You are on page 1of 19

CHAPTER - V

PHOTOCATALYSIS OF ZnO NANORODS


5.1 Introduction
Zinc oxide, with a high surface reactivity owing to a large number of native defect
sites arising from oxygen nonstoichiometry, has emerged to be an efficient photocatalyst
material compared to other metal oxides [1-3]. ZnO exhibits comparatively higher
reaction and mineralization rates [4] and can generate hydroxyl ions more efficiently than
titanium oxide (TiO2) [5]. ZnO has been often considered a valid alternative to TiO2
because of its good optoelectronic, catalytic and photochemical properties along with its
low cost. ZnO has a band gap of 3.0 eV that is lower than that of anatase. Due to the
position of the valence band of ZnO, the photo generated holes have strong enough
oxidizing power to decompose most organic compounds [6]. ZnO has been tested to
decompose aqueous solutions of several dyes [7–9], and many other environmental
pollutants [10–12]. So the ZnO has been chosen as a catalyst for this experiment for the
photocatalytic studies. In many cases, ZnO has been reported to be more efficient than
TiO2 [13–14] but the occurrence of photo corrosion and the susceptibility of ZnO to
facile dissolution at extreme pH values, have significantly limited its application in
photocatalysis.
Studies of ZnO hexagonal nanorod structure represent a significant research area,
and it is reported that one dimensional (1D) ZnO nanorods arrays could enhance
photocatalytic efficiency [15]. One dimensional nanostructure, such as nano wires and
nanorods, offer higher surface to volume ratio compared to nanoparticulate coatings on a
flat plate [16] which increase the photocatalytic efficiency. Because of the ultra-high
surface area, ZnO nanowire or nanorod arrays could potentially be a very good class of
catalyst support structures. The effective surface area (adsorbed amount of target
molecules) and the diffusivity are important indices to gauge photocatalytic activity [17].
Surface area and surface defects play an important role in the photocatalytic activity of
metal-oxide nanostructures, as the contaminant molecules need to be adsorbed on to the
photocatalytic surface for the redox reactions to occur. The higher the effective surface
127

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


area, the higher will be the adsorption of target molecules leading to better photocatalytic
activity. ZnO nanorods, which can be strongly attached to any type of substrates through
proper surface treatment before seeding [18], are an attractive option for photocatalytic
applications. From the above findings, the ZnO nanorods have been taken as a catalyst
for the photo degradation of the textile reactive dye Methylene Blue (MB). The
performance of the reactor can also improve ZnO nanorod arrays on the inner wall of the
capillaries for forming high efficiency photocatalysis [19].
Several studies have shown that ZnO was quite active under UV illumination for
the photo degradation of some organic compounds in aqueous solution [20]. Methyl
Green was successfully decolorized and degraded by ZnO under visible light irradiation
at low watt irradiation [21] and the addition of an oxidant (H 2O2) enhanced the
degradation rate of the pollutant.
Lu et al. [22] used ZnO to degrade Basic Blue 11 under visible light irradiation
and studied the effects of influential factors like initial dye concentration, catalyst dosage,
and initial pH. Sobana and Swaminathan [23] increased the photocatalytic activity of
ZnO for the solar assisted photocatalytic degradation of Direct Blue 53 by mixing ZnO
and activated carbon at different proportions in an aqueous suspension. The synergistic
effect increased the efficiency of the photocatalyst by a factor of 4.21.
ZnO nanoparticles, prepared using zinc acetate and NaOH as precipitant, were
tested for the photo degradation of Biebrich scarlet in aqueous phase [24]. The
comparison with other commercial semi-conductors (TiO2, ZnO, CdS and ZnS) indicated
that the nano sized ZnO was the best photocatalyst for the decolorization of the dye.
Sakthivel et al. [25] studied the solar photo degradation of Acid Brown 14 as the
model pollutant to evaluate the performance of both ZnO and TiO2. The photo
degradation rate was determined for each experiment and the highest values were
observed for ZnO, suggesting that it absorbs large fraction of the solar spectrum and
absorbs more light quanta than TiO2.
The degradation process was effective at pH 7 and 10, but it was rather slow at pH
4. ZnO nanoparticles prepared from zinc acetate by triethylamine template assisted sol–

