You are on page 1of 16

1

INTRODUCTION
1.1 Technological Overview
Nanoscience and technology are to be the dening technology of the 21
st
century.
Nanoscience is based on the fact that properties of materials change as a function of the
physical dimension of that material, and nanotechnology takes advantage of this by applying
selected property modifications of this nature to some beneficial endeavor.
Controlling the size and shape of the nanostructures are of both fundamental and
technological interest because it provides effective strategies for tuning the electronic,
magnetic, catalytic and optical properties of materials. One-dimensional (1D) nanostructures
have attracted considerable attention in recent years due to their potential as building blocks
for the assembly of electronic and optoelectronic device systems such as sensors, laser
diodes, eld-effect transistors, light-emitting diodes, and flat panel displays [1-2]. Among the
optoelectronic semiconductors, zinc oxide with a wide direct band gap (3.37eV) and a large
exciton binding energy (60 meV) is recognized as one of the most promising semiconductor
materials for electronic and photonic applications because of its unique optical,
semiconducting, magnetic and gas sensing properties [1-12]. Owing to its
noncentrosymmetric structure, ZnO is well-known as a piezoelectric material in surface
acoustic wave (SAW) devices for delay lines, filters, resonators in wireless communication,
and signal processing [3]. On the basis of the remarkable physical properties and the versatile
applications of the ZnO material, a number of 1D ZnO nanomaterials with different
morphologies such as nanorods, nanowires, nanobelts, nanotubes, nanorings, nanobridges,
and nanonails [4] have been successfully synthesized. Diverse morphologies of ZnO
structures have been synthesized by various processes such as RF-sputtering [5], pulsed laser
deposition (PLD) [6], molecular beam epitaxy (MBE) [7], thermal evaporation [8] and
chemical vapour deposition (CVD) [9] as well as by wet processes like electro-deposition
2

[10], spray pyrolysis [11], solgel [12] and aqueous chemical growth (ACG) [13]. ZnO
semiconductor nanostructures also play an important role in developing smart materials that
can simultaneously sense and destroy harmful chemical contaminants from our environment.
Waste water from textile, paper, and some other industries contain residual dyes,
which are not readily biodegradable and poses severe threats to aqueous environments when
released without treatment. The elimination of harmful components from waste water is an
important goal for environmental control in industrial settings. Adsorption and chemical
coagulation are the two common techniques of treatment of such waste water. However,
these methods merely transfer dyes from the liquid to the solid phase causing secondary
pollution and require further treatment [14]. Advanced Oxidation Processes (AOPs) are
alternative techniques of destruction of dyes and many other organics in waste water and
effluents. Semiconductors such as TiO
2
, ZnO, Fe
2
O
3
, CdS, and ZnS can act as sensitizers for
light-induced redox processes due to the electronic structure of the metal atoms in chemical
combination, which is characterized by a lled valence band, and an empty conduction band
[15]. Upon irradiation, valence band electrons are promoted to the conduction band leaving a
hole behind. These electron-hole pairs can either recombine or can interact separately with
other molecules. The holes may react either with electron donors in the solution, or with
hydroxide ions to produce powerful oxidizing species like hydroxyl or super oxide radicals
[16].
The catalytic properties of photocatalysts can be improved by designing
semiconductormetal composite nanoparticles or semiconductor crystals with engineered
defects. Contact of metal with the semiconductor indirectly inuences the energetics and
interfacial charge transfer processes in a favorable way. The deposition of a noble metal on
semiconductor nanoparticles is an essential factor for maximizing the efciency of
photocatalytic reactions [17]. It is commonly assumed that the noble metal (e.g., Pt) acts as a
3

