. Eur. J. 17 (2011) 1275-1282.
[11] G. Xi, J. Ye, Chem. Commun. 46 (2010) 1893−1895.
[12] J. Su, L. Guo, S. Yoriya, C.A. Grimes, Cryst. Growth Des. 10 (2010) 856−861.
[13] A. Zhang, J. Zhang, N. Cui, X. Tie, Y. An, L. Li, J. Mol. Catal. A-Chem. 304 (2009) 28−32.
[14] R. Li, F. Zhang, D. Wang, J. Yang, M. Li, J. Zhu, X. Zhou, H. Han, C. Li, Nat. Commun. 4 (2013) 1432.
[15] Y. Zhao, Y. Xie, X. Zhu, S. Yan, S. Wang, Chem. Eur. J. 14 (2008) 1601−1606.
[16] Y. Guo, X. Yang, F. Ma, K. Li, L. Xu, X. Yuan, Y. Guo, Appl. Surf. Sci. 256 (2010) 2215-2222.
[17] K. Byrappa, T. Adschiri, Prog. Cryst. Growth Charact. Mater. 53 (2007) 117-166.
[18] F.B. Caraciolo, M.A.V. Maciel, J.B. dos Santos, M.A. Rabelo, V. Magalhães, An. Bras. Dermatol. 87 (2012) 857-861.
[19] R.V. Rasmussen, V.G. Fowler, S. Jr Robert, N.E. Bruun, Future Microbiol. 6 (2012) 43-56.
[20] S.M. Thalluri, M. Hussain, G. Saracco, J. Barber, N. Russo, Ind. Eng. Chem. Res. 53 (2014) 2640-2646.
[21] S. Obregón, A. Caballero and G. Colón, Appl. Catal. B: Environ. 2012, 117, 59-66.
[22] K. Namratha, K. Byrappa, S. Byrappa, P. Venkateswarlu, D. Rajasekhar, B.K. Deepthi, J. Environ. Sci. 34 (2015) 248-255.
[23] C.S. Vicas, K. Namratha, K. Byrappa, H.S. Yathirajan, J. Chem. Pharma. Res. 7 (2015) 1114-1124.
[24] C.S. Vicas, K. Namratha, K. Byrappa, H.S. Yathirajan, J. Chem. Bio. Phy. Sci. Sec. A. 6 (2016) 105-116.
[25] K. Rajeshwar, N.R. de Tacconi, Chemical Society Reviews. 38 (2009) 1984-1998.
[26] Y. Zhao, Y. Xie, X. Zhu, S. Yan, S. Wang, Eur. J. 14 (2008) 1601-1606.
[27] M. Wang, Q. Liu, Y. Che, L. Zhang, D. Zhang, J. Alloy Compd. 548 (2013) 70-74.
[28] M. Long, W.M. Cai, J. Cai, B.X. Zhou, X.Y. Chai, Y.H. Wu, J. Phys. Chem. B. 110 (2006) 20211-20216.
[30] U.M.G. Pérez, S.S. Guzmán, A.M. de la Cruz, J. Peral, Int. J. Electrochem. Sci. 7 (2012) 9622-2632.
[31] J.B. Liu, H. Wang, S. Wang, H. Yan, Mater. Sci. Eng B. 104 (2003) 36-39.
[32] M. Gotic, M. Ivanda, M. Soufek, S. Popovic, J. Mol. Struct. 744 (2005) 535-540
[33] F. Stanculescu, A. Stanculescu, Nanoscale Res. Lett. 11 (2016) 87-99.
[34] O. Man, N. Haoyu, Z. Qin, Z. Shule, Z. Lei, Phys. Chem. Chem. Phys. 17 (2015) 28809-28817.
[35] K. Nagaveni, M.S. Hegde, N. Ravishankar, G.N. Subbanna, G. Madras, Langmuir. 20 (2004) 2900-2907.
[36] B. Xie, H. Zhang, P. Cai, R. Qiu, Y. Xiong, Chemosphere. 63 (2006) 956-963.
[37] C. Pablos, R. van Grieken, J. Marugán, I. Chowdhury, S.L. Walker, Catal. Today. 209 (2013) 140-146.
[38] W. Wang, Y. Yu, T. An, G. Li, H.Y. Yip, J.C. Yu, P.K. Wong, Environ. Sci. Technol. 46 (2012) 4599-4606.

You might also like