128

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


gel precipitation and further hydrothermal treatment exhibited high conversion values for
phenol photo oxidation [26].
Kislov et al. [27] showed that the photo activity and the photo stability of single
crystal ZnO samples strongly depended on the crystallographic orientation. High surface
area hexagonal ZnO nanoparticles demonstrated an enhanced photocatalytic degradation
of a tough pollutants compared with a commercial ZnO powder [28]. El-Kemary et al.
[29] synthesized ZnO nanoparticles by heating a mixture of zinc acetate dehydrate and
triethylamine in ethanol for 60 min at 50–60 ◦C. The photocatalytic activity for the
degradation of ciprofloxacin was investigated under UV light irradiation.
Kitture et al. [30] prepared polydispersed ZnO nanoparticles with two different
particle size distributions (120nm and 30 nm) that were tested for the degradation of MB
and Methelene Orange under UV irradiation. Shape and size selective ZnO nanorods with
high alignment and uniformity were grown by using a microwave-assisted chemical bath
deposition method on Indium Tin Oxide substrates [31].
The ZnO nanorods were efficient for the degradation of MB under UV irradiation
and exhibited a size-dependent activity. Hierarchically assembled porous ZnO spherical
nanoparticles showed a photo activity the degradation of phenol superior to that of TiO2
nanoparticles [32].
Li et al. [33] found that the photo reactivity of ZnO hollow spheres for the
degradation of reactive Brilliant Red X-3B increased by a factor 4.66 compared with that
of ZnO nanoparticles. Mohajerani et al. [34] synthesized ZnO nanostructures in the shape
of particle, rods, flower-like and microsphere that were tested for the decolorization of CI
Acid Red 27 under direct sunlight irradiation.
The photo activity of the ZnO nanorods was slightly superior to that of the
nanoparticles. The flower-like and microsphere 3D nanostructures showed much lower
photo activity. ZnO nano flowers were more efficient than ZnO nanorods for the
degradation of 4-chlorophenol under UV light irradiation [35].
The superior performance of the nano flowers resulted from the larger content of
oxygen vacancy on the surface of the 1D nano materials. Likewise, 3D flower-like ZnO

129

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


hierarchical microstructures prepared by a low-temperature aqueous solution route were
more active than other nano structured ZnO powders (nanoparticles, nano sheets, and
nanorods) [36].
Doping of metal oxides with metals and / or transition metals creates quasi-stable
energy states within the band gap (surface defects) [37], thereby affecting the optical and
electronic properties [38]. Increased electron trapping due to higher defect sites leads to
enhancement in the photocatalytic efficiency. This increase in photocatalytic efficiency is
possible provided the electron-hole pair recombination rate is lower than the rate of
electron transfer to adsorbed molecules. There are reports on the enhancement of visible
light absorption in ZnO by doping with, e.g., cobalt (Co) [39], manganese (Mn) [40], lead
(Pb) and silver (Ag) [37], etc. Photocatalytic activity comparable to doped ZnO was also
observed with engineered defects in ZnO crystals achieved by fast crystallization during
synthesis of the nanoparticles [41].

5.2 Experimental technique

Fig. 5.1 Photocatalytic experimental setup

Photo catalytic activity was carried out in a specially designed reactor in which the
light source was 8W UV lamp (Philips TUV-08 G5) shown in fig. 5.1. The wavelength
range and peak wavelength were determined to 225–265nm and 254 nm, respectively,
with an average intensity of 0.2mW/cm2, at the irradiation distance of 10 cm. All samples
to be irradiated were placed in a glass beaker of 2 cm inner diameter, at a volume of 20
ml. The hydrothermally grown ZnO nanorods were used as catalyst. 0.5 mol of MB dye

130

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


was taken in a beaker and ZnO nanorods were suspended into the beaker and treated with
UV lamp by varying irradiation time and area of catalyst. The absorption spectra were
recorded using JASCO V-570 UV-Vis spectrophotometer and rate of decolorization was
observed.