sink for photoinduced charge carriers and promotes interfacial charge transfer processes.
Metal islands or nanoparticles deposited on a semiconductor surface undergo Fermi level
equilibration following the charging with photogenerated electrons. The effect of Fermi level
equilibration is predominantly seen when the metal deposits consists of small islands or small
particles. Unlike bulk metals, the nanoparticles do not often exhibit ohmic contact with the
semiconductor surface which retain the charge before transferring them to the redox species.
During extended UV photolysis, electron capture by metal islands of Ag, Au and Cu becomes
inhibited as their Fermi-level shifts close to the conduction band of the semiconductor. Pt on
the other hand acts as an electron sink and fails to achieve Fermi-level equilibration. Based
on the photoelectrochemical study a Fermi-level shift of about ~150 mV was observed for the
TiO
2
/Au composite lm [18, 19]. The metal nanoparticles deposited on TiO
2
lms thus
proved benecial for improving the photoelectrochemical and photocatalytic performance.
The energy-level diagram illustrating the role of Pt and Au in dictating Fermi level
equilibration of ZnOMetal composite system is shown in figure 1.1.






Figure 1.1: Energy-level diagram illustrating the role of (a) Pt and (b) Au in dictating Fermi
level equilibration of ZnOMetal composite system [20].

1.2 Review of Metal Oxide Semiconductor Photocatalysts
Semiconductor photocatalysts such as metal oxide are typically used as activators that
catalyze the radical chain reaction in photocatalytic oxidation. Semiconductor photocatalysts
are preferred in photocatalytic treatments of dye wastewater for the following reasons:
4

(i) they are inexpensive; (ii) they contain low to no toxicity; (iii) they exhibit tunable
properties that can be modied by size reduction, doping, or sensitizers; (iv) they contain an
affording facility for a multielectron transfer process; and (v) they are capable of extending
their use without substantial loss in photocatalytic activity [21]. Several metal oxides have
been reported to be good photocatalysts, namely titanium dioxide (TiO
2
), zinc oxide (ZnO),
tungstate (WO
3
), vanadate (VO
4
), molybdate (MoO
4
) etc., [22-24]. The photocatalytic
properties of these metal oxides have been studied extensively, and results show that they
have active potential as photocatalysts in the degradation of dye in wastewater.
1.2.1 TiO
2
Semiconductor Photocatalyst
TiO
2
is used as a photocatalyst in dye wastewater treatment mainly because of its
ability to generate a high oxidizing electronhole pair, its good chemical stability, non-
toxicity and long-term photostability [21, 25, 26]. Up to now, several modications in the
structure of TiO
2
have been made to achieve the following: (i) reduce the band gap energy to
a solar light compatible level; [27, 28] (ii) enhance the efciency of electronhole production;
[29] and (iii) enhance the absorbency of organic pollutants onto TiO
2
by modifying the
surface structure [30]. The structure of TiO
2
can be modied by doping with certain metal
ions, such as rare earth metals, [29, 31, 32] noble metals [33] and transition metals [34] as
well as non-metals such as C, N, S and P [26, 27, 30, 35, 36]. Noble metals (Au, Pt, Ag and
Pd) are usually used as dopants to modify the TiO
2
structure. However, recently transition
metals (Cr, Mn, Fe, Co, Ni, Cu and Zn) have been used to replace noble metals and thus, to
reduce the overall catalyst production cost. Ghasemi et al. have found that TiO
2
-doped Fe
gave the best results, with more than 90% degradation and 75% Total Oxygen Demand
(TOC) removal efficiency of Acid Blue 92 upon UV light irradiation. The incorporation of
transition metal ions into the TiO
2
lattice was found to alter or lower the band gap energy and
shift the catalyst absorbance edge closer to the visible light region [34].
5

