You are on page 1of 27

Water Air Soil Pollut

DOI 10.1007/s11270-010-0456-3

Advances in Heterogeneous Photocatalytic Degradation


of Phenols and Dyes in Wastewater: A Review
Saber Ahmed & M. G. Rasul & Wayde N. Martens &
Richard Brown & M. A. Hashib

Received: 7 January 2010 / Accepted: 21 April 2010


# Springer Science+Business Media B.V. 2010

Abstract The heterogeneous photocatalytic water


purification process has gained wide attention due
to its effectiveness in degrading and mineralizing
the recalcitrant organic compounds as well as the
possibility of utilizing the solar UV and visible
light spectrum. This paper aims to review and
summarize the recently published works in the field
of photocatalytic oxidation of toxic organic compounds such as phenols and dyes, predominant in
wastewater effluent. In this review, the effects of
various operating parameters on the photocatalytic
degradation of phenols and dyes are presented.
Recent findings suggested that different parameters,

such as type of photocatalyst and composition, light


intensity, initial substrate concentration, amount of
catalyst, pH of the reaction medium, ionic components in water, solvent types, oxidizing agents/
electron acceptors, mode of catalyst application,
and calcinations temperature can play an important
role on the photocatalytic degradation of organic
compounds in water environment. Extensive research has focused on the enhancement of photocatalysis by modification of TiO2 employing metal,
non-metal, and ion doping. Recent advances in TiO2
photocatalysis for the degradation of various phenols
and dyes are also highlighted in this review.

S. Ahmed (*) : M. G. Rasul


Faculty of Science, Engineering and Health, CQ University,
Rockhampton QLD 4702, Australia
e-mail: s.ahmed@cqu.edu.au

Keywords Phenols . Azo dyes . Photocatalysis .


Water purification

1 Introduction
W. N. Martens
Discipline of Chemistry, Faculty of Science and Technology,
Queensland University of Technology,
Brisbane, Australia
R. Brown
School of Engineering System,
Queensland University of Technology,
Brisbane, Australia
M. A. Hashib
Department of Ecological Engineering,
Toyohashi University of Technology,
Toyohashi, Japan

The reuse and recycling of wastewater effluent is


rapidly growing and becoming a necessity for water
utilities both in Australia and in other parts of the
world to augment our limited fresh water supply,
which is currently under pressure due to rapid
population growth (Radcliff 2006; Mitchell et al.
2002; DEH 2002). Heightened concerns over public
health and associated environmental hazards due to
the presence of toxic organic compounds such as
phenols and dyes in wastewater have been reported

Water Air Soil Pollut

(Eriksson et al. 2007). Phenols and azo dyes are wellknown for their bio-recalcitrant and acute toxicity.
Phenols and azo dyes are being continuously introduced into the aquatic environment through various
anthropogenic inputs. The presence of toxic organic
compounds in wastewater effluent is still a major
hindrance as regards widespread acceptance of water
recycling (Mahmoodi et al. 2007; DEC 2006).
Furthermore, their variety, toxicity, and persistence
can directly impact the health of ecosystem and
present a threat to humans through contamination of
drinking water supplies, e.g., surface and ground
water. The response has been the drive to achieve
effective removal of persistent organic pollutants from
wastewater effluent to minimize the risk of pollution
problems from such toxic chemicals to enable its
reuse. Consequently, considerable efforts have been
devoted to developing a suitable purification method
that can easily destroy these bio-recalcitrant organic
contaminants. Because of incomplete removal during
primary and secondary treatment processes, they are
ubiquitous in secondary wastewater effluents at low
concentration. Despite their low concentration, these
contaminants have raised substantial concern in the
public and regulatory agencies due to their extremely
high endocrine disrupting potency and geno toxicity
(Arques et al. 2007). Moreover, the conventional endof-pipe techniques also generate wastes during the
treatment of contaminated water, which requires
additional steps and cost. Heterogeneous photocatalytic oxidation (HPO) process employing catalyst
such as TiO2, ZnO, etc., and UV light has demonstrated promising results for the degradation of
persistent organic pollutants producing more
biologically degradable and less toxic substances
(Garcia et al. 2006; Garca et al. 2007; Garcia et al.
2008; Hincapie et al. 2006; Maldonado et al. 2007;
Malato et al. 2002; Oller et al. 2006).This process is
largely dependent on the in situ generation of
hydroxyl radicals under ambient conditions which
are capable of converting a wide spectrum of toxic
organic compounds including the non-biodegradable
one into relatively innocuous end products such as
CO2 and H2O. In HPO process, destruction of
recalcitrant organics is governed by combined action
of semiconductor photocatalyst, an energetic radiation
source and an oxidizing agent. Moreover, the process
can be driven by solar UV/visible light. Near the earth
surface, the sun produces 0.20.3 mol photons m-2h-1

in the range of 300400 nm with a typical UV flux of


2030 Wm-2.This suggests the feasibility of using
sunlight as an economically and ecologically sensible
light source (Goslich et al. 1997). As a result,
development of efficient photocatalytic water
purification process for large-scale applications is a
recent research challenge in order to bring forward
a new and efficient end-of-pipe technology.
Compounds studied using photocatalytic oxidations
process in aqueous medium belong to various types of
pesticides, phenolic compounds, and cationic and
anionic dyes (Shifu and Yunzhang 2007; Echavia et
al. 2009; Bahnemann et al. 2007). In light of the basic
and applied researches reviewed, photocatalytic
oxidation method appears to be a promising route
for the treatment of wastewater contaminated with
phenols and dyes. However, low quantum efficiency
due to inefficient visible light harvesting catalyst (Sun
and Bolton 1996), the design of photoreactor
(Mukherjee and Ray 1999), the recovery and reuse
of titanium dioxide (Moctezumaa et al. 2007), the
generation of toxic intermediates (Konstantinou and
Albanis 2003), as well as concern over catalyst
deactivation (Legrini et al. 1993; Rincon and Pulgarin
2004; Doll and Frimmel 2005) are reported to be the
major drawbacks. Information from various investigations suggest that photocatalytic degradation of
pesticides, phenols, and dyes is largely dependent on
the solution pH, types of catalyst and composition,
organic substrate type and concentration, light
intensity, catalyst loading, ionic composition of
wastewater, types of solvent, oxidant concentration,
and calcinations temperature (Shakthivel et al. 2004;
Byrappa et al. 2006). Understanding the impacts of
various process parameters that govern the photocatalytic degradation efficiency is of paramount
importance from the design and the operational points
of view when choosing a sustainable, efficient
technique for the treatment of wastewater. This paper
aims to review and summarize the role of important
operating parameters on the photocatalytic degradation of phenolic compounds and dyes together
with recent achievements. Recent developments in
TiO2 photocatalysis for the degradation of various
phenols and dyes by means of metal, non-metal, and
ion doping are also highlighted in this review.
The existing limitation and future research needs
associated with the treatment technology are also
discussed.

Water Air Soil Pollut

2 Principle of Photocatalytic Oxidation Process


In the photocatalytic oxidation process, organic
pollutants are destroyed in the presence of semiconductor photocatalysts (e.g., TiO2 and ZnO), an
energetic light source, and an oxidizing agent such as
oxygen or air. Only photons with energies greater than
the band gap energy (E) can result in the excitation
of valence band (VB) electrons which then promote the
possible reactions. The absorption of photons with
energy lower than E or longer wavelengths usually
causes energy dissipation in the forms of heat. The
illumination of the photocatalytic surface with
sufficient energy leads to the formation of a positive
hole (hv+) in the valence band and an electron (e-) in
the conduction band (CB). The positive hole oxidizes
either pollutant directly or water to produce OH
radicals, whereas the electron in the conduction band
reduces the oxygen adsorbed on the photocatalyst
(TiO2). The activation of TiO2 by UV light can be
represented by the following steps.

TiO2 + h

e- + h+

2:1

e-+ O2

O2 -

2:2

In this reaction, h+ and e- are powerful oxidizing


and reductive agents, respectively. The oxidative and
reductive reaction steps are expressed as,
Oxidative reaction:

h+ + Organic

h + H2O

3 Influence of Parameters on the Photocatalytic


Degradation of Phenols and Dyes
2:3

CO 2

OH +H

2:4

Reductive reaction:

OH +Organic

adsorbed water where it is adsorbed as OH-, is the


primary oxidant, and the presence of oxygen can
prevent the recombination of an electronhole pair. In
the photocatalytic degradation of pollutants, when the
reduction process of oxygen and the oxidation of
pollutants do not advance simultaneously, there is an
electron accumulation in the CB, thereby causing an
increase in the rate of recombination of e- and h+.
Thus, it is of paramount importance to prevent electron
accumulation in efficient photocatalytic oxidation. In
photocatalysis, TiO2 has been studied extensively
because of its high activity, desirable physical and
chemical properties, low cost, and availability. Of
three common TiO2 crystalline forms, anatase and
rutile forms have been investigated extensively as
photocatalysts. Anatase has been reported to be more
active as a photocatalyst than rutile. Similar oxidation
pathways to those of TiO2 are confirmed in ZnO
photocatalyst including the formation of OH radical
and the direct oxidation by photogenerated holes, etc.
ZnO is reported to be as reactive as TiO2 under
concentrated sunlight, since the band gap energy of
ZnO is equal to that of TiO2, i.e., 3.2e V. Different
light sources such as UV lamps and solar radiation
have been used in previous investigations into the
photocatalysis of various organic contaminants dominant in wastewater effluent. The following sections
describe the influence of parameters on the photocatalytic degradation of various phenols and dyes.

CO2

2:5

Hydroxyl radical generation by the photocatalytic


oxidation process is shown in the above steps. In the
degradation of organic pollutants, the hydroxyl
radical, which is generated from the oxidation of

3.1 Crystal Composition and Catalyst Type


The photocatalytic activity of TiO2 is dependent on
surface and structural properties of semiconductor such
as crystal composition, surface area, particle size
distribution, porosity, band gap, and surface hydroxyl
density. Average crystal size is of primary importance
in heterogeneous catalysis because it is directly related
to the efficiency of a catalyst through the definition of
its specific surface area. A number of commercially
available catalysts have been investigated for the
photocatalytic degradation of phenolic compounds
and dyes in aqueous environment. The photocatalyst
titanium dioxide Degussa P25 has been widely used in
most of the experimental conditions; other catalyst

Water Air Soil Pollut

powders, namely, Hombikat UV100, PC 500, and


travancore titanium products (TTP) were also used for
degradation of toxic organic compounds. P25 contains
75% anatase and 25% rutile with a specific BrunauerEmmett-Teller (BET) surface area of 50 m2/g and a
primary particle size of 20 nm (Bahnemann et al.
2007). Hombikat UV 100 consists of 100% pure and
smaller anatase with a specific BET surface area of
250 m2/g and a primary particle size of 5 nm
(Bahnemann et al. 2007).The photocatalyst PC 500
has a BET surface area of 287 m2/g with 100% anatase
and primary particle size of 510 nm (Bahnemann et
al. 2007), and TiO2 obtained from TTP, India, has a
BET surface area of 9.82 m2/g. It has been demonstrated that the degradation rate of dyes proceeds much
more rapidly in the presence of Degussa P25 as
compared with other photocatalysts. The efficiency of
photocatalysts was shown to follow the order:
P25>UV100>PC500>TTP for the degradation of various dyes (Qamar et al. 2005a, b; Tariq et al. 2005,
2008; Faisal et al. 2007). In some cases, the order of
degradation was found to be UV100>P25>PC500
(Saquib and Muneer 2002; Saquib et al. 2008a, b).
The differences in the photocatalytic activity are likely
to be related to the variations in the BET surface,
impurities, existence of structural defects into crystalline frame work, or density of hydroxyl groups on the
catalyst's surface. These factors could influence the
adsorption behavior of a pollutant or degradation
intermediates and the lifetime and recombination rate
of electronhole pairs (Qamar et al. 2005a, b). Lu et al.
2007 compared the photocatalytic degradation of
methyl orange (MO) using natural rutile (rutile 93%)
and P25 (20% rutile) under visible light. After 2 h
irradiation, the degradation efficiency was reported to
be 82.33% and 94.85%, respectively. A possible reason
was the particle size of P25 (30 nm) which is
significantly smaller compared with the natural rutile
sample (7080 m). In suspended form, Lachheb et al.
(2008) reported that, in comparison to PC 500, P25
was more efficient for the degradation of phenols and
poly nitrophenols (4-NP, 2,4-DNP,2,4,6-TNP) in the
presence of either artificial or solar light. The photocatalytic degradation of the tested compounds was
shown in the following order: 2,4,6-TNP>2,4-DNP>4NP>phenol. For PC 500 supported on Ahstrom paper
1048, the order is different: phenol<4-NP<2,4-D,N,
P<2,4,6-TNP. The difference was related to the
variation in adsorption behavior. Khataee et al. (2009)

examined the photocatalytic degradation of three


monoazo anionic dyes (Acid Orane 8, Acid Orange10,
and Acid Orange12) using immobilized TiO2 on glass
plates and UV light. The photocatalytic degradation
rates of tested dyes are shown to follow the order:
AO10>AO12>AO8. This is related to the different
molecular structures of the dyes and their adsorption
behavior. Alinsafi et al. (2007) tested the efficiency of
TiO2 supported either on a glass slide or on a nonwoven glass fiber fabric for the decolorization of textile
wastewater under solar light. Decolorization of the
tested sample was reported to be 2174% with a range
of chemical oxygen demand (COD) removal rate 0.2
0.9 g COD/h/m2. Pare et al. (2008) compared the
efficacy of ZnO, TiO2, and CdS for the photocatalytic
degradation of Acridine Orange. The order of the
degradation rate was shown to be as ZnO>TiO2>CdS.
CdS was reported to be less efficient compared with
ZnO and TiO2, due to its smaller band gaps.
Muruganandham et al. (2006) studied the effect of
various photocatalyst on the decolorization and degradation of Reactive Yellow 14 using UV and solar light
(Muruganandham and Swaminathan 2006a, b). For
both processes, the order of activities of the photocatalysts is shown to be ZnO>P25>TiO2 anatase after
40 min irradiation. CdS, Fe2O3, and SnO2 were
reported to have negligible activity on the degradation
of RY 14 decolorization and degradation. An order of
P25>ZnO>TiO2 anatase has been reported for the
solar photocatalytic degradation of reactive orange 4
(Muruganandham and Swaminathan 2004). Guetta
and Ait Amar (2005) compared the performance of
five catalysts for the degradation of MO. The order
of efficiency was shown to be P25>Pt-UV100>
UV100>Mikro anatase>Prtios AV01. The photocatalytic activity of Degussa P25 was reported to be
higher due to slow recombination between electron and
holes where as Hombikat UV 100 has a high photoreactivity due to fast interfacial electron transfer rate.
The higher photoactivity of P25 has been attributed to
its crystalline composition of rutile and anatase. The
smaller band gap of rutile absorbs the photons and
generates electronhole pairs. Then, the electron
transfer takes place from the rutile CB to electron traps
in anatase phase. Recombination is thus inhibited,
allowing the hole to move to the surface of the particle
and react (Bahnemann et al. 2007). Guillard et al.
(1999) compared the photocatalytic efficiency of
different catalyst with various surface areas, crystallite,

Water Air Soil Pollut

particle size, and their chemical surface for the


degradation of 4-chlorophenol. The observed order of
efficiency based on the rate of degradation upon solar
exposure was shown to be: TiONA PC 10>P25>TiL
COM HC 120>Hombikat UV 100. In the presence of
P25, Selvam et al. (2007) indicated that complete
degradation of 4-flurophenol with P25 was achieved in
60 min while complete degradation occurred in 90 min
in the presence of ZnO. Zheng et al. 2007 compared
the photocatalytic degradation of reactive brilliant red
X-3B in water with pure ZnO, ZrO2ZnO, and P25
TiO2. The degradation of reactive brilliant red over
coupled ZrO2ZnO was shown to be higher compared
with pure ZnO and also slightly better than that of P25
and TiO2. Sobana and Swaminathan (2007) compared
the efficiencies of various catalysts (ZnO, TiO2
anatase, ZnS, SnO2, Fe2O3, and CdS for the degradation of Acid Red 18 under UV irradiation. The highest
degradation was achieved with ZnO over TiO2 anatase
catalyst due to higher surface area of ZnO (10 m2/g)
over TiO2 anatase (8.9 m2/g). SnO2, Fe2O3, CdS, and
ZnS have negligible actively on Acid Red 18 decolorization due to smaller band gap which permits rapid
recombination of hole and electron. Talebian and
Nilforoushan 2009 compared the photocatalytic degradation efficiency of methylene blue (MB) under UV
irradiation. The observed order of photocatalytic efficiency was shown to be SnO2<ZnO<TiO2<In2O3. The
same trend was also observed at different pH values for
the photocatalytic activities of different catalysts due to
their various micro-structures. Zhang et al. 2004
compared the efficiency of coupled oxide ZnOSnO2
with ZnO or SnO2 for the degradation of MO. The
photocatalytic efficiency of ZnOSnO2 was significantly higher compared with ZnO or SnO2 alone.
Cun et al. (2002) compared the efficacy of zinc
sulfide (ZS) and pure SnO2 for the photocatalytic
degradation of MO. The degradation efficiency was
reported to follow the order of Z2S>ZS>ZnO>SnO2.
3.2 UV Light Intensity
The light intensity determines the extent of light
absorption by the semiconductor catalyst at a given
wavelength. The rate of initiation for photocatalysis,
electronhole formation in the photochemical reaction
is strongly dependent on the light intensity (Cassano
and Alfano 2000). Light intensity distribution within
the reactor invariably determines the overall pollutant

conversion and degradation efficiency (Pareek et al.