Fig. 5.2 Phillips TUV-08 G5 lamp

The UV lamp used for photocatalytic application in our study is Philips TUV-08
G5 lamp shown in fig. 5.2. Phillips is the largest manufacturer of standard low pressure
mercury lamps. These Philips TUV lamps consist of a tubular glass envelope emitting
short-wave ultraviolet (UV) radiation with a peak at 254 nm (UVC) for germicidal action.
The Philips in-house made glass filters out the 185 nm ozone forming line thus
preventing the creation of ozone. Low pressure mercury lamps are very efficient, up to
40%. A protective coating on the inside limits the depreciation of the useful UVC output.
This allows application manufacturers to design their systems to the highest efficiency.
Philips invented and pioneered the use of technology to reduce the mercury level of the
lamps. As a result this has been brought down to by far the lowest mercury level in UV
lamps in the industry. Main applications of the Philips TUV-08 G5 lamp are:
 Residential drinking water units
 Stand alone air purifiers
 Germicidal actions.
131

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


The measurement conditions of the Philips TUV-08 G5 lamp is given in table 5.1.
Table 5.1 Measurement conditions of Phillips TUV-08 G5 lamp

Instrument name Philips TUV-08 G5


Tube diameter 16 cm
Arc length 237 mm
Lamp wattage 7 watts
Lamp voltage 56 volts
Lamp current 0.15 A
UV C 100h 2.1 watts
mW/cm2 at 1 meter 21
Depreciation 9000h 25 %
Bf – Bf mm 288.3 A

The photo catalyzed decolorization of a dye in solution is initiated by the photo


excitation of the semiconductor, followed by the formation of electron–hole pair on the
surface of catalyst (Eq. 1). The high oxidative potential of the hole h +VB in the catalyst
permits the direct oxidation of the dye to reactive intermediates (Eq. 2).
(MO/MO2) + hυ → (MO/MO2) (e-CB + h-CB) (1)
h+VB + dye → dye●+ → Oxidation of dye (2)
Another reactive intermediate which is responsible for degradation is hydroxyl
radical (OH•). It is either formed by the decomposition of water (Eq.3) or by reaction of
the hole with OH-(Eq. 4).
h+VB + H2O → H+ + ●OH (3)
h+VB + OH- → ●OH (4)

OH + dye → degradation of the dye (5)
The hydroxyl radical is an extremely strong, non-selective oxidant (E o = +3.06 V),
which leads to the partial or complete mineralization of several organic chemicals [9].
The photocatalytic activity of the materials was investigated by degrading an organic
pollutant MB. MB is one of the generally accepted organic pollutants for degradation
studies and is used as an industrial standard.

132

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


Methylene Blue (CI 52015) is a heterocyclic aromatic chemical compound with
the molecular formula C16H18N3SCl. It has many uses in a range of different fields, such
as biology and chemistry. At room temperature it appears as a solid, odorless, dark green
powder, which yields a blue solution when dissolved in water. The hydrated form has 3
molecules of water per molecule of MB [42].
MB is widely used as a redox indicator in analytical chemistry. Solutions of this
substance are blue when in an oxidizing environment, but will turn colorless if exposed to
a reducing agent. The redox properties can be seen in a classical demonstration of
chemical kinetics in general chemistry, the "blue bottle" experiment. Typically, a solution
is made of glucose (dextrose), MB, and sodium hydroxide. Upon shaking the bottle,
oxygen oxidizes methylene blue, and the solution turns blue.
In this work a model textile reactive dye methelene blue is taken for the
photocatalytic degradation using grown ZnO nanorods. The structure and the properties
of the MB dye were given in the fig.5.3 and table 5.2 respectively.