El-Bahy et al. [29] and Zhu et al. [31] took this step further by synthesizing TiO
2
with
solgel method and, at the same time incorporating rare earth metal ions into the catalyst
structure. The synthesized TiO
2
with the addition of the rare earth ions (La
3+
, Nd
3+
, Sm
3+
,
Eu
3+
, Gd
3+
or Yb
3+
) produced rare earth doped TiO
2
with Gd-TiO
2
, which exhibited the
highest decolorization (91.5%) of Direct Blue Dye (DB53) under UV irradiation [29]. Zhu et
al. [31] have found the decolorizing efciency of 91.2% for Methylene Blue after 180 min
irradiation with sunlight using Ce
3+
doped TiO
2
-SiO
2
.
Generally, there is growing concern over the recovery of TiO
2
from the reactor if the
AOP process is scaled up for industrial application. According to Pozzo et al. [37] although
TiO
2
is an effective catalyst in dye wastewater treatment, the post-treatment to separate the
catalyst from treated water is difcult and not cost effective. In addition TiO
2
particles
dispersed in the aqueous solution tend to coagulate in a prolonged photocatalytic degradation
process, causing reduction of both the surface area and treatment efciency [38]. In most
bench-scale studies, catalysts are often separated by simple centrifugation or ltration.
However, this is not feasible for separating catalyst at an industrial scale, mainly due to the
energy and time consumed by the separation processes [39]. Consequently, research on the
immobilization or synthesis of TiO
2
particles with magnetic behaviour is gaining importance.
Recently, TiO
2
particles were synthesized by coating a TiO
2
precursor onto the
surface of a ferromagnetic particle, acting as a magnetic core. Several magnetic TiO
2
species
were reported, from simple maghemite (-Fe
3
O
4
) coated with TiO
2
to a mixture of silica
(SiO
2
) and titania (TiO
2
) coatings on magnetite iron oxide (Fe
3
O
4
) [39-41]. These magnetic
TiO
2
species can easily be separated from treated waste-water using a magnetic eld.
Furthermore, this new type of photocatalyst usually exhibits high photocatalytic efciency;
for example, Kurinobu et al. [40] found that TiO
2
and SiO
2
coated on a Fe
3
O
4
magnetic core
could completely degrade three different types of dye in 240 min under UV irradiation.
6

1.2.2 Other Metal Oxide Semiconductor Photocatalysts
Apart from TiO
2
and ZnO, other metal oxide photocatalysts, such as vanadium,
tungsten, molybdenum, indium and cerium oxides, have been used to treat dye waste-water.
Some studies showed that these photocatalysts could be as effective as or even better than
conventional photocatalysts [59-64]. Metal vanadates, MVO
4
, have recently been considered
as potential semiconductors in dye waste-water treatment. Recently, the main focus has been
on the development of bismuth vanadate, BiVO
4
, which has been reported to be active in the
degradation of dye waste-water under visible light irradiation.
Guo et al. [60] have found that ms-BiVO
4
exhibited the highest UV light catalytic
activity regardless of its morphology; with Rhodamine B and Methylene Blue completely
degraded in 60 and 45 minutes respectively. Also, ms-BiVO
4
showed potential as a
photocatalyst in visible light because at pH 9 it could degrade up to 77.8% of Rhodamine B
and 65.0% of Methylene Blue after 6 h irradiation.
Tungsten oxide, WO
3
, is also reported to be photocatalytically active and effective for
the treatment of dye waste-water. According to Cruz et al. [64] the different morphologies
and physical properties of WO
3
, such as the band gap energy and surface area are the one
which would determine the treatment efciency. For example, samples with square-plate
morphologies and higher surface areas showed the highest TOC reduction of Indigo Carmine
up to 93%.
Coupling a WO
3
semiconductor to TiO
2
is also one of the ways to improve
photocatalytic degradation of dye waste-water [65, 66]. Sajjad et al. [66] have investigated
the operational parameters which could affect the degradation of Acid Orange 7 and Methyl
Orange by WO
x
-TiO
2
nanocomposites produced by a solgel method. They have found that a
catalyst loading of 1.0 gL
1
and 4.0% WO
x
-TiO
2
composites are able to degrade completely
7

Acid Orange 7 and Methyl Orange in 240 min and 300 min, respectively, under visible light
irradiation. This composite catalyst demonstrated a far superior performance than pure TiO
2.
Xu and Ni [67] have successfully fabricated azeolite-supported Bi
2
MoO
6
that was
photocatalytically active under visible light conditions. The synthesized catalyst was used to
treat Methyl Orange with more than 90% degradation obtained under visible light irradiation
with a 400W Xe lamp. The high photocatalytic activity of the synthesized catalyst could be
due to the increased formation of Bi
2
MoO
6
nanoparticles on the zeolite surface with a low
Si/Al ratio. Additional studies are being conducted by many researchers on other metal oxide
semiconductors to determine whether their role as photocatalysts can be expanded in the
future.
1.2.3 ZnO