2008). Consequently, the dependency of pollutant
degradation rate on the light intensity has been
studied in numerous investigations for various organic
pollutants; while in some cases, the rate of reaction
exhibited a square root dependency on the light
intensity, others observed a linear relationship
between the two variables (Terzian and Serpone
1991).Ollis et al. (1991) reviewed the effect of light
intensity on the organic pollutant degradation rate. It
has been reported (Hermann 1999) that the rate is
proportional to the radiant flux for <25 mW/cm2,
and above 25 mW/cm2, the rate has been shown to be
varied as 1/2, indicating a too-high value of the flux
and an increase of the electronhole recombination
rate. At high intensity, the reaction rate is independent
of light intensity. This is likely because at lowintensity reactions involving electronhole formation
are predominant, and electronhole recombination is
insignificant. Kaneco et al. (2004) investigated the
effect of light intensity on the solar photocatalytic
degradation of Bishphenol A in water with TiO2 on
sunny and cloudy days. The degradation efficiency is
shown to increase rapidly with increase in the light
intensity up to 0.35 mW/cm2, and then the efficiency
increased gradually. Muruganandham et al. (2006)
examined the effect of UV radiation power on the
decolorization and degradation of Reactive Yellow
14. Reported results indicated that the decolorization
and degradation increases between 16 and 32 W but
the increase is not significant in the range of 32
48 W, and it is independent of light intensity in the
range of 4864 W. Radiation of 30 W was found to be
the optimum intensity. The increase in light intensity
from 18.8 to58.5 Wm-2 increases the decolorization of
Acid Red 27 from 25% to 42% and the degradation
from 13% to 28% in an 80-cm length immobilized
continuous flow reactor (Behnajady et al. 2007).
Venkatachalam et al. (2007a) compared the mineralization efficiency of 4-CP using the lamps of
wavelength 365 and 254 nm over TiO2. The mineralization rate at 365 nm is reported to be slightly
higher than at 254 nm. The degradation rate constant
of Acridine Orange was shown to increase from 0.018
to 0.0892 min-1 as the light intensity increases from
5,000 to 45,000 LUX (Pare et al. 2008). Chiou et al.
(2008a) studied the effect of UV light intensity (20
400 W) on the phenol degradation. In the UV/TiO2
system, the degradation rate constants with a light

Water Air Soil Pollut

intensity of 20, 100, and 400 W are 8.310-3, 0.012,


and 0.031 min-1, respectively. Under the conditions
tested, an acceptably good linear correlation exists
between the apparent first-order-rate constant and
light intensity. The photocatalytic degradation of Acid
Red 88 using immobilized ZnO was shown to
increase with increase in light intensity (Behnajady
et al. 2009). At higher light intensity, the catalyst
absorbs more photons producing more electronhole
pairs in the catalyst surface, and this increases the
hydroxyl radical concentration and consequently
increases the degradation rate.
3.3 Pollutant Type and Concentration
The successful application of photocatalytic oxidation
system requires the investigation of the dependence of
photocatalytic degradation rate on the substrate
concentration. A summary of various phenols and
dyes studied under different initial concentration is
presented in Table 1. In the presence of TiO2 and
ZnO, the degradation rate of 4-flurophenol was
shown to decrease with the increase in initial

concentration from 0.022 to 0.09 mM (Selvam et al.


2007). Similar trends have been observed for the
photocatalytic degradation of phenol and mnitrophenol (Chiou et al. 2008b), Acid Red 114
(Nikazor et al. 2008), methyl red (Sahoo et al.
2005), Acid Blue 80 under solar irradiation (Su et
al. 2008), Amido Black (Qamar et al. 2005b), and
Acid Orange 7 (Daneshvar et al. 2007). The removal
of bisphenol A was reported to decrease from 100% to
97% as the initial substrate concentration increased
from 0.001 to 0.018 mM, and the removal efficiency
decreased from 97% to 67% with further increase in
BPA concentration to 0.044 mM (Tsai et al. 2009).
Venkatachalam et al. (2007a) observed that the
photocatalytic degradation of 4-CP increases with
increase in initial concentration up to 0.194 mM and
then decreases with further increase in 4-CP concentration from 0.194 to 0.233 mM. This effect has been
associated with the screening effects at 4-CP concentration greater than 0.194 mM. Consistently similar
results have been reported for the photocatalytic
degradation of Chromotrope 2B, Xylenol Orange,
Bromothymol, Acridine Orange, Acid Orange 8,

Table 1 Influence of initial pollutant concentration on the photocatalytic degradation of various pollutants
Pollutant type

Light
source

Photocatalyst

Range of initial
concentration, mM

Optimum concentration,
mM

Phenol

UV

m-Nitrophenol

UV

Chrysoidine Y

Reference

TiO2

0.130.71

0.13

Chiou et al. 2008b

TiO2

0.130.71

0.13

Chiou et al. 2008b

UV

TiO2

0.131.0

0.75

Qamar et al. 2004

Acid Orange 7

UV

ZnO

0.0030.009

0.003

Daneshvar et al. 2007

Remazol Brilliant Blue R

Solar

TiO2

0.120.5

0.12

Saquib and Muneer 2002

Disperse Blue 1

UV

TiO2

0.130.5

0.125

Saquib et al. 2008b

Amaranth

UV

TiO2

0.30.6

0.5

Tariq et al. 2005

Bismarck

UV

TiO2

0.30.6

0.6

Tariq et al. 2005

Reactive Orange 4

UV

F-TiO2

0.0150.035

0.3

Vijayabalan et al. 2009

Acid Blue 45

UV

TiO2

0.30.6

0.3

Tariq et al. 2008

Xylenol Orange

UV

TiO2

0.30.6

0.5

Tariq et al. 2008

Acridine Orange

UV

TiO2

0.10.5

0.25

Faisal et al. 2007

Ethidium bromide

UV

TiO2

0.10.4

0.1

Faisal et al. 2007

Bromothymol

UV

TiO2

0.150.5

0.35

Haque and Muneer 2007

Fast green FCF

UV

TiO2

0.0310.125

0.031

Saquib et al. 2008a

Phenol

UV

Pr-TiO2

0.111.7

0.11

Chiou and Juang 2007

Acid Blue 80

Solar

TiO2

0.030.2

0.03

Su et al. 2008

Acid Red 18

UV

ZnO

0.21.0

0.2

Sobana and Swaminathan 2007

Chromotrope 2B

UV

TiO2

0.250.75

0.35

Qamar et al. 2005b

Amido Black 10B

UV

TiO2

0.250.75

0.25

Qamar et al. 2005b

Water Air Soil Pollut

Amaranth, and Remazol Brilliant Blue R (Saquib and


Muneer 2002). At high pollutant concentration, the
adsorbed reactant molecules may occupy all the active
sties of catalyst surface, and this leads to decrease in
degradation rate (zero-order kinetics). As explained in
several investigations that, as the concentration of
target pollutant increases, more and more molecules of
the compound get adsorbed on the surface of the
photocatalyst. Therefore, the requirement of reactive
species (OH and O2-) needed for the degradation of
pollutant also increases. However, the formation of
OH and O2- on the catalyst surface remains constant
for a given light intensity, catalyst amount, and
duration of irradiation. Hence, the available OH
radicals are inadequate for the pollutant degradation
at higher concentration. Consequently, the degradation
rate of the pollutant decreases as the concentration
increases (Bahnemann et al. 2007). In addition, an
increase in substrate concentration can lead to the
generation of intermediates, which may adsorb on the
surface of the catalyst. Slow diffusion of the generated
intermediates from the catalyst surface can result in the
deactivation of active sites of the photocatalyst and
consequently, a reduction in the degradation rate. In
contrast, at low concentration, the number of catalytic
sites will not be a limiting factor, and the rate of
degradation is proportional to the substrate concentration, in accordance with apparent first-order kinetics
(Selvam et al. 2007). Under the conditions investigated, the degradation rate of Bismarck was reported
to increase with the increase in substrate concentration
(Tariq et al. 2005). This is in accordance with the L-H
law. Several investigations adequately described the
dependence of photocatalytic degradation rates on the
concentration of various organic compounds by the
L-H kinetics model (Galindo et al. 2000; Wenhua et al.
2000; Alaton and Balcioghu 2001). The L-H model
was established to describe the dependence of the
observed reaction rate on the initial solute concentrations (Turchi and Ollis 1990).
3.4 Influence of Catalyst Loading
A number of studies (Daneshvar et al. 2003, 2004,
2006) have shown that the photocatalytic rate initially
increases with catalyst loading and then decreases at
high dosage because of light scattering and screening
effects. The tendency toward agglomeration (particle
particle interaction) also increases at high solid

concentration, resulting in a reduction in surface area


available for light absorption and hence, a drop in
photocatalytic degradation rate. Although the number
of active sites in solution will increase with catalyst
loading, a point will appear to be reached where light
penetration is to be compromised because of excessive particle concentration. The trade-off between
these two opposing phenomena results in an optimum
catalyst loading for the photocatalytic reaction
(Adesina 2004). A further increase in catalyst loading
beyond the optimum will result in non-uniform light
intensity distribution, so that the reaction rate would
indeed be lower with increased catalyst dosage.
Table 2 summarizes the effect of catalyst concentration
on the photocatalytic degradation of various phenols
and dyes in numerous studies. Venkatachalam et al.
(2007a, b) observed that the degradation rate of 4-CP
increases linearly with catalyst concentration up to 2 g/
l and then decreases due to increase in solution
turbidity. Selvam et al. (2007) indicated that the
degradation rate constant of 4-fluorophenol increases
from 0.0152 to 0.0358 min-1 as the concentration of
TiO2 increases from 1 to 3 g/L in the presence of P25.
Furthermore, increase in catalyst loading from 3 to 4 g/
L results in a decrease in the rate constant from 0.0358
to 0.0296 min-1. In the case of ZnO, a similar
observation has been made. The optimum concentrations were found to be 3 g/L TiO2 and 4 g/L ZnO for
efficient removal of 4-fluorophenol. The highest
decolorization and degradation rate of Reactive Orange
16 was achieved at the concentration of 0.4 g/L of
TiO2 (Mahvi et al. 2009).
Sobana and Swaminathan (2007) studied the effect
of ZnO concentration on the decolorization and
degradation rate of Acid Red 18. The optimum
concentration of ZnO was found to be 4 g/l. The
observed low degradation rate at high catalyst loading
may also be due to deactivation of activated molecules
by collision with ground state molecules of titania. The
photocatalytic degradation of Acid Red 88 was found
to increase as the AgTiO2 loading increases from 0.2
to 1.8 g/l (Anandan et al. 2008). Under the conditions
tested, Muruganandham and Swaminathan (2006a)
reported that the decolorization and degradation of
Reactive Orange 4 (RO4) increased as the catalyst
loading increases from 1 to 4 g/l under UVA irradiation.
Consistent results have been observed for the solar
photocatalytic degradation of RO4 (Muruganandham
and Swaminathan 2004). The degradation rate was

Water Air Soil Pollut


Table 2 Influence of catalyst loading on the photocatalytic degradation of various pollutants
Pollutant type

Light
source

Reference
Photocatalyst TiO2 (g/L) Optimum TiO2
concentration, g/L

Reactive Yellow 14

UV

TiO2

1.06.0

Muruganandham and Swaminathan 2006a, b

BPA

Solar

TiO2

01.0

0.5

Kaneco et al. 2004

Chrysoidine R

UV

TiO2

0.55.0

5.0

Qamar et al. 2005a

Cibacron Yellow LS-R

UV

TiO2

0.14.0

4.0

Kositzi et al. 2007

Phenol

UV

TiO2

04.0

2.0

Chiou et al. 2008a

Supra Blue BRL

Visible

K+TiO2

0.252.0

1.5

Chen et al. 2007

Reactive Yellow 14

Solar

TiO2

1.06.0

4.0

Muruganandham et al. 2006

Acid Red 88

UV

ZnO

212

Remazol Brilliant Blue R UV/Solar TiO2

0.55.0

5.0

Disperse Blue 1

UV

TiO2

0.54.0

3.0

Saquib et al. 2008b

Acridine Orange

Visible

ZnO

0.25

Pare et al. 2008

Amaranth

UV

TiO2

0.54.0

1.0

Tariq et al. 2005

0.0.35

12

Behnajady et al. 2009


Saquib and Muneer 2002

Bismarck

UV

TiO2

0.54.0

1.0

Tariq et al. 2005

Acid Orange 8

UV

TiO2

0.55.0

2.0

Saquib and Muneer 2003

Acid Blue 45

UV

TiO2

0.53.0

2.0

Tariq et al. 2008

Xylenol

UV

TiO2

0.53.0

3.0

Tariq et al. 2008

Acridine Orange

UV

TiO2

0.53.0

2.0

Faisal et al. 2007

Bromothymol

UV

TiO2

0.53.0

3.0

Haque and Muneer 2007

Fast green FCF

UV

TiO2

0.54.0

4.0

Saquib et al. 2008a

Reactive Orange 4

UV

FTiO2

1.05.0

4.0

Vijayabalan et al. 2009

Acid Blue 80

Solar

TiO2

0.34.0

2.0

Su et al. 2008

Rhodamine B

Solar

ZnO

0.050.4

0.3

Byrappa et al. 2006

Chromotrope 2B

UV

TiO2

0.55.0

5.0

Qamar et al. 2005b

Amido Black 10B

UV

TiO2

0.55.0

5.0

Qamar et al. 2005b

found to increase with increasing catalyst concentration


up to a level and, on subsequent addition of catalyst,
leads to the leveling off the degradation rate (Kaneco et
al. 2004; Tsai et al. 2009). Regardless of static, slurry,
or dynamic flow reactors, the initial reaction rates were
demonstrated to be directly proportional to catalyst
concentration, indicating a heterogeneous regime
(Bahnemann et al. 2007). However, several researchers
have noticed that, above a certain concentration, the
reaction rate even decreases and becomes independent
of the catalyst concentration. Earlier studies indicated
that this limit depends on the nature of the pollutants,
reactor geometry, and operating conditions of the
photoreactor and the amount of illuminated surface of
TiO2 particles. In the case of high catalyst concentration, turbidity impedes penetration of light within the
reactor. It is assumed that the increase in the number of
TiO2 particles will result in increased number of
photons absorbed and the number of the pollutant

molecule absorbed. Consequently, the degradation


efficiency will be enhanced with increasing TiO2
concentration due to the increase in total surface area
available for contaminant adsorption. Excessive TiO2
dosage leads to opacity of the suspension, which
prevents the catalyst farthest in solution from being
illuminated (Haque and Muneer 2007; Saquib et al.
2008a, b). At high catalyst concentration, particle
particle aggregation may also reduce the catalytic
activity due to the scattering and screening effects
(Adesina 2004). When the concentration of TiO2 is
increased above a certain level, the number of active
sites on the TiO2 surface may become almost constant
because of the decreased light penetration, the increased
light scattering, and the loss in surface area occasioned
by aggregation (particleparticle interactions) at high
solid surface (Hermann 1999). In an immobilized
reactor, McMurray et al. (2006) reported an optimum
TiO2 loading of 1.1 mg/cm2 for the degradation of