Fig. 5.3 Structure of the methelene blue dye

Table 5.2 Properties of methylene blue dye

Dye name Methylene blue


Molecular formula C16H18N3SCl
Molar mass 319.85 g / mol
Melting point 100 - 110°C
Boiling point Decomposes

133

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


5.3 Results and discussion
The photo degradation of MB by undoped ZnO nanorods and ZnO nanorods
doped with aluminium, strontium and lithium under Phillips TUV-08 G5 lamp were
carried out. The efficiency of the dye degradation was reported by varying the irradiation
time and area of the catalyst. Extent of methylene blue degradation was estimated from
absorbance spectra as intact MB shows strong absorbance at 665 nm. Loss of intensity
and shift in this peak position was considered as degradation of methylene blue. The
percentage degradation (% D) was calculated using Equation (6).
Percentage of degradation = A 0 - At / A0 x100 (6)
Where A0 = absorbance at t = 0 minute
At = absorbance at t minute
The degradation of methylene blue dye was carried out using ZnO nanorods with
varying irradiation time, area of the catalyst and the effect of doping with aluminium,
strontium and lithium.

5.3.1 Effect of Irradiation time

2.0
B methelene blue
1.8 C ZnO-1 hour
D ZnO-2 hour
1.5 E ZnO-3 hour

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
450 500 550 600 650 700 750
Wavelength(nm)

Fig. 5.4 Time-dependent UV–Vis absorption spectra for decolorization of methylene blue
using ZnO nanorods

134

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


Fig. 5.4 shows the time dependent UV–Vis spectra of methylene blue dye during
photo irradiation with ZnO nanorods. The rate of decolorization was recorded with
respect to the change in the intensity of absorption peak in visible region. The prominent
peak was observed at λmax i.e., 665 nm which decreased gradually with increase in
irradiation time from 1 hour, 2 hour and 3 hour indicating that the dye had been degraded.
The decolorization of dye was achieved as 42%, 47% and 74% for the irradiation time of
1hour, 2hour and 3 hour respectively.

100

90

80

70
% D ecolorization

60

50 UV + ZnO

40

30

20

10

0
0 20 40 60 80 100 120 140 160 180
T im e (m in s )

Fig. 5.5 Photo catalytic decolorization of methelene blue dye with various irradiation
time
Fig 5.5 shows the effect of irradiation time on the decolorization of methelene blue
at natural pH. It can be seen that initial slopes of the curves representing rate of
decolorization, increase greatly by increasing irradiation time from 1hour, 2hour and 3
hour for MB.

5.3.2 Effect of Area of catalyst


Fig. 5.6 shows the area dependent UV–Vis spectra of MB dye during photo
irradiation with ZnO nanorods. The ZnO nanorods are used as catalyst. The area of the
ZnO nanorod thin films are varied and dipped in MB and exposed to UV-light. The rate
of decolorization was recorded with respect to the change in the intensity of absorption

135

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


peak. No shift in absorbance at 665 nm was recorded. In this case, the absorbance of MB
shows the decrease in intensity of MB.

2.0
B m ethelene blue
1.8 C ZnO -1cm x 1cm
D ZnO -2cm x 2cm
1.5

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
500 550 600 650 700 750
W avelength(nm )

Fig. 5.6 Area-dependent UV–Vis absorption spectra for decolorization of methelene blue
using ZnO nanorods
When the mixture of ZnO nanorods and MB was exposed to UV-light with
variation in the area of the catalyst, 75% decomposition of MB was recorded. The
absorbance observed at 665 nm has been decreased gradually with increase in area of the
catalyst from 1cm2 to 2cm2, indicating the degradation of the dye.

100

90

80

70
% Decolorisation

60

50
UV + ZnO
40

30

20

10

0
0 10 20 30 40
Area of catalyst (m m x m m )

Fig. 5.7 Photo catalytic decolorization of methelene blue dye with varying area of
catalyst

136

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


The decolorization of dye was achieved as 43% and 75% for the variation in area
of the catalyst from 1cm2 to 2cm2. Fig 5.7 shows the effect of area of catalyst on the
decolorization of methelene blue. The photo catalytic destruction of other organic
pollutants has also exhibited the same dependency on catalyst dose [43]. This can be
explained on the basis of catalyst loading. It is found to be dependent on initial solute
concentration because with the increase in catalyst dosage, total active surface area
increases. Hence availability of more active sites on catalyst surface increases [44].
The results indicate that ZnO exhibits higher photo catalytic activity when
irradiation time and area of catalyst is increased for the decolorization of MB. The
decolorization efficiencies are given in tables 5.3 and 5.4.
Table 5.3 Photo degradation efficiencies of ZnO nanorods with various irradiation time