Semiconductor Photocatalyst
ZnO exists in three different types of crystal structure; hexagonal wurtzite, cubic zinc
blende, and cubic rock salt. Wurtzite is the most stable structure and is abundantly available,
unlike the other two structures. ZnO has a wide band gap, and due to its low production cost
and its unique electrical, optoelectronic and luminescence properties, it has recently been
studied as a potential photocatalyst to degrade dyes.
The quantum efciency of ZnO is signicantly larger than that of TiO
2
. Thus, ZnO
absorbs over a larger fraction of the solar spectrum, rendering it a more suitable photocatalyst
for use under visible light conditions [22, 43-47]. Some of the most recent experimental
results also showed that ZnO actually exhibits a higher photocatalytic activity than TiO
2
,
especially when the degradation of industrial effluents occurs at neutral pH [45].
Most of the photocatalytic studies were conducted using commercial ZnO under UV
irradiation, [22, 31, 46-48], studies under visible light conditions have also been investigated
[48-50]. Other studies compared the ZnO photocatalytic activity with that of TiO
2
and other
semiconductor photocatalysts, such as SnO
2
, ZnS and CdS. There are many reports indicating
8

that, ZnO is generally a better photocatalyst under visible light conditions than the other
semiconductor metal oxides [49, 51-53]. Some researchers have found that ZnO showed
promising results as a photocatalyst for the degradation of azo dyes (Orange II and Direct
Yellow 12) [47, 54] and some reactive dyes (Remazol Black B and Remazol Brilliant Blue R)
[46] under UV conditions. For example, Nishio et al. [47] and Gouvea et al. [46] have found
that using ZnO, the model dye could be completely bleached by 60 min of UV light
irradiation. Also, high TOC removal of Remazol Brilliant Blue R (up to 90%) could be
achieved using ZnO with 120 min of UV light irradiation.
The photocatalytic efciency of ZnO can be enhanced by incorporating a noble metal.
Chen et al. [55] have found that the ZnO (synthesized from zinc acetate using a solvothermal
method) doped by Ag could mineralize Methyl Orange completely in 60 min with UV light
irradiation at 254 nm. Subsequently, a report by Pawinrat et al. [68] showed that by doping
ZnO with a small amount of Au (3 wt%) using one-step ame spray pyrolysis, the efciency
of the synthesized ZnO was enhanced degrading up to 71% of Methylene Blue in 1 h with
UV irradiation at 365 nm, close to the visible light region. The addition of the metal cluster
could enhance the photocatalytic activity of the semiconductor oxide by improving its charge
transfer by the metal cluster functioning as a trap for photoinduced charge carriers.
Several studies have attempted to modify the size and morphology of ZnO for
example, Wang and Xie [57] synthesized ZnO particles by using a mixture of zinc acetate
and lithium hydroxide, followed by agitation with magnetic beads. The photocatalytic
activity tested on Methyl Orange solution, found that the colloidal form was more reactive in
comparison with ZnO that had undergone an annealing process at various temperatures. A
smaller particle size of ZnO was found to enhance photoreactivity, with a degradation
efciency of 80% using 50 nm ZnO compared with results for 200 nm and 1000 nm samples
[58]. However, the smallest particle size, 10 nm, gave a lower efficiency. Thus, although the
9