Water Air Soil Pollut

atrazine. For any application, the optimum catalyst


concentration has to be determined in order to avoid
excess catalyst (Hermann 1999) and to ensure efficient
absorption of photons.
3.5 Solution pH
Organic compounds in wastewater differ greatly in
several parameters, particularly in their speciation
behavior, solubility in water, and hydrophobicity. While
some compounds are uncharged at common pH
conditions typical of natural water or wastewater, other
compounds exhibit a wide variation in speciation (or
charge) and physico-chemical properties. At pH below
its pKa value, an organic compound exists as a neutral
species. Above this pKa value, organic compound
attains a negative charge. Some compounds can exist
in positive, neutral, as well as negative forms in aqueous
solution. This variation can also significantly influence
their photocatalytic degradation behavior. pH of the
wastewater can vary significantly. pH of aquatic
environment plays an important role on the photocatalytic degradation of organic contaminants since it
determines the surface charge of the photocatalyst and
the size of aggregates it forms( Bahnemann et al. 2007).
The surface charge of photocatalyst and ionization or
speciation (pKa) of an organic pollutant can be
profoundly affected by the solution pH. Electrostatic
interaction between semiconductor surface, solvent
molecules, substrate, and charged radicals formed
during photocatalytic oxidation is strongly dependent
on the pH of the solution. In addition, protonation and
deprotonation of the organic pollutants can take place
depending on the solution pH. Sometimes, protonated
products are more stable under UV radiation than its
main structures (Saien and Khezrianjoo 2008). Therefore, the pH of the solution can play a key role in the
adsorption and photocatalytic oxidation of pollutants.
The ionization state of the surface of the photocatalyst
can also be protonated and deprotonated under acidic
and alkaline conditions, respectively, as shown in the
following reactions:

pH<Pzc

TiOH + H+

TiOH2+

pH>Pzc

TiOH + OH-

TiO-+H2O

5:1

5:2

The point of zero charge (Pzc) of the TiO2


(Degussa P25) is widely investigated/reported at
pH6.25 (Zhu et al. 2005). Table 3 presents a
summary of phenols and dyes investigated under
various solution pH. While under acidic conditions,
the positive charge of the TiO2 surface increases as
the pH decreases (Eq. 5.1); above pH 6.25, the
negative charge at the surface of the TiO2 increases
with increasing pH. Moreover, the pH of the solution
affects the formation of hydroxyl radicals by the
reaction between hydroxide ions and photo-induced
holes on the TiO2 surface. The positive holes are
considered as the major oxidation steps at low pH,
whereas hydroxyl radicals are considered as the
predominant species at neutral or high pH levels
(Shifu and Gengyu 2005; Mathews 1986). It would
be expected that the generation of OH radicals are
higher due to the presence of more available
hydroxyl ions on the TiO 2 surface. Thus, the
degradation efficiency of the process will be logically enhanced at high pH. The degree of electrostatic attraction or repulsion between the
photocatalyst's surface and the ionic forms of
organic molecule can vary with the change in
solution pH, which can result in enhancement or
inhibition on the photocatalytic degradation of
organic pollutants in the presence of TiO2. In the
presence of P25 and ZnO, the effect of pH on the
photocatalytic degradation of 4-fluorophenol was
observed to be pH7>pH9>pH4 (Selvam et al.
2007). As shown elsewhere in (Venkatachalam et
al. 2007a), that the degradation rate of 4-CP was
faster in the acidic pH range compared with the
alkaline pH which was attributed to enhanced
adsorption of 4-CP on the surface of nano-TiO2.
Minimization of electronhole recombination in the
acidic pH was also indicated to be an important
factor for the enhancement of degradation.
Mahmoodi and Arami 2009 tested the effect of pH
(3.510.5) on the photocatalytic degradation of Acid
Blue 25. Under the conditions examined, the optimum
pH was reported to be 10.5 for the degradation of Acid
Blue 25. Using Z2S, Cun et al. 2002 examined the
influence of pH (4.2212.57) on the photocatalytic
degradation of MO and noticed an optimum pH of 4.2.
Soutsas et al. (2009) studied the effect of pH (39)
on the decolorization rates and efficiency of Remazol
Red RR, Remazol Yellow, Procion Crimson H-exl, and
Procion Yellow H-exl. At pH 3, nearly complete

Water Air Soil Pollut


Table 3 Influence of pH on the photocatalytic degradation of phenols and dyes
Pollutant type

Light source

Photocatalyst

pH range

Optimum pH

Reference

Phenol

UV

TiO2

4.112.7

7.4

Chiou et al. 2008a, b

Chrysoidine Y

UV

TiO2

3.09.0

9.0

Qamar et al. 2004

m-Nitrophenol

UV

TiO2

4.112.7

8.9

Reactive Blue 4

UV

NdZnO

3.013.0

11

BPA

Solar

TiO2

2.010.0

Methylene Blue

Visible

La TiO2

210

10

Supra Blue BRL

Visible

K+TiO2

4.511.8

7.2

Chen et al. 2007

3+

Chiou et al. 2008b


Zhou et al. 2009
Kaneco et al. 2004
Parida and Sahu 2008

Reactive Orange 4

UV

FTiO2

1.09.0

3.0

Vijayabalan et al. 2009

Acid Red 88

Visible

AgTiO2

0.21.8

1.8

Anandan et al. 2008

Remazol Brilliant Blue R

Solar

TiO2

3.011.0

3.0

Saquib and Muneer 2002

Disperse Blue 1

UV

TiO2

311

3.0

Saquib et al. 2008b

Methyl Orange

UV

PtTiO2

2.511.0

2.5

Huang et al. 2008

Amaranth

UV

TiO2

3.459.31

7.73

Tariq et al. 2005

Bismarck

UV

TiO2

3.257.85

3.25

Tariq et al. 2005

Acid Orange 8

UV

TiO2

3.011.0

9.0

Saquib and Muneer 2003

Acid Blue 45

UV

TiO2

2.0510.05

5.8

Tariq et al. 2008

Acridine Orange

Visible

ZnO

2.97.1

7.1

Pare et al. 2008

Bromothymol

UV

TiO2

2.29.0

4.35

Haque and Muneer 2007

Fast green FCF

UV

TiO2

3.011.0

4.4

Saquib et al. 2008a

Methyl Red

UV

AgTiO2

3.013.0

310.0

Acid Blue 80

Solar

TiO2

2.010.0

10.0

Su et al. 2008

Acid Red 29

UV

TiO2

310.5

10.5

Qamar et al. 2005a

decolorization (>90%) of all dyes was achieved in only


15 min UV irradiation. Sobana and Swaminathan
(2007) suggested that the decolorization and degradation of Acid Red 18 by ZnO increases as the pH
increases from 3 to 11. At acidic pH, the removal
efficiency is less due to the dissolution of ZnO. AR
18 was efficiently removed at pH 11 due to the
abundance of OH- ions in the particle surfaces as well
as in the reaction medium. Using ZnO and solar light,
the degradation of Rodhamine B was reported to favor
at acidic and basic pH (Byrappa et al. 2006).
Muruganandham and Swaminathan (2004) observed
that the photocatalytic decolorization and degradation of
RO4 was efficient/faster in alkaline pH in comparison to
acidic pH range in both UV and solar irradiation. At low
pH value, TiO2 particle agglomeration reduces the dye
adsorption as well as photon absorption. In RO4 dye,
the azo linkage (-N=N-) was indicated to be susceptible to electrophilic attack by OH radical. Talebian and
Nilforoushan 2009 examined the effect of pH (28) on
the photocatalytic degradation of MB using SnO2,

Gupta et al. 2006

ZnO, TiO2, and In2O3 as catalysts. For all the catalysts


tested, an observed order of pH: pH 8>pH 6>pH 4>pH
2 was found to exist in the degradation of MB.

4 Co-occurring Substances
4.1 Ionic Components
The amount of UV absorption is influenced by water
transmittance over the spectral UV range of interest.
Some common constituents that affect water transmittance are dissolved organic matter, nitrate, and ferrous/
ferric ions. The presence of spectator's components in
water can adversely affect the contaminant degradation
rates. Inorganic anions, such as phosphate, sulfate,
nitrate, and chloride, have been reported to limit the
performance of solar-based photocatalysis (Abdullah et
al. 1990). Bicarbonate, in particular, is detrimental to
reactor performance as it acts as hydroxyl radical
scavenger (Abdullah et al. 1990). Long term experi-

Water Air Soil Pollut

ence with photocatalytic oxidation system showed that


humic substances in contaminated water can adsorb
strongly titanium dioxide particles and reduce activity
toward the target substances (Goswami 1997). Therefore, the incorporation of pre-treatment process has
been indicated to have a beneficial effect on the
performance of photocatalytic oxidation process. Gupta
et al. (2006) studied the influence of various ionic
components on the degradation of crystal violate (CV)
and methyl red (MR) dyes using AgTiO2 as photocatalyst. Cl-, NO3-, SO42-, HPO42-, and humic acid
were shown to have negligible effect on the degradation rate, whereas Ca2+ and Fe2+ interfered significantly. In case of the MR, Cl-, Ca2+, and Fe2+ were
reported to affect the degradation with the exception of
NO3-, SO42-, and HPO42-(Sahoo et al. 2005). The
effect of transition metal ions on the degradation of
4-flurophenol was shown to be in the order of Mg2+>
Fe3+>Fe2+>Cu2+, and the inhibition of inorganic anions
on the degradation of 4-FP was demonstrated to be
CO32->HCO3->Cl->NO3->SO42- (Selvam et al. 2007).
Muruganandham et al. (2006) observed that the
addition of 0.471.89 mM Na2CO3 decreases the
degradation of Reactive Yellow 14 from59.2% to
48.2% due to hydroxyl radical scavenging property of
carbonate ion.

CO32-+.OH

OH-+CO3.-

6:1

HCO3-+.OH

H2O+CO3.-

6:2

As a result, the primary oxidant hydroxyl radical


decreases gradually with the increase in carbonate
ion, and consequently, there is significant decrease in
photocatalytic degradation. Addition of Cl- ion from
0.47 to 1.89 mM Na2CO3 to reaction decreases the
degradation from 80.1% to 72.3% by time of 80 min.
The decrease in degradation efficiency is due to the
hole scavenging and hydroxyl radical scavenging
properties of chloride ion.

Cl-+hVB+

Cl.

6:3

Cl. +Cl-

Cl2.-

6:4

The inhibitive effect of CO32- ion is shown to be


greater than the inhibitive effect of Cl- ion. Lin and Lin

(2007) observed that the presence of humic acid caused


a significant retardation on the photocatalytic degradation of 4-chlororphenol. The observed retardations of
humic acids were related to the inhibition (surface
deactivation), competition, and light attenuation
effects. Moreover, the presence of humic acid in the
reaction mixture has been reported to significantly
reduce light transmittal and therefore, the photooxidation rate. Humic acid can also compete with organohalides for the active sites on the TiO2 surface. Mahvi
et al. (2009) studied the effect of anions on the
photocatalytic decolorization Reactive Orange 16.
The observed order of negative impact of anions were
shown to be as SO42-<NO3-<Cl-<HCO3-<CO32-.
Naeem and Feng (2008) studied the inhibition effect
of anions on the photocatalytic degradation of phenol,
which is reported to be Cl->SO42->NO3->CO32-. At
certain concentration, Fe3+ is reported to improve the
degradation rate, whereas Ca2+, Cu2+, and Mg2+ hinder
the degradation process. The effect of surfactants
(0.05 mmol/L) on the degradation rate of phenol was
reported to be sodium dodecyl benzene sulfate>sodium
dodecyl sulfonate>sodium dodecyl sulfate (SDS). The
presence of anions is reported to alter the ionic strength
of the solution, and therefore, influence the catalytic
activity, and hence the photocatalytic degradation
(Calza and Pelizzetti 2001). The reaction of hvb+ and
.
OH with anions which behave as scavengers and thus
inhibit the degradation. Pare et al. (2008) investigated
the effect of carbonate ions on the photocatalytic
degradation of Acridine Orange for a range of
concentration from 2.010-3 to 810-3 mM. The
degradation rate constant was shown to decrease from
0.023 to 0.015 min-1 under the above concentration
range. In the presence of NaCl, similar degradation
trends have been reported for a certain concentration
range of 2 to 10 mM due to the hole scavenging
properties of these ions. The presence of anionic (SDS)
and cationic surfactants (C16TAB) retarded the rate of
degradation due to the preferential adsorption of the
amphiphilic surfactant molecules on the photocatalyst
surface. Epling and Lin (2002) examined the inhibition
effect of humic substances and inorganic anions on the
degradation of several cationic and anionic dyes. Of
the ionic species (Cl-, NO3-, PO43-, and HCO3-)
studied, Cl- showed the strongest inhibition effect on
both cationic and anionic dyes, followed by PO43-. The
overall retardation has been attributed to the surface
deactivation mechanism. A significant retardation on