Irradiation time Efficiency (%)


1 hour 42 %
2 hour 47 %
3 hour 74 %

Table 5.4 Photo degradation efficiencies of ZnO nanorods with varying area of catalyst

Area of catalyst Efficiency (%)


1 cm2 43 %
2 cm2 75 %

5.3.3 Effect of doping with aluminium, strontium and lithium


Fig. 5.8 and 5.9 show time and area dependent UV–Vis spectra of MB dye during
photo irradiation with aluminium doped ZnO nanorods. The rate of decolorization was
recorded with respect to the change in the intensity of absorption peak.
 For Al doped ZnO nanorods the degradation efficiency was found to increase from
28%, 32% to 34.2% for the irradiation time of 1hour, 2hour and 3 hour
respectively.

137

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


 It was also found that the degradation efficiency increases from 32% to 37% as the
area of catalyst was increased from 1cm2 to 2cm2.

2.0

B methelene blue
1.8 C Al doped ZnO-1hour
D Al doped ZnO-2hour
1.5 E Al doped ZnO-3hour

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
450 500 550 600 650 700 750
Wavelngth(nm)

Fig. 5.8 Time-dependent UV–Vis absorption spectra for decolorization of methylene blue
using aluminium doped ZnO nanorods

2.0

B methelene blue
1.8 C Al doped ZnO-1cm x 1cm
D Al doped ZnO-2cmx2cm
1.5

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
500 550 600 650 700 750
Wavelength(nm)

Fig. 5.9 Area-dependent UV–Vis absorption spectra for decolorization of methylene blue
using aluminium doped ZnO nanorods

Fig. 5.10 and 5.11 show time and area dependent UV–Vis spectra of MB dye during
photo irradiation with strontium ZnO nanorods. Exposure of the mixture of strontium
doped ZnO nanorods and MB to UV light resulted in 42% and 45% degradation of MB.

138

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


2.0

B mehtelene blue
1.8 C St doped ZnO - 1hour
D St doped ZnO - 2hour
1.5 E St doped ZnO - 3hour

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
450 500 550 600 650 700 750
Wavelength(nm)

Fig. 5.10 Time-dependent UV–Vis absorption spectra for decolorization of methylene


blue using strontium doped ZnO nanorods

2.0
B methelene blue
1.8 C St dop ZnO-1cmx1cm
D St dop ZnO-2cmx2cm
1.5

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
500 550 600 650 700 750
Wavelength(nm)

Fig. 5.11 Area-dependent UV–Vis absorption spectra for decolorization of methylene


blue using strontium doped ZnO nanorods

 For Sr doped ZnO nanorods the degradation efficiency was found to increase from
27%, 32% to 42% for the irradiation time of 1hour, 2hour and 3 hour respectively.
 It was also found that the degradation efficiency increases from 30.5% to 44.7% as
the area of catalyst increases from 1cm2 to 2cm2.

139

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


Fig. 5.12 and 5.13 show time and area dependent UV–Vis spectra of MB dye
during photo irradiation with lithium doped ZnO nanorods. Interesting results were
obtained when lithium doped ZnO nanorods and MB mixtures were exposed to UV light.
Rapid degradation of MB was recorded which showed maximum of 34.2% and 30.5%
degradation of MB.

2.0
B methelene blue
1.8 C Li doped ZnO-1hour
D Li doped ZnO-2hour
1.5 E Li doped ZnO-3hour

1.3
Absorbance(a.u)

1.0

0.8

0.5

0.3

0.0
500 550 600 650 700 750
Wavelength(nm)

Fig. 5.12 Time-dependent UV–Vis absorption spectra for decolorization of methylene


blue using lithium doped ZnO nanorods

2.0
B methelene blue
1.8 C Li dop ZnO-1cmx1cm
D Li dop ZnO-2cmx2cm
1.5

1.3
Absorption(a.u)

1.0

0.8

0.5

0.3

0.0
500 550 600 650 700 750
Wavelength(nm)