10 nm catalyst showed good degradation efficiency during the initial treatment, the efficiency
decreased to a rate even lower than the 200 nm ZnO after about 70 min operation. This
decrease was due to inherent instability in the photodegradation process. In conclusion, the
photocatalytic activity of ZnO is not solely determined by the particle size and the
morphology but also the method of preparation plays a key role in enhancing its degradation
efficiency.
1.3 Motivation for the Work
In this millennium, we are facing the challenge of cleaning our natural water and air
resources. While we enjoy the comforts and benets that technology has provided us, from
composites to computers, from drugs to dyes, we also have the task of treating wastes
generated during manufacturing processes and the proper disposal of various products and
byproducts. Therefore, the elimination of toxic chemicals from wastewater is currently one of
the most important subjects. Universal attention has been given to the application of
photocatalysis in environmental water treatment and energy development fields. Many
industries produce wastewater which contains different types of organic and inorganic
pollutants. For example, textile industry uses large volumes of water in wet processing
operations and, thereby, generates substantial quantities of wastewater containing large
amounts of dissolved dyestuffs and other products, such as dispersing agents, dye bath
carriers, salts, emulsifiers, leveling agents and heavy metals [68-73]. Dyes are extensively
used for dyeing, printing and several other coloring purposes in industries. Synthetic dyes
are used in the textile, paper, tanning, pharmaceutical, cosmetics and food industries.
Over 50,000 tons of approximately 10,000 different dyes and pigments produced
annually worldwide are discharged into the environment [74]. It is estimated that
approximately 15% of the dyestuffs are lost in the industrial effluents during manufacturing
and processing operations [75, 76]. The color of the effluent discharged into ecosystem
10

affects the aquatic flora and fauna and causes many environmental diseases. Some dyes are
toxic and/or carcinogenic, and the biodegradation of many dyes yields aromatic amines,
which may be carcinogenic or otherwise toxic [77, 78]. Moreover, dyes can accumulate in
sediment and soil, especially at locations of wastewater discharge, which causes different
problems to the ecological balance in the environment. Ground water systems are also
affected by these pollutants due to leaching from the soil [79].
Nanotechnology can provide us ways to purify the air and water resources by utilizing
semiconductor nanoparticles as catalysts and/or sensing systems. Furthermore the utilization
of light and ultrasound to activate such nanoparticles opens up new ways to design green
oxidation technologies for environmental remediation. The size and shape dependent optical
and electronic properties of nanoparticles make an interesting case for environmental
chemists to exploit their role in environmental remediation processes [80-84]. Of particular
interest is their application in advanced oxidation processes [85]. Various metal oxide
semiconductor nanoparticles have been extensively studied for oxidative and reductive
transformation of organic and inorganic species present as contaminants in air and water.
Technological advances in this area have already led to product development for a variety of
day-to-day operations. Commercialization of products such as self-cleaning glasses,
disinfectant tiles and lters for air purication demonstrates the early successes of
nanosystems for environmental applications [86]. Semiconductor nanostructures can also
play an important role in developing smart materials that can simultaneously sense and
destroy harmful chemical contaminants present in the environment. Such an application
seems to be important as the concern over chemical contamination of water and air needs to
be addressed.
Maximizing the efficiency of photo-induced charge separation in semiconductor
systems remains to be a major challenge to the scientic community. Obtaining insight into
11

charge transfer processes is important to improve the photoconversion efficiencies in
semiconductor based nanostructures. 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 [87-89]. ZnO exhibits
comparatively higher reaction and mineralization rates [90] and can generate hydroxyl ions
more efficiently than titania (TiO
2
) [91]. 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. Photocatalytic activity comparable to doped ZnO was also observed in ZnO crystals
with engineered defects achieved by a slow growth process during synthesis of the
nanoparticles, thereby tuning the structural, optical and electronic properties. Hydrothermally
grown ZnO nanostructures have inherent crystalline defects primarily due to oxygen
vacancies, interstitial defects and oxygen nonstoichiometry. Therefore, these interesting and
inevitable reasons motivated to make an attempt towards the study on morphology controlled
synthesis of zinc oxide semiconductor nanostructures using hydrothermal technique and their
influences on photocatalytic degradation of model textile dyes.
1.4 Details of the Present Work
The main objective of the present work is to synthesize and study the potential of
various hydrothermally grown zinc oxide semiconductor nanostructures on photocatalytic
degradation of model textile dyes for photocatalytic treatment of water pollutants.
ZnO nanostructures have been synthesized using zinc nitrate hexahydrate
[Zn(NO
3
)
2
6H
2
O] and hexamethylenetetramine (HMT) [(CH
2
)
6
N
4
]