Water Air Soil Pollut

the degradation of MB in the presence of either Aldrich


humic acid or natural humic substances has been
reported. Guillard et al. (2003) studied the inhibition of
inorganic salts on the photocatalytic degradation of
MB at neutral and basic pH. Under the conditions
tested, the order of inhibition was reported to be
CO32->PO43->SO42->Cl->NO3-. In the presence of
200 M nitrate salt, the effect of metal ions on the
photocatalytic degradation of Azure A was observed to
be Cu2+>Al3+>Co2+>Zn2+, and the order was shown to
be Cu2+>Co2+>Fe3+ in the presence of 200 M
chloride salt (Aarthi et al. 2007). Papadam et al.
(2007) examined the effect of water matrix on the
decolorization of Acid Orange 20. The presence of
17.1 mM NaCl or 7.04 mM Na2SO4 led to decreased
decolorization which was attributed to the trapping of
photogenerated valence band holes and the hydroxyl
radicals by the respective anions. Aarthi and Madras
(2007) investigated the effect of metal ions (Cu2+,Fe3+,
Zn2+, and Al3+) on the photocatalytic degradation of
Rhodamine B and was shown to follow the order of
Cu2+>Al3>Zn2>Fe3+. In the presence of Cu2+ and
Zn2+, the degradation of Malachite Green was reduced
by 76% and 66%, respectively.
4.2 Solvent Types
Organic solvents are mostly present in many industrial
wastewaters. Therefore, their effects on the photocatalytic degradation system have been investigated in
several studies. Lin and Lin 2007 examined the effects
of acetonitrile or isopropanol on the photocatalytic
oxidation of 4-chlorophenol. The degradation rate of
4-chlorophenol was shown to decrease from 1.0436 to
0.0525 h-1 as the acetonitrile content increases in the
reaction mixture. A similar retardation effect was also
observed when 5% isopropanol was added into the
acetonitrile/water (1:1) reaction mixture. Behnajady et
al. (2009) observed that the pseudo first-order-rate
constants of Acid Red 88 and water ethanol were
0.0257 and 0.0271 min-1, respectively. The use of
ethanol instead of water as a solvent did not show any
significant effect on loading morphology of ZnO on
glass plate. Shohrabi et al. (2009) studied the effect of
various solvents on the photocatalytic degradation of
Benzidene Yellow. The observed order of solvents
effects on the degradation efficiency are reported to be
t-butyl alcohol>n-butanol>2-propanol>acetonitrile>Nhexane. Of the alcohols, t-butyl alcohol was shown to

have the most inhibitive effect due to the spatial


hindrance effect of the t-butyl alcohol. In the presence
of ethanol and methanol, Aarthi et al. (2007) investigated the effect of solvents and mixed-solvents systems
on the photocatalytic degradation of dye Sudan III. As
the ethanol content in the system increases, the
degradation rate was reported to decrease. This effect
was attributed to the scavenging of OH radicals by
these solvents. Liu et al. (2009) studied the effect of
solvent type on the photocatalytic activity of N-doped
TiO2 under visible light irradiation for the degradation
of MO. Under the conditions tested, it was observed
that the anatase gradually changed into rutile with
increase in carbon chain of methanol, ethanol,
1-propanol, and 1-butanol as the solvent under
solvothermal conditions. The bicrystalline TiO 2
(19.5% rutile and 80.5% anatase) prepared in methanol
showed the highest photocatalytic activity for the
degradation of MO. Aarthi and Madras (2007) examined the effect of solvent effects on the photocatalytic
degradation of rodamine B (RB) using combustionsynthesized TiO2. In the mixture of ethanol and water,
when ethanol was 10% by volume, the rate decreased
by 91.5%, and when the ethanol increased to 50% by
volume, no photodegradation was observed. Similarly,
the rate decreased by 76%, 96%, and 100% when
acetonitrile in the mixture of water and acetonitrile was
increased from 10%, 20%, and 50% by volume,
respectively. This effect is attributed to the lesser
solvation of excited electrons in the organic solvents
compared with the aqueous solution.

5 Influence of Oxidants/Electron Acceptor


The electronhole recombination is one of the main
drawbacks in the application of TiO2 photocatalysis
as it causes waste of energy. In the absence of suitable
electron acceptor or donor, recombination step is
predominant, and thus, it limits the quantum yield.
Thus, it is crucial to prevent the electronhole
recombination to ensure efficient photocatalysis.
Molecular oxygen is generally used as an electron
acceptor in heterogenous photocatalytic reactions.
Addition of external oxidant/electron acceptors into
a semiconductor suspension has been shown to
improve the photocatalytic degradation of organic
contaminants by (1) removing the electronhole
recombination by accepting the conduction band

Water Air Soil Pollut

electron; (2) increasing the hydroxyl radical concentration and oxidation rate of intermediate compound;
and (3) generating more radicals and other oxidizing
species to accelerate the degradation efficiency of
intermediate compounds (Qamar et al. 2005a, b;
Saquib and Muneer 2002; Saquib et al. 2008a, b;
Tariq et al. 2005, 2008; Faisal et al. 2007; Haque and
Muneer 2007). Since hydroxyl radicals appear to play
an important role in the photocatalytic degradation,
several researchers have investigated the effect of
addition of electron acceptors such as H2O2, KBrO3,
and (NH4)2S2O8 on the photocatalytic degradation of
various dyes and phenolic compounds (Qamar et al.
2005a, b; Saquib and Muneer 2002; Saquib et al. 2008a,
b; Tariq et al. 2005, 2008; Faisal et al. 2007; Haque and
Muneer 2007) to enhance the formation of hydroxyl
radicals as well as to inhibit the electron/hole(e-/h+) pair
recombination. In all cases, the addition of oxidants has
resulted in higher pollutant degradation rate compared
with the molecular oxygen. In most of the cases, the
order of enhancement is reported to be UV/TiO2/
H2O2>UV/TiO2/BrO3->UV/TiO2/S2O82-. The enhancement of degradation rate is due to the reaction between
BrO3- and conduction band electron. This reaction
reduces the recombination of electronhole pair.

BrO3-+6e-CB+6H+

Br-+3H2O

7:1

S2O82- can generate sulfate radical anion (SO4-)


both thermally and photolytically in aqueous solution.
SO4- then reacts with H2O to produce .OH radicals.

S2O82-

.SO4 + H2O

.2SO4

.
OH+ SO42 + H+

7:2

7:3

With the addition of H2O2, the enhancement of


degradation is due to the increase in the hydroxyl
radical concentration as shown by Eqs. 7.4 and 7.5:

H2O2+eCB-

OH+OH-

7:4

7:5
The degradation efficiency of UV/TiO2/oxidant
process is slightly more in acidic medium than in

basic medium. Muruganandham et al. (2006) examined


the effect of H2O2, KBrO3, and (NH4)2S2O8 addition
on the solar photocatalytic degradation of Reactive
Yellow 14 by varying the amount of oxidant concentration. The optimum concentrations of the oxidants
were 15 mM, 0.90 mM, and 0.88 mM for H2O2,
KBrO3, and (NH4)2S2O8, respectively. Of the oxidants
studied, (NH4)2S2O8 was found to be efficient to
oxidize the Reactive Yellow 14 azo dye. Under visible
light irradiation, addition of optimum amount of H2O2
and K2S2O8 was shown to increase the degradation rate
of Acridine Orange in the presence of ZnO photocatalyst (Pare et al. 2008). Sobana and Swaminathan
2007 observed that addition of KBrO3 and (NH4)2S2O8
was shown to increase the photocatalytic degradation
of Acid Red 18 by ZnO. However, the addition of H2O2
was reported to decrease the photocatalytic degradation
rate which was attributed to its low adsorption on the
ZnO surface. Both decolorization and degradation of
Reactive Yellow 14 by UV/TiO2 system were reported
to increase in the presence of H2O2, (NH4)2S2O8, and
KBrO3, up to a certain dosage beyond which the
enhancement was negligible (Muruganandham and
Swaminathan 2006b). The effect of oxidants on the
degradation of 4-flurophenol was shown to be in the
order of IO4->BrO3->S2O82->H2O2>ClO3- (Selvam et
al. 2007). After 15 min irradiation at pH 4, the
degradation of 4-flourophenol were observed to be
79.86%, 66.89%, 60.88%, 53.85%, and 46.35% in UV/
TiO2/IO4-, UV/TiO2/BrO3-, UV/TiO2/H2O2, UV/TiO2/
S2O82-, and in UV/TiO2/ClO3- processes, respectively.
IO4- was found to be most efficient oxidants due to
the generation of a number of highly reactive
intermediate radicals such as IO3., OH., and IO4.. The
optimum concentration of H2O2 addition is found to be
about 1.2 mM/L for the degradation of MO in the
presence of 0.5 wt.% PtTiO2/Zeolite systems (Huang
et al. 2008). Under visible light irradiation, the
influence of oxidants on the degradation of Acid Red
88 by AgTiO2 was found to be peroxomonosulfate>
peroxodisulfate>hydrogen peroxide (Anandan et al.
2008). Chiou et al. (2008a) investigated the effect of
H2O2 addition from 1.77 to 88.2 mM on the
photocatalytic degradation of phenol. The addition of
H2O2 from 1.77 to 8.82 mM leads to an increase in
removal efficiency from 58% to 84% within 3 h. In
contrast, phenol is completely degraded within 2.5 and
1 h when the level of H2O2 increases to 44.1 and
88.2 mM, respectively.

Water Air Soil Pollut

6 Combined Approach
In recent years, photocatalysis has been combined with
other physical and chemical means of treatment to
improve the efficiency of photocatalysis. Generally,
integration of several methods provides high treatment
efficiency compared with individual treatment. TiO2 has
been combined with ozonation, membrane filtration,
and ultrasound technology for the treatment of various
organic compounds. Chen and Smirniotis (2002)
indicated the enhancement of photocatalytic degradation of phenol and chlorophenol by ultrasound, UV
irradiation with photocatalysis results in 76% destruction in 3 h of treatment. Combination of ultrasound and
photocatalysis results in almost complete destruction of
phenol in about 150 min. Synergistic effects observed
in combined approach are attributed to desorption of
organic substrate and degradation by products from the
catalyst surface and de-agglomeration of the catalyst.
Mrowetz et al. (2005) reported that the degradation rate
constant of Acid Orange 8 in ultrasonic irradiation and
the combined technique was 0.1, 1.81, and 3.329
10-4 s-1, respectively, clearly indicating the synergistic
effects. For Acid Red 1 and 2 chlorophenol, the rate
constants in same order are 0.038, 1.22, and 2.89
(10-4s-1) and 0.59, 0.88, and 1.58(10-4s-1), respectively. Li et al. (2003) reported that the degradation rate
constants of reactive Brilliant Red X-3B using TiO2/
UV/O3 system is shown to be 2.563.2 times higher in
comparison to the maximal rate constants of TiO2/UV/
O2, 4.689.8 times of maximal rate constants with
TiO2/UV. For the degradation of catechol, the order of
removal efficiency was demonstrated to be O3>UV/TiO2/
O3>UV/TiO2/O2>UV/TiO2. Within 6 min reaction time,
the decolorization efficiency of Reactive Red 2 was
significantly higher in the case of UV/TiO2/O3 system
compared with UV/TiO2 system (Wu et al. 2008).
Oyama et al. (2009) indicated significant mineralization
of bisphenol A in UV/TiO2/O3 system as compared with
either ozonation or UV/O3 and to the UV/TiO2/O2
system. Fernando et al. (2005) reported that the
degradation rate constant of photocatalytic ozonation
process was higher compared with TiO2 photocatalysis
for phenol, p-chlorophenol, and p-nitrophenol. Photocatalytic ozonation process exhibited the highest mineralization efficiency of phenols. Zou and Zhu (2008)
investigated the comparative efficacy of UVA/TiO2
and UVA/TiO2/O3 system for the removal of color,
A254, and total organic carbon (TOC). By means of

TiO2/UVA process, color removal was 33% after


40 min treatment. For UVA/TiO2/O3 system, color
removal was greater than 50% with a significant
removal within first 10 min. The highest color removal
rate (70.2%) was achieved by the combined treatment of
photocatalytic oxidation and ozonation. The cotreatment of ozonation and photocatalytic oxidation
increased TOC removal rate by nearly 50% compared
with the process when ozone was used as a pretreatment for photocatalysis. Mozia et al. (2009)
compared the effectiveness of photocatalysis and hybrid
process photocatalysis-membrane distillation (MD) for
the degradation of Acid Red 18. The degradation rate
constants were higher for 10% or less than during the
hybrid process. This effect was attributed to the
concentration of feed solution during hybrid process.
Grzechulska-Damszel et al. (2009) studied the effectiveness of TiO2 photocatalysisnanofiltration (NF) membrane system for the degradation of Acid Red 18.
However, the integration of post-NF membrane system
did not result in complete removal of organic matter
from solution when catalyst was used in immobilized
form. Suspended TiO2/UF/MD system was reported to
be efficient for the removal of organic matter from the
solution treated by TiO2 photocatalysis compared with
the TiO2 photocatalysisNF system. The ion rejection
efficiency was reported to be similar for both systems.
Gonzalez and Martinez (2008) compared the efficiency
of sonophotocatalytic system using nano-size anatase
and rutile TiO2 powders with ultrasound alone for the
degradation of Methyl Orange. Under the conditions
studied, the pseudo first-order-rate constants of the three
cases were shown to be 0.0179, 0.0068, and
0.0017 min-1, respectively. The sonophotocatalytic
activity of nano-size anatase was found to be efficient
compared with nano-rutile TiO2 powder. Bejarano-Perez
and Suarez-Herrera (2007) reported that the kinetic
constant of the Congo Red oxidation is 2.6 times
higher than the one obtained in the absence of
ultrasound and in the presence of UV light and TiO2
P25. It was also observed that the degradation rate of
MO in the presence of both ultrasound and UV light
is significantly higher compared with either sonocatalytic or photocatalysis. Zhang et al. (2006) studied
the combined photocatalytic membrane technique for
the removal of Direct Black 168 dye. The removal of
Direct Black 168 with UV irradiation alone was
achieved at 70%, and the retention of the dye with
membrane filtration alone was reached at 67% after

Water Air Soil Pollut

300 min in the continuous system. However, when


the photocatalysis system was combined with membrane techniques, the removal efficiency of the dye
was increased to 82% after 300 min under the
conditions investigated. Wang et al. (2008) compared
the sonophotocatalytic degradation of Methyl Orange
using Degussa P25 (55.07 m 2 /g), Yilli TiO 2
(10.45 m2/g), and AgTiO2 (Ag loaded on Yili
TiO2). The degradation rate constants obtained with
combined approach were reported to be 5.810-3,
1.1710-2, and 3.5810-2 min-1, respectively for Yill
TiO2, P25, and AgTiO2. However, the degradation
rate constants were shown to be 3.210-3, 9.810-3,
and 2.5910-2 min-1 in photocatalysis in the same
order. Reported results indicated that the combination
of sonolysis and photocatalysis enhanced the MO
degradation. Comparatively, the photocatalytic activity of AgTiO2 nanoparticles was much higher than
that of Degussa P25 and Yilli TiO2. At pH 7, the
decolorization rates of Reactive Red 2 were shown to
follow the order of UV/US/TiO2(0.94 h-1)>UV/
TiO2(0.85 h-1)>US/TiO2(0.25 h-1)>US (Wu and Yu
2009).