Fig. 5.13 Area-dependent UV–Vis absorption spectra for decolorization of methylene


blue using lithium doped ZnO nanorods

140

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


 For Li doped ZnO nanorods the degradation efficiency was found to increase from
25.2%, 28% to 34.2% for the irradiation time of 1hour, 2hour and 3 hour
respectively.
 It was also found that the degradation efficiency increases from 25% to 30.5% as
the area of catalyst increases from 1cm2 to 2cm2.
The results indicate that ZnO nanorods doped with aluminium, strontium and lithium
exhibit photo catalytic activities when irradiation time and area of catalyst is increased
for the decolorization of MB. In the above three cases, strontium doped ZnO nanorods
exhibit better photocatalytic efficiency than the aluminium and lithium doped ZnO
nanorods. The photo degradation efficiencies were tabulated for various irradiation time
and area of catalyst in tables 5.5 and 5.6
Table 5.5 Photo degradation efficiencies of doped ZnO nanorods for various irradiation
time
Irradiation time Efficiency (%)
Al doped ZnO Sr doped ZnO Li doped ZnO
1 hour 28% 27% 25.2%
2 hour 32% 32% 28%
3 hour 34.2% 42% 34%

Table 5.6 Photo degradation efficiencies of doped ZnO nanorods with varying area of
catalyst
Area of catalyst Efficiency (%)
Al doped ZnO Sr doped ZnO Li doped ZnO
2
1 cm 32% 30.5% 25%
2 cm2 37% 44.7% 30.5%
When comparing the photocatalytic activity of the undoped ZnO nanorods and the
ZnO nanorods doped with aluminium, strontium and lithium, the efficiency of the photo
degradation of methylene blue dye is high in the undoped ZnO nanorods. The above
degradation efficiencies were the result of 8W- UV irradiation. The textile reactive dye
MB is degraded at the maximum of 75% efficiency with 8W-UV source. It is an
encouraging result and it can be taken for future research.

141

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


References
[1] Ali, A. M.; Emanuelsson, E. A. C.; Patterson, D. A. Appl. Catal. B, 97, 168–181. doi:
10.1016 /j.apcatb.2010.03.037 (2010).
[2] Pardeshi, S. K.; Patil, A. B. J. Mol. Catal. A: Chem. 2009, 308, 32–40. doi:10.1016
/j.molcata. 2009.03.023.
[3] Qamar, M.; Muneer, M. Desalination 2009, 249, 535–540, doi:10.1016
/j.desal.01.022 (2009).
[4] Poulios, I.; Makri, D.; Prohaska, X. Global NEST, 1, 55–62 (1999).
[5] Carraway, E. R.; Hoffman, A. J.; Hoffmann, M. R. Environ. Sci. Technol., 28, 786–
793. doi:10.1021/es00054a007 (1994).
[6] M. Miyauchi, A. Nakajima, T. Watanabe, K. Hashimoto, Photocatalysis and photo
induced hydrophilicity of various metal oxide thin films, Chem. Mater. 14, 2812–2816
(2002).
[7] A. Akyol, M. Bayramoglu, Photocatalytic degradation of Remazol Red F3B using
ZnO catalyst, J. Hazard. Mater. B 124, 241–246 (2005).
[8] N. Daneshvar, D. Salari, A.R. Khataee, Photocatalytic degradation of azo dye acid red
14 in water on ZnO as an alternative catalyst to TiO2, J. Photochem. Photobiol. A: Chem.
162, 317–322 (2004).
[9] R. Comparelli, E. Fanizza, M.L. Curri, P.D. Cozzi, G. Mascolo, G. Agostiano, UV-
induced photocatalytic degradation of azo dyes by organic-capped ZnO nanocrystals
immobilized onto substrates, Appl. Catal. B: Environ. 60, 1–11 (2005).
[10] V. Kandavelu, H. Kastien, K.R. Thampi, Photocatalytic degradation of isothiazolin-
3-ones in water and emulsion paints containing nanocrystalline TiO2 and ZnO catalysts,
Appl. Catal. B: Environ. 48, 101–111 (2004).
[11] P. Percherancier, R. Chapelon, B. Pouyet, Semiconductor-sensitized pho-
todegradation of pesticides in water: the case of carbetamide, J. Photochem. Photobiol. A:
Chem. 87, 261–266 (1995).