in a controlled manner by
varying the pH of the precursor solution using hydrothermal technique. The morphological
changes of the prepared ZnO nanostructures have been investigated in the range of pH 510.
Radial hexagonal rod-like shape is formed at lower pH values of 5 and 6 whereas, flower-like
12

shape is obtained for higher pH values of 9 and 10. Flake-like structure is observed at
moderate pH of 8. The prepared ZnO nanostructures have been characterized using x-ray
diffraction technique (XRD), energy dispersive x-ray analysis (EDAX), scanning electron
microscope (SEM) and FTIR spectroscopy. UV-Vis spectroscopy and photoluminescence
have been applied to study the optical properties.
The potential of the prepared semiconductor ZnO nanostructures as an effective
catalyst for the photo degradation of model textile dyes: Methylene Blue, Methyl Orange and
Rodamine B, in water under UV irradiation has been studies. The efficiency of synthesized
flake-like, flower-like and rod-like photocatalyst on the extent of photocatalytic degradation
of model dyes after 5h have been determined by the reduction in absorbance of the solution.
The chemical oxygen demand (COD) test is applied to measure the organic strength of initial
and treated model dye solutions.













13

References
1. Tian Yu, Lu Hong-Bing, Liao Lei, Li Jin-Chai, Wu Yun, Fu Qiang, Physica E 41 729
(2009)
2. M.H. Kim, Y.H. Cho, H. Lee, S.I. Kim, S.R. Ryu, D.Y. Kim, T.W. Kang, K.S.
Chung, Nano Lett. 4 1059 (2004)
3. W. Zhang-lin, J. Mater. Chem. 15 1021 (2005)
4. Wang Zhong-lin, J. Materials Today 7 26 (2004)
5. W.C. Shih, M.S. Wu, J. Crystal Growth 137 319 (1994)
6. Y. Sun, G.M. Fuge, M.N.R. Ashfold, Chem. Phys. Lett. 396 21 (2004)
7. Y.W. Heo, V. Varadarajan, M. Kaufman, K. Kim, D.P. Norton, Appl. Phys. Lett. 81
3046 (2002)
8. J.Q. Hu, X.L. Ma, Z. Y. Xie, N.B. Wong, C.S. Lee, S.T. Lee, Chem. Phys. Lett. 344
97 (2001)
9. J.J. Wu, S.C. Liu, Adv. Mater. 14 215 (2002)
10. B. Illy, B.A. Shollock, J.L. MacManus-Driscoll, M.P. Ryan, Nanotechnology 16 320
(2005)
11. S.A. Studenkin, N. Golego, M. Cocivera, J. Appl. Phys. 83 2104 (1998)
12. M. Ohyama, H. Kozuka, T. Yoko, Thin Solid Films 306 78 (1997)
13. D. Vernardou, G. Kenanakis, S. Couris, A.C. Manikas, G.A. Voyiatzis, M.E. Pemble,
E. Koudoumas, N. Katsarakis, Journal of Crystal Growth 308 105 (2007)
14. K. Tanaka, K. Padermpole, T. Hisanaga, Water Res. 34 327 (2000)
15. M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahenemann, Chem. Rev. 95 69 (1995)
16. W.Z. Tang, An. Huren, Chemosphere 31 4157 (1995)
17. A.J. Bard, J. Phys. Chem. 86 172 (1982)
18. N. Chandrasekharan, P.V. Kamat, J. Phys. Chem. B 104 10851(2000)
19. V. Subramanian, E.E. Wolf, P.V. Kamat, J. Phys. Chem. B 107 7479 (2003)
20. M. Jakob, H. Levanon, P.V. Kamat, Nano Lett. 3 353 (2003)
21. Chatterjee D and Dasgupta S, J Photochem Photobiol C: Phot chem Rev 6 186 (2005)
22. Chakrabarti S and Dutta BK, J Hazard Mater 112 269 (2004)
23. Janitabar-Darzi S and Mahjoub AR, J Alloys Compd 486 805 (2009)
24. Guo Y, Yang X, Ma F, Li K, Xu L, Yuan X, Appl Surf Sci 256 2215 (2010)
25. Han F, Kambala VSR, Srinivasan M, Rajarathnam D and Naidu R, Appl Catal A
Gen 359 25 (2009)
14