7 Film Fabrication Technique and Calcination


Temperature
Nanocrystalline titanium oxide powder for catalytic
purposes can be applied by a variety of techniques
including solgel method, solvothermal process, reverse
miscellar, hydrothermal method, and electrochemical
methods. There have been several investigations on
photocatalytic degradation of various dyes by titania
nanoparticles; however, information on the titania film
fabrication routes on the photocatalytic degradation of
dyes is still desirable. Song et al. (2009) investigated the
effect of titania films with different nanostructures of
nanorods (NR), solgel film, nanotubes (NT), and
nanoparticle aggregates on the photocatalytic degradation of RB, Methylene Blue, and Methyl Orange. In
the case of RB degradation, the observed order was
reported to be NR>DP>SG>NT, and the order was
shown to be NR>SG>DP>NT for the degradation of
MB. Nonetheless, an order of NT>DP>SG>NR was
found to observe for the degradation of MO.
Bessekhouad et al. (2004) studied the effect of alkaline
(Li, Na, and K) doped TiO2 prepared by solgel route
and impregnation method on the photocatalytic degra-

dation of malachite green oxalate. The crystallinity


levels of catalysts are found to be largely dependent on
both the nature and the concentration of alkaline. The
best crystallinity is obtained for Li-doped TiO2 and is
lowest for K-doped TiO2. For a given alkaline
concentration, the catalyst prepared by the impregnation technique was shown to be more efficient than
those prepared by solgel route. For 5% Li-doped TiO2
catalyst prepared by impregnation method, the half life
of malachite green is reported to be twice shorter than
that observed for P25 TiO2. However, the catalyst
prepared by solgel route, the presence of Li reduces
the catalyst photoactivity. Indeed, the photoactivity of
undoped catalyst is shown to be better than Li-doped
TiO2. Structure and size of TiO2 crystallites are
significantly dependent on calcinations temperature.
Thermal treatment of TiO2 gels at higher temperature
promotes phase transformation from thermodynamically metastable anatase to more stable and condensed
rutile phase. As the dehydration process occurs during
heat treatment, crystallites grow to dimensions larger
than those of the original particles. Pecchi et al. (2001)
studied the effect of calcinations temperature on the
photocatalytic degradation of pentachlorophenol using
solutiongel method under different pH. Increasing
calcinations temperature results in reduced surface area
and increased crystallinity and particle size. The
degradation rate constant was shown to increase as
the calcinations temperature increased from 300C to
500C for pH values 3 and 5. Nonetheless, the rate
constant was higher at 300C and 500C compared
with 400C at pH 9. Peng et al. (2005) noticed that
compared with undoped sample and P25, highest
degradation of RB was achieved using La-doped
TiO2 at optimum calcinations temperature of 300C.
This effect was ascribed to larger surface area due to
the presence of lanthanum ion in the high thermal
stable mesostructures TiO2. However, the degradation
rate was shown to decrease as the calcinations
temperature increases from 300C to 700C. The
reduction of photocatalytic activity was reported to be
related with the decrease in specific surface area from
460.6 to 102.1 m2/g. After calcinations at 900C, it
was also indicated that the anatase crystallites begin to
grow extensively and transform to rutile phase and
then segregate from mesostructures. Subsequently, the
whole mesostructure is completely destroyed (Chen et
al. 2007). The surface areas are in the decreasing order
of TiO2>4.6% K+ doped>14.3% K+ doped, 4.6% K+

Water Air Soil Pollut

doped>TiO2>14.3% K+ doped, 4.6% K+ doped>14.3%


K+ doped>TiO2, and 14.3%K+ >4.6%K+>TiO2 when
the samples were calcined at 400C, 600C,700C, and
850C, respectively. Similar trends of the dependence
of pore volume, micropore volume, external surface
area, and micropore surface area on the K+ content and
calcination temperature are reported. Of all the samples
tested, 4.6% K+ doped TiO2 calcined at 700C was
shown to be efficient for the degradation of Supra Blue
dye under the conditions studied. Behnajady et al.
(2009) investigated the effect of furnace temperature
on the photoactivity of immobilized ZnO for the
degradation of Acid Red 88 for range of 400C to
500C. Degradation rate was shown to increase from
0.0257 to 0.0208 min-1 with increasing temperature
from 400C to500C, which is attributed to the
increase in particle size from 42 to 67 nm and thereby
reduced availability of surface area. Zheng et al. (2007)
studied the effect of calcinations temperature on the
photocatalytic degradation of reactive Brilliant Red
X-3B for a range of 250550C. Under the conditions
investigated, the highest degradation was achieved at
350C. Using a carbon doped-TiO2, Xiao and Ouyang
(2009) observed that the reduction in surface area from
20 to15.5 m2/g is due to the increase in calcination
temperature from 500800C. This would lead to the
decrease of active sites on which reactants could
adsorb. Under the conditions tested, decolorization of
MB was reported to decrease from 49.16% to 39.16%
with increasing calcination temperature. The order of
photocatalytic activity and the formation rate of .OH
radicals per unit surface area was shown to be closely
related. An optimum temperature of 600C was
also noticed for the photocatalytic degradation of
Methylene Blue and the formation rate of .OH radicals.
Wen et al. (2009) indicated that no significant changes
on the photocatalytic degradation of MB were observed as the calcination temperature increases from
500C to 600C. The degradation rates were higher
in the case of Co doped TiO2 calcined at 500C and
600C, compared with that of one calcined at 700C.
This effect was attributed to the presence of larger
surface area. It is well-known from the literature that
calcination temperature is an important factor that
probably influences the crystalline sizes, morphology,
and surface area of TiO2 which can clearly affect the
photocatalytic activity. Lim et al. (2007) reported that
the BET surface area of C-doped TiO2 decreased from
200 m2/g (200C) to 80 m2/g (400C). A rapid

decrease of specific surface area from 300C onwards


is a consequence of crystallization and growth of
anatase. At calcination temperature above 700C, the
specific surface area was <1 m2/g. No discernible
changes were observed in the visible light activity.
Baiju et al. (2007) indicated that the crystallite size of
TiO2 increases as the calcinations temperature increases
for both pure and Ta2O5 doped-TiO2. For any given
temperature, crystallite size TiO2 decreases as the
amount of Ta2O5 increases from 1 to 10 mol%. All
the Ta2O5-added titania was shown to have higher
surface area than pure titania. The addition of Ta2O5
was reported to increase the anatase phase stability
above 1,000C by inhibiting the increase in crystallite
size of titania. The optimum calcinations temperature
was found to be 500C for the degradation of MB.
Silveyra et al. (2005) tested the effect of calcinations
temperature (600800C) on the photocatalytic
degradation of phenol using N-doped TiO2. Highest
degradation was obtained at 600C and no variation
was observed at 700C and 800C. Huang et al. (2008)
examined the effect of calcinations temperature on the
decolorization rate of Methyl Orange solution under
UV irradiation. In case of the undoped TiO2, the
decolorization rate decrease from 73% to 31% as the
calcination temperature increases from 110C to 500C.
In contrast, the decolorization rate decreases from
87.8% to 43.2% with increase in temperature for the
above range in the presence of 1.5% pt-TiO2 catalyst.
This effect was related to the formation of TiO2 anatase
phase and conglomeration of catalyst particles which
reduces the specific surface areas of samples, resulting
in the decrease of photocatalytic activity at lower
calcinations temperature. Wawrzyniak et al. (2006)
studied the effect of calcination temperature (100
800C) on the photocatalytic degradation of phenol and
two azo dyes (Reactive Red, Direct Green) by N-doped
TiO2 under visible light irradiation. The highest
effectiveness of phenol degradation was achieved at
calcinations temperature 700C. In case of degradation
of both azo dyes, the highest degradation was observed
at 500C and 600C. Temperature of 600C was
reported to be the optimum calcinations temperature
for the highest photocatalytic degradation of phenol
using 5 mol% La-doped TiO2 (Liqiang et al. 2004). In
the presence of STiO2ZrO2, the optimum calcinations temperature was 500C for the degradation (85%)
of rodhamine B in the visible irradiation. The optimum
calcinations temperature was shown to be 400C when

Water Air Soil Pollut

STiO2 was the visible light catalyst (Tian et al. 2009).


Talebian and Nilforoushan (2009) studied the effect of
annealing temperature (250550C) on the photocatalytic activity of In2O3 film for the degradation of MB.
The optimum annealing temperature was 550C, and
the activity of In2O3 film was directly related to degree
of crystallinity at higher temperature. Gorska et al.
(2009b) studied the effect of calcinations temperature
(350750C) on the photocatalytic degradation of
0.21 mM phenol under UV (25<<400 nm) and visible
(>400 nm) light irradiation. After 60 min irradiation,
maximum amount of phenol degradation was about
80% when P25 TiO2 calcinated at 350C in visible
light. The sample calcined at 350C was shown to have
the largest BET surface area (205.8 m2/g), the smallest
crystallite size (8.4 nm) and an average pore diameter
of 8.3 nm of TiO2. This effect was attributed to the
efficient absorption of light vis region due to the
existence of largest amount of carbon in aromatic C-C
bonds (10.1 at %). However, the photocatalyst calcinated at 450C was found to have the highest activity
in UV light irradiation. Lettmann et al. (2001) observed
that, about 30% 4-CP (Co =0.25 mM) was degraded
after 100 min irradiation (>400) in the presence of a
catalyst-obtained hydrolysis, followed by calcination at
250C for 3 h. Using a carbon doped TiO2 prepared by
TiCl4 hydrolysis in tetrabutylammonium hydroxide
(calcinations at 400C, 1 h), 70% TOC reduction of
0.25 mM 4-CP was reported by Shakthivel et al.
(2004). Porter et al. (1999) examine the microstructural
changes in Degussa P25 due to heat treatment. The
TiO2 powder was annealed from 600C to 1,000C.
Under UV irradiation, the apparent crystallite size and
rutile content of catalyst increased with increasing
calcination temperatures, whereas the specific surface
area and the rate of phenol degradation was reported to
decrease. Sonawane and Dongare (2006) tested the
effect of calcinations temperature (275500C) on the
photocatalytic activity of 2%-Au/TiO2 for the degradation of phenol under solar light. The catalyst calcinated
at 275C exhibited the highest activity compared with
the sample calcinated at higher temperature. Using ZnO
upon solar irradiation, the photocatalytic degradation of
4-nitrophenol increases from 52.9% to 74.4% as the
calcinations temperature increases from 110C to 300C
due to higher surface area and lowest crystallite sizes
(Parida et al. 2006). Zhang et al. (2004) studied the
effect of calcinations temperature (300900C) on the
photocatalytic degradation of MO using ZnOSnO2.

At 350C, the higher photocatalytic activity of coupled


oxides was reported due to variation in phase
composition and particle size. Using Z2S, an optimum
temperature of 600C was found for the degradation of
MO (Cun et al. 2002).

8 Doping and Mixed Semiconductor


The recombination of photogenerated electrons and
holes are believed to be reason for the low photoactivity of TiO2. Structural imperfections in the TiO2
lattice generate trap sites which may act as recombination centers, leading to a decrease in the levels of
electrons and holes. In order to enhance the TiO2
photocatalysis as well as the response into the visible
spectrum of solar light, TiO2 has been doped with
certain transition metals, non-metals, and ionic components. Doped ions can also act as charge trapping
sites and thus reduce electronhole recombination.
The effect of doping on the activity of photocatalyst is
governed by several factors, e.g., the type and
concentration of dopant, preparation method, the
structure and the initial concentration of the pollutants
(dyes and phenols), and physico-chemical properties
of the catalyst. Both positive and negative results
have been reported from doping with metal ions. The
increase in charge separation efficiency will enhance
the formation of both free hydroxyl radicals and
active oxygen species (Kato et al. 2005). The amounts
of doping concentration along with a summary of the
dyes and phenolic compounds degraded using doped
catalyst are shown in Table 4. Upon 45 min UV
irradiation, >99% degradation of Methyl Red was
reported with Ag doped TiO2 whereas only 85% was
degraded in the presence of TiO2 (Sahoo et al. 2005).
Liu et al. (2005) compared the photocatalytic activities of N-doped TiO2 and Degussa P25 for the
degradation of three azo dyes: AO7, MX-5B, and
RB5 in the presence of UV and solar light. Some 95%
decolorization of AO7 was reported to occur in 1 h
under UV illumination and the doped TiO2 was
reported to be four times more efficient compared
with bare TiO2. In addition, 70% dye removal was
achieved within 3 h whereas no detectable dye
decolorization was reported to obtain under solar
irradiation. Under the tested conditions, 0.2 mmol
fluoride ion is shown to be the optimum concentration
for the efficient removal of Reactive Orange 4

Water Air Soil Pollut


Table 4 The influence of dopant concentration on photocatalytic activity of photocatalyst
Pollutant

Light source

Photocatalyst

Doping (%)

Optimum doping (%)

Reference

Methylene Blue

Visible

C-TiO2

03.2

1.25

Wong et al. 2008

Supra blue BRL dyes

Visible

K+TiO2

014.3

4.6

Chen et al. 2007

XRG dyes

UV/Visible

Fe3+/Fe2+TiO2

0.030.15

0.09

Zhu et al. 2004

Methyl Orange

Solar

Eu2+TiO2

0.51.0

1.0

Yi et al. 2007

Methyl orange

Solar

Gd3+TiO2

0.11.0

0.5

Yi et al. 2007

p-chlorophenol

UV

Ag TiO2

0.10.5

0.5

Kirilov et al. 2006

p-chlorophenol

UV

Pd2+TiO2

0.10.5

0.5

Kirilov et al. 2006

Methelyne Blue

Visible

Pd2+TiO2

05.0

5.0

Chan et al. 2009

Phenol

Visible

Fe3+TiO2

0.45.1

1.0

Adan et al. 2007

Methyle Orange

UV

Mo6+TiO2

01.5

1.0

Yang et al. 2004

4-Chlorophenol

Visible

Ce4+TiO2

01.0

0.6

Silva et al. 2009

Methylyne Blue

UV

V5+TiO2

0.255.0

0.5

Tian et al. 2009

2,4-Dichlorophenol

Visible

V5+TiO2

0.255.0

1.0

Tian et al. 2009

Methelyne Orange

UV

Pt6+TiO2

03.0

1.5

Huang et al. 2008

(Vijayabalan et al. 2009). Ag and Pd modified TiO2


were shown to have improved photocatalytic activity
for the degradation of p-chlorophenol compared with
Degussa P25 which was coupled with better charge
separation and therefore, to slower recombination
(Kirilov et al. 2006). Under the conditions tested,
Gupta et al. (2006) reported that more than 90%
degradation of CV and MR mixtures was achieved in
the presence of Ag doped TiO2 compared with only
about 70% with TiO 2 . Anandan et al. (2008)
compared the effectiveness of 8 wt.% AgTiO2 for
the degradation of Acid Red 88 in the presence of
electron acceptors such as under visible light
illumination. In comparison to TiO2, the nano-sized
AgTiO2 exhibited seven times higher photodegradation rate in the presence of peroxomonosulfate while
2.7 times higher in the case peroxodisulfate. Andronic
and Duta (2007) noticed that optimum photocatalytic
behavior of Cd doped-TiO2 was found to be at 0.1 at.
% for cadmium acetate and 0.5 at.% for cadmium
chloride for the degradation of MO and MB,
respectively. Carbon doped TiO2 was reported to be
efficient for the degradation phenol under visible light
irradiation (Lee et al. 2008). This effect was attributed
to the shifting of absorption edge to a lower energy,
thus improving the photocatalytic activity in the
visible region.
Ren et al. (2007) observed a high surface area and
significant absorption in the visible light (>420 nm)
for a carbon doped TiO2 compared with undoped

counterpart. The doped TiO2 was reported to show


higher photocatalytic activity for the degradation of
RhB compared with undoped TiO2. In comparison to
undoped TiO2, 0.5% AgTiO2 was shown to achieve
1.53 times higher degradation of 2,4,6-trichlorophenol
by Rengaraj and Li (2006). Nonetheless, Ag content
beyond the stated level was reported to have a
detrimental effect on the photocatalytic activity of
TiO2. Under the conditions tested, Silva et al. (2009)
noticed an optimum Ce content of 0.6% (w/w) into
TiO 2 matrix for enhanced degradation of 4chlorophenol in the visible light. In addition to the
retardation of phase transformation, this behavior was
related to the shift of TiO2 absorption edge towards
longer wavelengths, by reducing the band gap of
original material. Subramanian et al. (2008) examined
the effect of Co doping on the structural and optical
properties of TiO2 to obtain a nanocrystalline films of
pure and doped TiO2 with grain size 10 to 25 nm. The
films were reported to be transparent in the visible
region, and the absorption edge shifted towards
the higher wavelength with increasing Co content.
The observed decrease in band gap was attributed to
the formation of Co impurity band into the TiO2
energy bands. Venkatachalam et al. (2007a) observed
that under identical and optimal experimental conditions, 1 mol% Mg2+ and Ba2+ doped nano-TiO2, nanoTiO2, and P25 required 300,360, and 450 min,
respectively, for the complete mineralization of
4-chlorophenol. The photocatalytic degradation of 4-