142

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


[12] A.A. Khodja, T. Sehili, J.F. Pihichowski, P. Boule, Photocatalytic degradation of 2-
phenylphenol on TiO2 and ZnO in aqueous suspensions, J. Photochem. Photobiol. A:
Chem. 141, 231–239 (2001).
[13] C.C. Chen, Degradation pathways of ethyl violet by photocatalytic reaction with
ZnO dispersions, J. Mol. Catal. A: Chem. 264, 82–92 (2006).
[14] C.A.K. Gouvêa, F. Wypych, S.G. Moraes, N. Durán, N. Nagata, P. Peralta-Zamora,
Semiconductor assisted photocatalytic degradation of reactive dyes in aqueous solution,
Chemosphere 40, 433–440 (2000).
[15] G. Wang, D. Chen, H. Zhang, J.Z. Zhang, J.H. Li, Tunable photocurrent spectrum in
well-oriented zinc oxide nanorod arrays with enhanced photocatalytic activity, J. Phys.
Chem. C 112, 8850–8855 (2008).
[16] Hornyak, G. L.; Dutta, J.; Tibbals, H. F.; Rao, A. Introduction to nanoscience; CRC
Press: Boca Raton, (2008).
[17] Kimura, T.; Yamauchi, Y.; Miyamoto, N. Chem.–Eur. J., 16,12069–12073.
doi:10.1002/chem.201001251 (2010).
[18] Baruah S, Thanachayanont C and Dutta J.Sci. Technol. Adv. Mater. 9, 025009
(2008).
[19] Z.Y.He, Y.G.Li, Q.H.Zhang, H.Z.Wang, Capillary microchannel-based
microreactors with highly durable ZnO/TiO2 nanorod arrays for rapid, high efficiency
and continuous flow photocatalysis, Appl.Catal.B:Environ. 93, 376–382 (2009).
[20] A. Sharma, P. Rao, R.P. Mathur, S.C. Ameta, Photocatalytic reactions of xylidine
ponceau on semiconducting zinc oxide powder, J. Photochem. Photobiol. A 86, 197–200
(1995).
[21] F.D. Mai, C.C. Chen, J.L. Chen, S.C. Liu, J. Chromatogr. A 1189, 355–365 (2008).
[22] C. Lu, Y. Wu, F. Mai, W. Chung, C. Wu, W. Lin, C. Chen, Degradation efficiencies
and mechanisms of the ZnO-mediated photocatalytic degradation of Basic Blue 11 under
visible light irradiation, J. Mol. Catal. A: Chem. 310, 159–165 (2009).