26. Zhang S and Song L, Catal Commun 10 1725 (2009)
27. Lv Y, Yu L, Huang H , Liu H and Feng Y, J Alloy Compd 488 314 (2009)
28. Hamadanian M, Reisi-Vanani A and Majedi A, Appl Surf Sci 256 1837 (2010)
29. El-Bahy ZM, Ismail AA and Mohamed RM, J Hazard Mater 166 138 (2009)
30. Janus M Choina J and Morawski AW, J Hazard Mater 166 1 (2009)
31. Zhu J, Xie J, Chen M, Jiang D and Wu D, Colloids Surf A-Physicochem Eng Asp 355
178 (2010)
32. Du P, Bueno-Lopez A, Verbaas M, Almeida AR, Makkee M , Mouljin JA, J Catal
260 75 (2008)
33. Yang X, Ma F, Li K, Guo Y , Hu J, Li W, J Hazard Mater175 429 (2010)
34. Ghasemi S, Rahimnejad S, Setayesh SR, Rohani S and Golami MR, J Hazard Mater
172 1573 (2009)
35. Zheng R, Guo Y, Jin C, Xie J, Zhu Y and Xie Y, J Mol Catal AChem 319 46 (2010)
36. Znad H and Kawase Y, J Mol Catal AChem 314 55 (2009)
37. Pozzo RL, Baltan MA and Cassano AE, Catal Today 39 219 (1997)
38. Puma GL, Bono A, Krishnaiah D and Collin JG, J Hazard Mater 157 209 (2008)
39. GaoY,Chen B,Li H and Ma Y, Mater Chem Phys 80 348 (2003)
40. Kurinobu S, Tsurusaki K, Natui Y, Kinamata M and Hasegawa M, J Magn Magn
Mater 310 1025 (2007)
41. Gad-Allah T A, Fujimura K, Kato S, Satokawa S and Kojima T, J Hazard Mater 154
572 (2008)
42. Muruganandham M , Chen I S and Wu JJ, J Hazard Mater 172 61 (2009)
43. Qiu R, Zhang D, Mo Y, Song L, Brewer E, Huang X , J Hazard Mater 156 80 (2007)
44. Behnajady MA, Moghaddam SG, Modirshahla N and Shokri M, Desalination 249
1371 (2009)
45. Chen X, He Y, Zhang Q , Li L , Hu D and Yin T, J Mater Sci 45 953 (2010)
46. Gouvea CAK, Wypych F, Moraes SG , Duran N, Nagata N and Peralta-Zamora P,
Chemosphere 40 433 (2000)
47. Nishio J, Tokumura M, Znad HT and Kawase Y , J Hazard Mater 138 106 (2006)
48. Pare B, Singh P and Jonnalgadda SB, Indian J C hem Sect A Inorg Bio-Inorg Phys
Theor Anal Chem 48 1364 (2009)
49. Lu C, Wu Y, Mai F, Chung W, Wu C, xsLin W, J Mol Catal A C hem 310 159
(2009)
50. Mai F D, Chen CC, Chen JL and Liu SC, J Chromat A 1189 355 (2008)
15