Water Air Soil Pollut

CP over TiO2 and Mg2+ and Ba2+ doped TiO2


indicated the higher activity for doped TiO2. Enhanced
adsorption of 4-CP on the catalyst surface and smaller
particle size due to Mg2+ and Ba2+ loadings are
indicated to be the cause for higher activity of the
catalysts. The BET surface area of both S-doped TiO2
and undoped TiO2 is shown to decrease significantly as
the calcinations temperature increases from 300C to
500C (Raileanu et al. 2009). The presence of S dopant
onto TiO2 matrix slightly improve their photocatalytic
activities and led to better chlorobenzene removal
compared with undoped TiO2. The photocatalytic
activity is also increased slightly as the sintering
temperature increases. Regardless of the organic
chloride pollutant, the removal has been reported to
vary between 96% and 100% under the conditions
studied. Lattice defects induced by dopant were
indicated to strongly influence the photocatalytic
activity. Irrespective of the number of coatings and
the irradiation times, the Ag and S-doped TiO2
exhibited higher removal compared with undoped
TiO2. The S-doped TiO2 were shown to be more
efficient than the Ag doped-TiO2 which was attributed
to the presence of anatase crystalline phase of TiO2 and
to a more significant surface concentration of dopant.
Xie and Zhao (2008) indicated that increasing reactive
temperature from 40C to 140C was reported to
enhance the crystalline size from 3.8 to 6.0 nm. F-N
and S-N co-doping resulted in a new strong absorption
band in the visible range of 400620 nm, and with the
decrease in reactive temperature, the absorption edge
shifted toward higher wavelength. In the visible region
(420 nm), photocatalytic activity of doped-TiO2 for
the degradation of MO was reported to be three times
more effective than that of bare P25. Sheng et al.
(2008) noticed that N-Br-codoped TiO2 readily
photodegrade MB under visible light irradiation in
comparison to N-doped TiO2. This behavior was
attributed to its anatase crystalline framework, low
electronhole recombination rate, and high absorbance
in the visible light. In contrast to P25 TiO2, photocatalytic activity of N, C-doped TiO2 was four times
higher for the degradation of phenol under visible light
(>400 nm) irradiation (Gorska et al. 2009a). The
degradation rate of RB by La-doped-TiO2 was higher
compared with undoped TiO2 for all the calcination
temperature investigated (Peng et al. 2005). TiO2
doped with 4.6% K+ and calcined at 700C was
demonstrated to exhibit higher activity compared

with other samples when the doping level of K+ and


calcinations temperature are 0143.% and 400
1,000C, respectively (Chen et al. 2007).
Wong et al. (2008) studied the influence of
crystallinity and carbon content on the photocatalytic
degradation of MB under visible light by varying the
carbon content from 03.2%. Under the conditions
tested, the highest degradation was achieved at 1.25%
carbon content. Zhu et al. (2004) observed that TiO2
doped with 0.09% FeCl3 and 0.09% FeCl2 exhibited
higher photocatalytic activity than undoped TiO2 for
the decolorization of Active Yellow XRG dye diluted
in water under both UV and visible light irradiation.
Fe2+/Fe3+ doped TiO2 was reported to have high
specific surface areas and small crystal sizes. TiO2
doped with FeCl3 was found to have better catalytic
activity for the degradation of XRG than those doped
with FeCl2. Yi et al. (2007) studied the effect of
Eu2+/Gd3+-co-doping onto TiO2 for the degradation of
MO under natural sunlight. The suitable doping
content is reported to be 0.51.0% for Eu2+ and
0.11.0% for Gd3+; the optimum doping content for
Eu2+ is 1.0% and for Gd3+ 0.5%. Coleman et al.
(2005) noticed that the addition of Ag or Pt had no
effect on the photocatalytic degradation of endocrine
disrupting chemicals at levels found in water. This
was attributed to the high concentration of holes and
hydroxyl radicals in the systems compared with the
low level of organic matters to be degraded. At high
concentration, a significant increase in the reaction
rate observed for bisphenol A and resorcinol over Pt/
TiO2. Under UV and visible light irradiation, the
mesoporous PdTiO2 film with a molar ratio of Pd/Ti=
0.05 exhibited the highest activity for the degradation
of MB due to smaller crystallite size and lower
crystallinity in mesoporous TiO2 film that that of
nonporous TiO2 film (Chan et al. 2009). The photocatalytic activity for aqueous phenol degradation of
anatase titania is shown to be enhanced upon doping
with Fe3+ up to ca. 1 wt.% (Adan et al. 2007). In
particular, maximum activity is observed for the
0.7 wt.% Fe3+ sample; in contrast, doping above ca.
3 wt.% does not produce any significant enhancement
in the photocatalytic activity. Under solar radiation,
La3+-TiO2 containing 0.4 mol % lanthanum showed
highest surface area(124.8 m2/g), lowest crystallite
size (8 nm), and exhibited 50% more degradation of
Methylene Blue compared with pure TiO2(Parida and
Sahu 2008). Zheng et al. (2007) noticed that the

Water Air Soil Pollut

coupled ZrO2ZnO system with 2.5% Zr was more


efficient compared with ZnO for the degradation of
reactive Brilliant Red X-3B. Wen et al. (2009)
investigated the effects of I and F codoped TiO2
(prepared by solgel-impregnation method) on the
photocatalytic degradation of MB. Co doping with I
and F was shown to enhance the degradation of MB
under simulated sunlight irradiation. The photocatalytic activity of as prepared I-F codoped TiO2 was
reported to be much higher than that of pure, I-doped,
and F-doped TiO2 when the molar ratios of I and F to
Ti were maintained at 10. The difference in the
photocatalytic activity was ascribed to larger surface
area and stronger absorbance in the visible light range
by doping with I and F. The photocatalytic activity of
co-doped TiO2 was higher than pure TiO2 under UV
and visible lights.
Bouras et al. (2007) compared the photocatalytic
degradation of Basic Blue in the presence of pure and
three cationic doped (Fe3+, Cr3+, and Co2+) TiO2
under UV-vis and visible light irradiation. In all three
studied cases, the efficiency of UV-vis photodegradation dramatically decreased in the presence of
dopants. The decrease was faster in the case of Fe3+
and Co2+ dopants and slower in the case of Cr3+. Li et
al. (2003) indicated that the photocatalytic activity of
Carbon-Black-modified nano-TiO2 (CBTiO2) thin
films was 1.5 times higher than that of TiO2 thin
films in degrading Reactive Brilliant Red X-3B. Tian
et al. (2009) studied the photocatalytic activity of V
doped-TiO2 for the degradation of MB and 2,
4-dichlorophenol using UV and visible light. In the
presence of UV light, 0.5% V-TiO2 was reported to be
efficient for the degradation of MB relative to pure
TiO2. Under visible light irradiation, the degradation
rate of 2,4-dichlorophenol over 1% V-TiO2 is shown
to be two times higher compared with undoped TiO2.
Baiju et al. (2007) tested the photocatalytic activity of
Ta2O5 doped TiO2 for the degradation of MB by
varying the amount of Ta2O5 from 1 to 10 mol %.
Under the conditions examined, 2 mol % of tantalum
was shown to be efficient compared with pure TiO2.
Zhou et al. (2009) studied the photocatalytic activity
of Nd-doped ZnO for the degradation of Reactive
Blue 4 in aqueous suspension by varying the dopant
concentration from 13 mol%. Some 2.5 mol %
Nd-doped ZnO was demonstrated to be efficient for
the enhanced photocatalytic activity under the experimental conditions. Naeem and Ouyang (2009)

studied the effect of Fe3+ doping (03.0 mol% Fe3+)


on the photocatalytic activity of TiO 2 for the
degradation of phenol under UV light. Fe3+-doped
TiO2 was reported to possess the anatase structure
with a range of crystal size 811 nm. The highest
degradation efficiency was found at 0.5 mol% Fe3+doped TiO2. After 2-h calcination at 600C, the order
of photocatalytic activity of La-doped TiO2 for phenol
degradation was demonstrated to be 1>1.5>3>0.5>
5>0 mol% La (Liqiang et al. 2004). Mohammad and
Al-Esaimi (2006) observed that 98% degradation of
MB was achieved after 75 min UV light irradiation in
the presence of 2V/TiO2SO4 catalyst, whereas only
78% degradation was obtained using 2V/TiO2 in the
same time. The variation in degradation efficiency
was related to the difference in vanadium and sulfate
content. Under visible light irradiation, the photocatalytic activity of S-doped TiO2ZrO2 was shown to
be higher compared with unmodified STiO2 and
Degussa P25 for the degradation of Rhodamine B
(Tian et al. 2009). The observed effect has been
attributed to the larger surface area, smaller crystal
size, porous structure, and more surface hydroxyl
groups in the catalyst. Ghasemia et al. (2009)
tested the effect of transition metal ions on the
photocatalytic degradation of Acid Blue 92 under
UV light. The effect of transition metals ions on the
photocatalytic degradation rate and efficiency was
shown to be Fe-TiO 2 >CoTiO 2 >CrTiO 2 >Mn
TiO2>CuTiO2>NiTiO2>ZnTiO2>TiO2. The photocatalytic degradation of MB by various TiO2/SnO2
thin films photocatalyst was investigated by varying
the amount of tin. For all TiO2/SnO2 thin films, the
degradation efficiency was higher compared with
pure TiO2 films, exhibiting the highest for TiO2/
SnO2 films with 9% Sn. The behavior was related to
greater porosity and smaller band gap energies
compared with pure TiO2 films (Martinez et al. 2005).
Yuan et al. (2002) studied the influence of
co-doping of Zn2+ and Fe3+ on the photocatalytic
degradation of phenol under solar light irradiation.
The co-doping of 0.5 mol% Zn2+ and 1 mol % Fe3+
onto TiO2 was found to be two times more efficient
compared with pure TiO2. This behavior was attributed to the coupled influenced of the co-dopant and
titania energy bands. Qu et al. (2009) studied the
effects of the co-addition of Zn2+ and sodium
dodecylbenzenesulfonate (DBS) on photocatalytic
degradation of RB relative to the undoped anatase

Water Air Soil Pollut

TiO2 nanoparticle films. In comparison to undoped


TiO2 film, the addition of Zn2+ was shown to improve
both the photocatalytic activity and the hydrophilicity,
which was attributed to surface oxygen vacancies.
However, the co-addition of Zn2+ and DBS resulted in
a film with super hydrophilic behavior. The degradation efficiency was reported to decrease from 43.5%
to 35% due to the co-addition of Zn2+ and DBS. Lee
et al. (2005) compared the photocatalytic degradation
of p-nitrophenol over TiO2 and TiO2/SiO2 nanoparticles prepared by the micro-emulsion method
using PFPE-NH4 surfactant. TiO 2 /SiO 2 (80:20)
nanoparticles were reported to have a higher photocatalytic activity than pure TiO2 and the TiO2/SiO2
(90:10) particles. This effect has been attributed to the
decrease of crystallite size from 17 to 133 nm with an
increase in silica content. Chen et al. (2009) studied
the effect of Cu2+ doping (00.2 wt.%) on the
photocatalytic degradation of Methyl Orange using
TiO2/SiO2. The optimum dopant content was about
0.10 wt.%. The photocatalytic activity of Cu2+-doped
TiO2/SiO2 system was higher than that of TiO2/SiO2.
This behavior was attributed to the trapping of both
photogenerated holes and electrons, and inhibition of
the electronhole recombination, leading to the
increase of photocatalytic activity. Arabatzis et al.
(2003) observed a two times faster degradation of
Methyl Orange in the presence of Au/TiO2 compared
with that obtained with the original TiO2 material.
Sonawane and Dongare (2006) tested the effect of
Au doping (12% Au/TiO2) on the photocatalytic
degradation of phenol in solar light irradiation of
4.55 Wm2/day. In comparison to undoped TiO2,
TiO2 doped with 12% Au showed 22.3 times
higher photocatalytic activity. Using solgel method,
Venkatachalam et al. (2007b) examined the effect of
Zr4+ doping(0.55.0 mol % Zr4+) onto TiO2 matrix
for the photocatalytic degradation of 4-chlorophenol.
The photocatalytic activity of 3.0 mol% doped TiO2
samples were reported to be higher than that of nanoTiO2 and Degussa P25. Enhanced adsorption of 4CP over the catalyst surface and decrease in particle
size are indicated to be the reason for high activity of
the catalyst. Bellardita et al. (2007) investigated the
effect of metal loading (W, Co, and Sm) onto TiO2
matrix on the photocatalytic degradation of 4nitrophenol. In comparison to Degussa P25, TiO2
loaded with 1% W and Sm caused a significant
improvement in the photocatalytic activity under UV

irradiation whereas Co was beneficial only under


visible light irradiation. This effect was ascribed to
an increased charge separation of the photogenerated
electronhole pairs.

9 Conclusion
Based on the recent representative studies, this review
focuses on the role of various operating parameters on
the photocatalytic degradation of various dyes and
phenols, and reports the main advances.TiO2 has been
suggested to be efficient for the degradation and
mineralization of various toxic organic pollutants, e.
g., phenols and dyes in water in the presence of UV,
visible or solar light, and oxygen. The findings also
suggest that various operating parameters such as type
of photocatalyst, light intensity, pollutant types, and
initial concentration, amount of catalyst, initial pH of
the reaction medium, mode of catalyst application,
oxidizing agents/electron acceptors, and presence of
ionic components in solution can significantly influence the photocatalytic degradation rate of phenols
and dyes. Optimization of degradation parameters is
crucial from the perspective of efficient design and
application of photocatalytic oxidation process to
ensure sustainable operation. The potential of this
technique under multi-component pollutants needs
further attention to yield stable pollutant removal
through the optimization of process parameters. Metal
and non-metal doped TiO2 have been reported to be
efficient for the improved degradation rate. In spite of
extensive investigations, the commercial exploitation
of photocatalysis has been hindered by the lack of
efficient and low-cost visible light harvesting catalyst,
a relatively poor understanding of the reactor design,
and inadequate scale-up strategies. Future research
should focus on the development of a more reliable
photocatalyst that can be activated by visible and
solar light or both. In addition, more work is required
on the modeling of photoreactor to optimize its design
for pollutant degradation. In the literature, there is
currently little information on the modeling of photoreactor to optimize its performance. Present research
activities at CQ University and QUT, Australia focus
on the computational fluid dynamics modeling of a
flat plate reactor to optimize its design and to predict
its performance. Although this review is not exhaustive in the scope of photocatalytic degradation of

Water Air Soil Pollut

organic pollutants; however, it addresses the fundamental principles and recent applications in this area.