143

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


[23] N. Sobana, M. Swaminathan, Combination effect of ZnO and activated carbon for
solar assisted photocatalytic degradation of Direct Blue 53, Sol. Energy Mater. Sol. Cells
91, 727–734 (2007).
[24] S.K. Kansal, A.H. Ali, S. Kapoor, Photocatalytic decolorization of biebrich scarlet
dye in aqueous phase using different nanophotocatalysts, Desalination 259, 147–155
(2010).
[25] S. Sakthivel, B. Neppolian, M.V. Shankar, B. Arabindoo, M. Palanichamy, V.
Murugesan, Solar photocatalytic degradation of azo dye: comparison of pho-tocatalytic
efficiency of ZnO and TiO2, Sol. Energy Mater. Sol. Cells 77, 65–82 (2003).
[26] G. Colón, M.C. Hidalgo, J.A. Navío, E. Pulido Melián, O. González Díaz, J.M.
Do˜na Rodríguez, Highly photoactive ZnO by amine capping-assisted hydrothermal
treatment, Appl. Catal. B 83, 30–38 (2008).
[27] N. Kislov, J. Lahiri, H. Verma, D.Y. Goswami, E. Stefanakos, M. Batzill,
Photocatalytic degradation of methyl orange over single crystalline ZnO: orientation
dependence of photoactivity and photostability of ZnO, Langmuir 25, 3310–3315 (2009).
[28] S. Daniele, M.N. Ghazzal, L.G. Hubert-Pfalzgraf, C. Duchamp, C. Guillard, G.
Ledoux, Mater. Res. Bull. 41, 2210–2218 (2006).
[29] M. El-Kemary, H. El-Shamy, I. El-Mehasseb, Photocatalytic degradation of
ciprofloxacin drug in water using ZnO nanoparticles, J. Lumin. 130, 2327–2331 (2010).
[30] R. Kitture, S.J. Koppikar, R. Kaul-Ghanekar, S.N. Kale, Catalyst efficiency, pho-
tostability and reusability study of ZnO nanoparticles in visible light for dye degradation,
J. Phys. Chem. Solids 72, 60–66 (2011).
[31] V. Shinde, T.P. Gujar, T. Noda, D. Fujita, A. Vinu, M. Grandcolas, J. Ye, Growth of
shape- and size-selective zinc oxide nanorods by a microwave-assisted bath deposition
method: effect on photocatalysis properties, Chem. Eur. J. 16, 10569–10575 (2010).
[32] F. Xu, P. Zhang, A. Navrotsky, Z.-Y. Yuan, T.-Z. Ren, M. Halasa, B.-L. Su, Hierar-
chically assembled porous ZnO nanoparticles: synthesis, surface energy, and
photocatalytic activity, Chem. Mater. 19, 5680–5686 (2007).

144

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.


[33] X. Li, K. Lv, K. Deng, J. Tang, R. Su, J. Sun, L. Chen, Synthesis and
characterization of ZnO and TiO2 hollow spheres with enhanced photoreactivity, Mater.
Sci. Eng. B 158, 40–47 (2009).
[34] M.S. Mohajerani, A. Lak, A. Simchi, Effect of morphology on the solar photocat-
alytic behavior of ZnO nanostructures, J. Alloys Compd. 485, 616–620 (2009).
[35] Y. Wang, X. Li, N. Wang, X. Quan, Y. Chen, Controllable synthesis of ZnO
nanoflowers and their morphology-dependent photocatalytic activities, Sep. Purif.
Technol. 62, 727–732 (2008).
[36] B. Li, Y. Wang, Facile synthesis and enhanced photocatalytic performance of
flower-like ZnO hierarchical microstructures, J. Phys. Chem. C 114, 890–896 (2010).
[37] Wang, R.; Xin, J. H.; Yang, Y.; Liu, H.; Xu, L.; Hu, J. Appl. Surf. Sci., 227, 312–
317. doi:10.1016/j.apsusc.2003.12.012 (2004).
[38] Vanheusden, K.; Warren, W. L.; Voigt, J. A.; Seager, C. H.; Tallant, D. R. Appl.
Phys. Lett., 67, 1280–1282. doi:10.1063/1.114397 (1995).
[39] Colis, S.; Bieber, H.; Begin-Colin, S.; Schmerber, G.; Leuvrey, C.; Dinia, A. Chem.
Phys. Lett., 422, 529–533. doi:10.1016/j.cplett.2006.02.109 (2006).
[40] Ullah, R. Dutta, J. J. Hazard. Mater., 156, 194–200, doi:10.1016 /j.jhazmat
2007.12.0 33 (2008).
[41] Baruah, S.; Rafique, R. F.; Dutta, J. NANO, 3, 399–407. doi:10.1142
/S17932920080 0126X (2008).
[42] http://www.methylene-blue.com/substance.php
[43] A. Akyol, H.C. Yatmaz, M. Bayramoglu, Appl. Catal. B Environ. 54, 19 (2004).
[44] M.S.T. Gonclaves, A.M.F. Oliveira-Campose, E.M.M.S. Pinto, P.M.S. Plasencia,
M.J.R.P Queiroz, Chemosphere 39:781, doi:10.1016/S0045-6535(99)00013-2, (1999).

145

Please purchase PDF Split-Merge on www.verypdf.com to remove this watermark.

You might also like