51. Akyol A and Bayramoglu M, Chem Eng Process 47 2150 (2008)
52. Kansal SK, Kaur N and Singh S, Nanoscale Res Lett 4 709 (2009)
53. Kansal SK, Singh M and Sud D, J Hazard Mater 141 581 (2007)
54. Rao AN, Sivasankar B and Sadasivam V, J Mol Catal AChem 306 77 (2009)
55. Chen T, Zheng Y , Lin JM and Chen G , J Am Soc Mass Spectrum 19 997 (2008)
56. Pawinrat P, Mekasuwandumrong O and Panpranot J, Catal Commun 10 1380 (2009)
57. Wang H and Xie C, J Phys C hem Solids 69 2440 (2008)
58. Wang H, Xie C, Zhang W, Cai S, Yang Z and Gui Y, J Hazard Mater 141 645 (2007)
59. Seddigi ZS, Bull Environ Cont Toxicol 84 564 (2010)
60. Guo Y, Yang X, Ma F, Li K, Xu L, Yuan X Appl Surf Sci 256 2215 (2010)
61. Yin W, Wang W, Zhou L, Sun S and Zhang L J Hazard Mater 173 194 (2010)
62. Castillo NC, H eel A, Graule T and Pulgarin C, Appl Catal B Environ 95 335 (2010)
63. Xu H, Li H, Sun G , Xia J, Wu C, Ye Z, Chem Eng J 160 33 (2010)
64. Cruz AM-dl, Martnez DS and Cuellar EL, Solid State Sci 12 88 (2010)
65. Sajjad AKL, Shamaila S, Tian B, Chen F and Zhang J, Appl Catal B Environ 91
397 (2009)
66. Sajjad AKL, Shamaila S, Tian B, Chen F and Zhang J, J Hazard Mater 177 781
(2010)
67. Xu X and Ni Q, Catal Commun 11 359 (2010)
68. S.A. Figueiredo, J.M. Louriero and A.R. Boaven-tura, Water Res., 39 4142 (2005)
69. C. Sourja, D. Sirshendu, D. Sunando and K.B. Jayanta, Chemosphere, 58 1079 (2005)
70. C. Namasivayam and S. Sumithra, J. Environ. Managem., 74 207 (2005)
71. O. Abdelwahab, A. El Nemr, A. El-Sikaily and A. Khaled, Chem. Ecol., 22 253
(2006)
72. A. El-Sikaily, A. Khaled, A. El Nemr and O. Abdel-wahab, Chem. Ecol., 22 149
(2006)
73. B. Noroozi, G.A. Sorial, H. Bahrami and M. Arami, J. Hazard. Mater B 139 167
(2007)
74. Kaushik P, Malik A. Environ Int., 35 127 (2009)
75. A. Reife and H.S. Freemann, Wiley, Canada (1996)
76. C. Sudipta, C. Sandipan, P.C. Bishnu, R.D. Akhil and K.G. Arun, J. Coll. Interf. Sci.,
288 30 (2005)
77. M.F. Boeniger, Department of Health and Human Services (NIOSH), Pub. No. 8-119,
Cincinnati, OH, (1980)
16

78. T. Zheng, T.R. Holford, S.T. Mayne, P.H. Owens, P. Boyle and B. Zhang, J. Cancer,
38 1647 (2002)
79. C. Namasivayam and S. Sumithra, J. Environ. Managem., 74 207 (2005)
80. M.L. Steigerwald, L.E. Brus, Acc. Chem. Res. 23 183 (1990)
81. A. Henglein, J. Phys. Chem. 97 5457(1993)
82. P. Alivisatos, J. Phys. Chem. 100 13226 (1996)
83. P.V. Kamat, D. Meisel (Eds.), Elsevier Science, Amsterdam, (1997)
84. P.V. Kamat, J. Phys. Chem. B 106 7729 (2002)
85. P.V. Kamat, D. Meisel, Curr. Opin. Colloid Interface Sci. 7 282 (2002)
86. A. Fujishima, K. Hashimoto, T. Watanabe, Bkc, Inc. Tokyo, Japan, (1999)
87. Ali A. M., Emanuelsson E. A. C., Patterson D. A. Appl. Catal. B 97 168 (2010)
88. Pardeshi, S. K., Patil, A. B., J. Mol. Catal. A: Chem. 308 32 (2009)
89. Qamar M, Muneer M, Desalination 249 535 (2009)
90. Poulios I, Makri D, Prohaska X, Global NEST 1 55 (1999)
91. Carraway E. R, Hoffman A. J, Hoffmann M. R, Environ. Sci. Technol. 28 776 (1994)

You might also like