Acknowledgement This study is supported under an Australian Research Council (ARC) linkage grant in collaboration
with CM Concrete Private limited and Department of Public
Works, QLD Government. The authors gratefully acknowledge
the financial support of ARC project. One author is also
grateful for the financial support of the Queensland Government through the Smart State fellowship scheme.

References
Aarthi, T., & Madras, G. (2007). Photocatalytic degradation of
Rhodamine yes with nano-TiO2. Industrial & Engineering
Chemistry Research, 46, 714.
Aarthi, T., Narahari, P., & Madras, G. (2007). Photocatalytic
degradation of Azure and Sudan dyes using nano TiO2.
Journal of Hazardous Materials, 149, 725734.
Abdullah, M., Low, G., & Mathews, R. W. (1990). Effects of
common inorganic ions on rates of photocatalytic oxidation of organic carbon over illuminated titanium dioxide.
Journal of Physical Chemistry, 94, 6820.
Adan, C., Bahamonde, A., Fernandez-Garica, M., & MartinezArias, A. (2007). Structure and activity of nanosized irondoped anatase TiO2 catalysts for phenol photocatalytic
degradation. Applied catalysis B: Environmental, 72, 1117.
Adesina. (2004). Industrial exploitation of photocatalysis:
progress, perspectives and prospects. Catalysis Surveys
from Asia, 8(4), 265273.
Alaton, I. A., & Balcioghu, I. A. (2001). Photochemical and
heterogeneous photocatalytic degradation of waste vinylsulphone dyes: A cases study with hydrolysed Reactive
black 5. Journal of Photochemistry and Photobiology A:
Chemistry, 141, 247254.
Alinsafi, A., Evenou, F., Abdulkarim, E. M., Pons, M. N.,
Zahraa, O., Benhammou, A., et al. (2007). Treatment of
textile industry wastewater by supported photocatalysis.
Dyes & Pigments, 74, 439445.
Anandan, S., Satish Kumar, P., Pugazhenthiran, N., Mdhavan, J.,
& Maruthamuthu, P. (2008). Effect of loaded silver nanoparticles on TiO2 for photocatalytic degradation of Acid Red
88. Solar Energy Materials & Solar Cell, 92, 929937.
Andronic, L., & Duta, A. (2007). TiO2 thin films for dyes
photodegradation. Thin Solid Films, 515, 62946297.
Arabatzis, I. M., Stergiopoulos, T., Andreeva, D., Kitova, S.,
Neophytides, S. G., & Falaras, P. (2003). Characterization
and photocatalytic activity of Au/TiO2 thin films for azodye degradation. Journal of Catalysis, 220, 127135.
Arques, A., Amat, A. M., Garcia-Ripoll, & Vicente, R. (2007).
Detoxification and/or increase of the biodegradability of
aqueous solutions of dimethoate by means of solar photocatalysis. Journal of Hazardous Materials, 146, 447452.
Bahnemann, W., Muneer, M., & Haque, M. M. (2007).
Titanium dioxide-mediated photocatalysed degradation of
few selected organic pollutants in aqueous suspensions.
Catalysis Today, 124, 133148.

Baiju, K. V., Shajesh, P., Wunderlich, W., Mukundan, P.,


Kumar, S. R., & Warrier, K. G. K. (2007). Effect of
tantalum addition on anatase phase stability and photoactivity of aqueous solgel derived mesoporous titania.
Journal of Molecular Catalysis A: Chemical, 276, 4146.
Behnajady, M. A., Modirshahla, N., Daneshvar, N., & Rabbani,
M. (2007). Photocatalytic degradation of an azo dye in
atubular continuous-flow photoractor with immobilized
TiO2 on glass plates. Chemical Engineering Journal, 127,
167176.
Behnajady, M. A., Moghaddam, S. G., Modirshahla, N., &
Shokri, M. (2009). Investigation of the effect of heat
attachment method parameters at photocatalytic activity of
immobilized ZnO nanoparticles on glass plate. Desalination. doi:10.1016/j.desal.2009.06.021.
Bejarano-Perez, N. J., & Suarez-Herrera, M. F. (2007).
Sonophotocatalytic degradation of Congo Red and Methyl
Orange in the presence of TiO2 as a catalyst. Ultrasonics
Sonochemistry, 14, 589595.
Bellardita, M., Addamo, M., Paola, A. D., & Palmisano, L. (2007).
Photocatalytic behaviour of metal-loaded TiO2 aqueous
dispersions and films. Chemicla Physiscs, 339, 94103.
Bessekhouad, Y., Robert, D., Weber, J.-V., & Chaoui, N.
(2004). Effect of alkaline-doped TiO2 on photocatalytic
efficiency. Journal of Photochemistry and Photobiology
A: Chemistry, 167, 4957.
Bouras, P., Stathatos, E., & Lianos, P. (2007). Pure versus
metal-ion-doped nanocrystalline titania for photocatalysis.
Applied Catalysis B: Environment, 73(12), 5159.
Byrappa, K., Subramania, A. K., Ananda, S., Rai, K. M. L.,
Dinesh, R., & Yushimura, M. (2006). Photocatalytic
degradation of Rodamine B dye using hydrothermally
synthesized ZnO. Bulletin of Materials Science, 29(5),
433438.
Calza, P., & Pelizzetti, E. (2001). Photocatalytic transformation
of organic compounds in the presence of inorganic ions.
Pure Applied Chemistry, 73(12), 18391848.
Cassano, A. E., & Alfano, O. M. (2000). Reaction engineering
of suspended solid heterogenous photocatalytic reactors.
Catalysis Today, 58, 167197.
Chan, C.-C., Chang, C.-C., Hsu, W.-C., Wang, S.-K., & Lin, J.
(2009). Photocatalytic activities of Pd-loaded mesoporous
TiO2 thin films. Chemical Engineering Journal, 152, 492
497.
Chen, Y.-C., & Smirniotis, P. (2002). Enhancement of photocatalytic degradation of phenol and chlorophenols by
ultrasound. Industrial & Engineering Chemistry Research,
41(24), 59585965.
Chen, L.-C., Huang, C.-M., & Tsai, F.-R. (2007). Characterization and photocatalytic activity of K+-doped TiO2
photocatalysts. Journal of Molecular Catalysis A: Chemical, 265, 133140.
Chen, R.-F., Zhang, C.-X., Deng, J., & Song, G.-Q. (2009).
Preparation and photocatalytic activity of Cu2+-doped
TiO2/SiO2. International Journal of Minerals, Metallurgy
and Materials, 16(2), 220225.
Chiou, C.-H., & Juang, R.-H. (2007). Photocatalytic degradation of phenol in aqueous solutions by Pr-doped TiO2
nanoparticles. Journal of Hazardous Materials, 149, 17.
Chiou, C.-H., Wu, C.-Y., & Juang, R.-S. (2008a). Influence of
operating parameters on photocatalytic degradation of

Water Air Soil Pollut


phenol in UV/TiO2 process. Chemical Engineering Journal,
139, 322329.
Chiou, C.-H., Wu, C.-Y., & Juang, R.-S. (2008b). Photocatalytic degradation of phenol and m-nitrophenol using
irradiated TiO2 in aqueous solutions. Separation and
Purification Technology, 62, 559564.
Coleman, H. M., Chiang, K., & Amal, R. (2005). Effects of Ag
and Pt on photocatalytic degradation of endocrine disrupting chemicals in water. Chemical Engineering Journal,
113, 6572.
Cun, W., Zhao, J., Xinming, W., Bixian, M., Guoying, S., An,
P. P., et al. (2002). Preparation, characterization and
photocatalytic activity of nano-sized ZnO/SnO2 coupled
photocatalysis. Applied Catalysis B; Environmental, 39,
269279.
Daneshvar, N., Salari, D., & Khataee, A. R. (2003). Photocatalytic
degradation of azo dye Acid Red 14 in water: Investigation of
the effect of operational parameters. Journal of Photochemistry and Photobiology A, 157, 111116.
Daneshvar, N., Salari, D., & Khataee, A. R. (2004). Photocatalytic degradation of azo dye acid red 14 in water on
ZnO as an alternative catalyst to TiO2. Journal of
Photochemistry and Photobiology A, 162, 317322.
Daneshvar, N., Salari, D., Niaei, A., & Khataee, A. R. (2006).
Photocatalytic degradation of the herbicide erioglaucine in
the presence of nanosized titanium dioxide: comparison
and modeling of reaction kinetics. Journal of Environmental Science and Health B, 41, 12731290.
Daneshvar, N., Rasoulifard, M. H., Khataee, A. R., & Hosseinzadeh, F. (2007). Removal of C.I. Acid Orange 7 from aqueous
solution by UV irradiation in the presence of ZnO nano
powder. Journal of Hazardous Materials, 143, 95101.
DEC (Department of Environment Conservation). (2006).
Managing urban stormwater: harvesting and reuses,
NSW, DEC2006/137
DEH (Department of Environment and Heritage). (2002).
Introduction to urban stormwater management in Australia, prepared under the stormwater initiative of the
living cities PROGRAM 2002
Doll, T. E., & Frimmel, F. H. (2005). Photocatalytic degradation of carbamazepine, clofibric acid and iomeprol with
P25 and Hombikat UV100 in the presence of natural
organic matter (NOM) and other organic water constituents. Water Research, 39, 403411.
Echavia, R. G. M., Matzusawa, F., & Negishi, N. (2009).
Photocatalytic degradation of organophosphate and phosphonoglycine pesticides using TiO2 immobilized on silica
gel. Chemosphere, 76, 595600.
Epling, G. A., & Lin, C. (2002). Investigation of retardation
effects on the titanium dioxide photo degradation system.
Chemosphere, 46, 937944.
Eriksson, E., Baun, A., Mikkelsen, P. S., & Ledin, A. (2007).
Risk assessment of xenobiotics in stormwater discharged
to Harrestup Ao, Denmark. Desalination, 215(2007), 187
197.
Faisal, M., Tariq, M. A., & Muneer, M. (2007). Photocatalysed
degradation of two selected dyes in UV-irradiated aqueous
suspensions of titania. Dyes and Pigments, 72, 233239.
Fernando, J., Beltrn, F., Rivas, J., & Gimeno, O. (2005).
Comparison between photocatalytic ozonation and other
oxidation processes for the removal of phenols from water.

Journal of Chemical Technology and Biotechnology, 80,


973984.
Galindo, C., Jaques, P., & Kalt, A. (2000). Photodegradation of the
amino-azobenzene acid orange 52 by three advanced oxidation processes: UV/H2O2, UV/TiO2 and Vis/TiO2 comparative mechanistic and kinetic investigations. Journal of
Photochemistry and Photobiology A: Chemistry, 130, 3547.
Garcia, A., Amat, A. M., Arques, A., Sanchis, R., Gernjak, W.,
Maldonado, M. I., et al. (2006). Detoxification of aqueous
solution of herbicide Sevnol by solar photocatalysis.
Environmental Chemistry Letters, 3, 169172.
Garcia, A., Arques, A., Vicente, R., Domenech, A., & Amat, A.
M. (2008). Treatment of aqueous solutions containing four
commercial pesticides by means of TiO2 solar photocatalysis. Journal of Solar Energy Engineering, 130,
041011041015.
Garca, A., Amat, A. M., Arques, A., Vicente, R., Lpez, M.
F., & Oller, I. (2007). Increased biodegradability of
Ultracide in aqueous solutions with solar TiO2 photocatalysis. Chemosphere, 68, 293300.
Ghasemia, S., Rahimnejada, S. R., Setayesha, S. R., &
Gholamia, M. R. (2009). Transition metal ions effect on
the properties and photocatalytic activity of nanocrystalline TiO2 prepared in an ionic liquid. Journal of
Hazardous Materials. doi:10.1016/j.jhazmat.2009.08.029.
Gonzalez, A. S., & Martinez, S. S. (2008). Study of the
photocatalytic degradation of Basic Blue 9 in industrial
textile dye over slurry titanium dioxide and influencing
factors. Ultrasonic Sonochemistry, 15, 386392.
Gorska, P., Zaleska, A., & Hupka, J. (2009a). Photodegradation
of phenol by UV/TiO2 and Vis/N, C-TiO2 process:
Comparative mechanistic and Kinetic Studies. Separation
and Purification Technology, 68, 9096.
Gorska, P., Zaleska, A., Kowalska, E., Klimczuk, T., Sobczak,
J. W., Skwarek, E., et al. (2009b). TiO2 photoactivity in
vis and UV light: The influence of calcinations temperature and surface properties. Applied catalysis B: Environmental, 84, 440447.
Goslich, R., Dillert, R., & Bahnemann, D. (1997). Solar water
treatment principles and reactors. Water Science and
Technology, 35(4), 137148.
Goswami, D. Y. (1997). A review of engineering developments
of aqueous phase solar photocatalytic detoxification and
disinfection processes. Journal of Solar Energy Engineering, 119, 101107.
Grzechulska-Damszel, J., Tomaszewska, M., & Morawski, A.
W. (2009). Integration of photocatalysis with membrane
processes purification of water contaminated with organic
dyes. Desalination, 241, 118126.
Guetta, N., & Ait Amar, H. (2005). Photocatalytic oxidation
of methyl orange in presence of titanium dioxide in
aqueous suspension. Part II: kinetics study. Desalination,
185, 439448.
Guillard, C., Disdier, J., Herrmann, J. M., Lehaut, C., Chopin,
T., Malato, S., et al. (1999). Comparison of various titania
samples of industrial origin in the solar photocatalytic
detoxification of water containing 4-chlorophenol. Catalysis Today, 54, 217228.
Guillard, C., Lachheb, H., Houas, A., Ksibi, M., Elaloui, E., &
Herrman, J.-M. (2003). Influence of chemical structure of
dyes of pH and of inorganic salts on their photocatalytic

Water Air Soil Pollut


degradation by TiO2 comparison of efficiency of powder
and supported TiO2. Journal of Photochemistry and
Photobiology A: Chemistry, 158, 2736.
Gupta, A. K., Pal, A., & Sahoo, C. (2006). Photocatalytic
degradation of a mixture of Crystal Violet (Basic Violet 3)
and Methyl Red dye in aqueous suspensions using Ag-C
doped TiO2. Dyes and Pigments, 69, 224232.
Haque, M., & Muneer, M. (2007). TiO2-mediated photocatalytic degradation of a textile dyederivative, Bromothymol Blue, in aqueous suspensions. Dyes and Pigments,
75, 443448.
Hermann, J. M. (1999). Heterogeneous photocatalysis: Fundamentals and applications to the removal of various types of
aqueous pollutants. Catalysis Today, 53, 115129.
Hincapie, M., Peuela, P. G., Maldonado, M. I., Malato, O.,
Fernndez-Ibez, P., & Oller, I. (2006). Degradation of
pesticides in water using solar advanced oxidation processes. Applied Catalysis B: Environment, 64, 272281.
Huang, M., Xu, C., Wu, Z., Huang, Y., Lin, J., & Wu, J. (2008).
Photocatalytic discolorization of Methyl Orange solution
by Pt modified TiO2 loaded on natural zeolite. Dyes and
Pigments, 77, 327334.
Kaneco, S., Rahman, M. A., Suzuki, T., Katsumata, H., &
Ohta, K. (2004). Optimization of solar photocatalytic
degradation conditions of bisphenol A in water using
titanium dioxide. Journal of photochemistry and photobiology A: Chemistry, 163, 419424.
Kato, S., Hirano, Y., Iwata, M., Sano, T., Takeuchi, K., &
Matsuzawz, S. (2005). Photocatalytic degradation of gaseous
sulphur compounds by silver-deposited titanium dioxide.
Applied Catalysis B: Environmental, 57, 109115.
Khataee, A. R., Pons, M. N., & Zahraa, O. (2009). Photocatalytic degradation of three azo dyes using immobilized
TiO2 nanoparticles on glass plates activated by UV light
irradiation: Influence of dye molecular structure. Journal
of Hazardous Materials, 168, 451457.
Kirilov, M., Koumanova, L., Spasov, L., & Petrov. (2006).
Effects of Ag and Pd modifications of TiO2 on the
photocatalytic degradation of p-chlorophenol in aqueous
solution. Journal of the University of Chemical Technology and Metallurgy, 41(3), 343348.
Konstantinou, I. K., & Albanis, T. A. (2003). Photocatalytic
transformation of pesticides in aqueous titanium dioxide
suspensions using artificial and solar light: Intermediates
and degradation pathways. Applied Catalysis B: Environmental, 42, 319335.
Kositzi, M., Poulios, I., Samara, K., Tsatsaroni, E., & Darakas,
E. (2007). Photocatalytic oxidation of Cibacron Yellow
LS-R. Journal of Hazardous Materials, 146, 680685.
Lachheb, H., Houas, A., & Herrmann, J.-M. (2008). Photocatalytic degradation of polynitrophenols on various
commercial suspended or deposited tiatania catalyst using
artificial and solar light. International Journal of Photoenergy, Article ID 497895. doi:1155/2008/497895
Lee, M. S., Lee, G.-D., Park, S. S., Ju, C.-S., Lim, K. T., &
Hong, S.-S. (2005). Synthesis of TiO2/SiO2 nanoparticles
in a water-in-carbon-dioxide microemulsion and their
photocatalytic activity. Research of Chemical Intermediates, 31(4), 379389.
Lee, S., Yun, C. Y., Hahn, M. S., Lee, J., & Yi, J. (2008).
Synthesis and characterization of carbon-doped titania as a

visible-light-sensitive photocatalyst. Korean Journal of


Chemical Engineering, 25(4), 892896.
Legrini, O., Oliveros, E., & Braun, A. M. (1993). Photochemical for water treatment. Chemical Reviews, 93, 671698.
Lettmann, C., Hildenbrand, K., Kisch, H., Macyk, W., & Maier,
W. F. (2001). Visible light photodegradation of 4chlorophenol with a coke-containing titanium dioxide
photocatalyst. Applied Catalysis B: Environmental, 32,
215227.
Li, L., Zhu, W., Zhang, P., Chen, Z., & Han, W. (2003).
Photocatalytic oxidation and ozonation of catechol over
Carbon-Black-modified nano-TiO2 thin films supported on
Al sheet. Water Research, 37, 36463651.
Lim, S. H., Ferrarisa, C., Schreyera, M., Shihc, K., Leckiec, J.,
& White, T. J. (2007). The influence of cobalt doping on
photocatalytic nano-titania: Crystal chemistry and amorphicity. Journal of Solid State Chemistry, 180, 29052915.
Lin, C., & Lin, K. (2007). Photocatalytic oxidation of toxic
organohalides with TiO2/UV: The effects of humic substances and organic mixtures. Chemosphere, 66, 1872
1877.
Liqiang, J., Xiaojun, S., Baifu, X., Baiqi, W., Weimin, C., &
Hongganga, F. (2004). The preparation and characterization of La doped TiO2 nanoparticles and their photocatalytic activity. Journal of Solid State Chemistry, 177,
33753382.
Liu, Y., Chen, X., Li, J., & Burda, C. (2005). Photocatalytic
degradation of azo dyes by nitrogen-doped TiO2 nano
catalysts. Chemosphere, 61, 1118.
Liu, J., Qin, W., Zuo, S., Yu, Y., & Hao, Z. (2009).
Solvothermal-induced phase transition and visible photocatalytic activity of nitrogen-doped titania. Journal of
Hazardous Materials, 163, 273278.
Lu, A., Li, Y., Lv, M., Wang, C., Yang, L., Liu, J., et al. (2007).
Photocatalytic oxidation of methyl orange by natural Vbearing rutile under visible light. Solar Energy Materials
& Solar Cells, 91, 8491855.
Mahmoodi, N. M., & Arami, M. (2009). Degradation and
toxicity reduction of textile wastewater using immobilized
titania nanophotocatalysis. Journal of Photochemistry and
Photobiology B: Biology, 94, 2024.
Mahmoodi, N. M., Armani, M., Lymaee, N. Y., & Gharanjig,
K. (2007). Photocatalytic degradation of agricultural Nheterocyclic organic pollutants using immobilized nanoparticles of titania. Journal of Hazardous Materials, 145
(1-2), 6571.
Mahvi, A. H., Ghanbarian, M., Nasseri, S., & Khairi, A.
(2009). Mineralization and discoloration of textile wastewater by TiO2 nanoparticles. Desalination, 239(2009),
309316.
Malato, S., Blanco, J., Cceres, J., Fernndez-Alba, A. R.,
Agera, A., & Rodiguez, A. (2002). Photocatalytic
treatment of water-soluble pesticides by photo-Fenton
and TiO2 using solar energy. Catalysis Today, 76, 209
220.
Maldonado, M. I., Passarinho, P. C., Oller, I., Gernjak, W.,
Fernndez, P., & Blanco, J. (2007). Photocatalytic degradation of EU priority substances: a comparison between
TiO2 and Fenton plus photo-Fenton in a solar pilot plant.
Journal of Photochemistry and Photobiology A: Chemistry, 185, 354363.

Water Air Soil Pollut


Martinez, A. I., Acosta, D. R., & Cedillo, G. (2005). Effect of
SnO2 on the photocatalytic properties of TiO2 films. Thin
Solid Films, 490, 118123.
Mathews, R. W. (1986). Photooxidation of organic material in
aqueous suspensions of titanium dioxide. Water Research,
20(5), 569578.
McMurray, T. A., Dunlop, P. S. M., & Byrne, J. A. (2006). The
photocatalytic degradation of atrazine on nanoparticulate
TiO2 films. Journal of Photochemistry and Photobiology
A: Chemistry, 182, 4351.
Mitchell, V. G., Mein, R. G., & Mcmahon, T. A. (2002).
Utilising storm water and wastewater resources in urban
areas. Australian Journal of Water Resources, 6, 3143.
Moctezumaa, E., Leyva, E., Palestino, G., & Lasa, H. D. (2007).
Photocatalytic degradation of methyl parathion: Reaction
pathways and intermediate reaction products. Journal of
Photochemistry and Photobiology A: Chemistry, 186, 7184.
Mohammad, M. M., & Al-Esaimi, M. M. (2006). Characterization, adsorption and photocatalytic activity of V-doped
TiO2 and sulphated TiO2 (rutile) catalysts: Degradation of
Methylene Blue dye. Journal of Molecular Catalysis:
Chemical, 255, 5361.
Mozia, S., Morawski, A. W., Toyoda, M., & Tsumura, T.
(2009). Integration of photocatalysis and membrane
distillation for removal of mono- and poly-azo dyes from
water. Desalination. doi:10.1016/j.desal.2009.06.075.
Mrowetz, Pirola, M. C., & Selli, E. (2005). Degradation of
organic water pollutants through sonophotocatalysis in the
presence of TiO2. Ultrasonics Sonochemistry, 10, 247.
Mukherjee, P. S., & Ray, A. K. (1999). Major challenges in the
design of a large scale photocatalytic reactor for water
treatment. Chemical Engineering Technology, 22, 253
260.
Muruganandham, M., & Swaminathan, M. (2004). Solar
photocatalytic degradation of a reactive Azo dye in TiO2suspension. Solar Energy Materials & Solar Cells, 81,
439457.
Muruganandham, M., & Swaminathan, M. (2006a). Photocatalytic decolourisation and degradation of Reactive
Orange 4 by TiO2-UV process. Dyes and Pigments, 68,
133142.
Muruganandham, M., & Swaminathan, M. (2006b). TiO2UV
photocatalytic oxidation of Reactive Yellow 14: Effect of
operational parameters. Journal of Hazardous Materials,
B135, 7886.
Muruganandham, M., Shobana, N., & Swaminathan, M.
(2006). Optimization of solar photocatalytic degradation
conditions of Reactive Yellow 14 azo dye in aqueous
TiO2. Journal of Molecular Catalysis A: Chemical, 246,
154161.
Naeem, K., & Feng, O. (2008). Parameters effect on heterogenous photocatalysed degradation of phenol in aqueous
dispersion of TiO2. Journal of Environmental Science, 21,
527533.
Naeem, K., & Ouyang, F. (2009). Preparation of Fe3+-doped
TiO2 nanoparticles and its photocatalytic activity. Physica
B: Condensed Matter, 405(1), 221226.
Nikazor, M., Gholivand, K., & MahanPoor, K. (2008). Photocatalytic degradation of azo dye Acid Red 114 in water
with TiO2 supported on clinoptilolite as a catalyst.
Desalination, 219, 293300.

Oller, I., Gernjak, W., Maldonado, M. I., Prez-Estrada, L. A.,


Snchez-Prez, J. A., & Malato, S. (2006). Solar photocatalytic degradation of some hazardous water-soluble
pesticides at pilot-plant scale. Journal of Hazardous
Materials, B138, 507517.
Ollis, D. F., Pelizzetti, E., & Serpone, N. (1991). Photocatalyzed destruction of water contaminants. Environmental Science and Technology, 25(9), 15221529.
Oyama, T., Yanagisawa, I., Takeuchi, M., Koike, T., Serpone,
N., & Hidaka, H. (2009). Remediation of simulated
aquatic sites contaminated with recalcitrant substrates by
TiO2/ozonation under natural sunlight. Applied Catalysis
B: Environmental, 91, 242246.
Papadam, T., Xekoukoulotakis, N. P., Poulios, I., & Mantzavinos, D. (2007). Photocatalytic transformation of acid
orange 20 and Cr(VI) in aqueous TiO2 suspensions.
Journal of Photochemistry and Photobiology A: Chemistry, 186, 308315.
Pare, B., Jonnalagadda, S. B., Tomar, H., Singh, P., & Bhagwat,
V. W. (2008). ZnO assisted photocatalytic degradation of
Acridine Orange aqueous solution using visible irradiation. Desalination, 232, 8090.
Pareek, V., Chong, S., Tade, M., & Adesina, A. (2008). Light
intensity distribution in heterogeneous photocatalytic reactors.
Asia-Pacific Journal of Chemical Engineering, 3, 171201.
Parida, K. M., & Sahu, N. (2008). Visible light induced
photocatalytic activity of rare earth titania nanocomposites. Journal of Molecular Catalysis A: Chemical, 287,
151158. 2008.
Parida, K. M., Dash, S. S., & Das, D. P. (2006). Physicochemical characterization and photocatalytic activity of
zinc oxide presented by various methods. Journal of
Colloid and Interface Science, 298, 787793.
Pecchi, G., Reyes, P., Sanhueza, P., & Villasensor, J. (2001).
Photocatalytic degradation of pentachlorophenol on TiO2
solgel catalysts. Chemosphere, 43, 141146.
Peng, T., Zhaoa, D., Song, H., & Yan, C. (2005). Preparation of
lanthana-doped titania nanoparticles with anatase mesoporous walls and high photocatalytic activity. Journal of
Molecular Catalysis A: Chemical, 238, 119126.
Porter, J. E., Li, Y., & Cahn, C. K. (1999). The effect of
calcination on the microstructural characteristics and
photoreactivity of Degussa P25 TiO2. Journal of Materials
Science, 34, 15231531.
Qamar, M., Saquib, M., & Muneer, M. (2004). Semicondcutormediated photocatalytic degradation of an azo dye, chrysoidine Y in aqueous suspensions. Desalination, 171, 185193.
Qamar, M., Saquib, M., & Muneer, M. (2005a). Titanium
dioxide mediated photocatalytic degradation of two selected azo dye derivatives, chrysoidine R and acid red 29
(chromotrope 2R), in aqueous suspensions. Desalination,
186, 255271.
Qamar, M., Saquib, M., & Muneer, M. (2005b). Photocatalytic
degradation of two selected dye derivatives, chromotrope
2B and Amido Black 10B, in aqueous suspensions of
titanium dioxide. Dyes and Pigments, 6, 19.
Qu, Y., Song, S., Jing, L., Luan, Y., & Fu, H. (2009). Effects of
the co-addition of Zn2+ and sodium dodecylbenzenesulfonate on photocatalytic activity and wetting performance of
anatase TiO 2 nanoparticle films. Thin Solid Films.
doi:10.1016/j.tsf.2009.09.006.

Water Air Soil Pollut


Radcliff, J. (2006). Future directions for water recycling in
Australia. Desalination, 187, 7787.
Raileanu, M., Crisan, M., Dragaon, N., Crisan, D., Galtayries,
A., Braileanu, A., et al. (2009). Solgel doped TiO2
nanomaterials: A comparative study. Journal of Sol-Gel
Science and Technology, 51, 315329.
Ren, W., Ai, Z., Jia, F., Zhang, L., Fan, X., & Zou, Z. (2007).
Low temperature preparation and visible light photocatalytic activity of mesoporous carbon-doped crystalline
TiO2. Applied Catalysis B: Environmental, 69, 138114.
Rengaraj, S., & Li, X. Z. (2006). Enhanced photocatalytic
activity of TiO2 by doping with Ag for degradation of 2,
4, 6-trichlorophenol in aqueous suspension. Journal of
Molecular Catalysis A: Chemical, 243, 6067.
Rincon, A.-G., & Pulgarin, C. (2004). Effect of pH, inorganic
ions, organic matter and H2O2 on E. coli K12 photocatalytic inactivation by TiO2. Applied Catalysis B:
Environment, 51, 283302.
Sahoo, C., Gupta, A. K., & Pal, A. (2005). Photocatalytic
degradation of Methyl Red dye in aqueous solution under UV
irradiation using Ag+ doped TiO2. Desalination, 181, 91100.
Saien, J., & Khezrianjoo, S. (2008). Degradation of the
fungicide carbendazim in aqueous solutions with UV/
TiO2 process: optimization, kinetics and toxicity studies.
Journal of Hazardous Materials, 157, 269276.
Saquib, M., & Muneer, M. (2002). Semiconductor mediated
photocatalysed degradation of an anthraquinone dye,
Remazol Brilliant Blue R under sunlight and artificial
light source. Dyes and Pigments, 53, 237249.
Saquib, M., & Muneer, M. (2003). Titanium dioxide mediated
photocatalyzed degradation of a textile dye derivative,
Acid Orange 8, in aqueous suspensions. Desalination,
155, 255263.
Saquib, M., Tariq, M. A., Faisal, M., & Muneer, M. (2008a).
Photocatalytic degradation of two selected dye derivatives
in aqueous suspensions of titanium dioxide. Desalination,
219, 301311.
Saquib, M., Tariq, M. A., Haque, M. M., & Muneer, M.
(2008b). Photocatalytic degradation of disperse blue 1
using UV/TiO2/H2O2 process. Journal of Environmental
Management, 88, 300306.
Selvam, K., Muruganandham, M., Muthuvel, I., & Swaminathan,
M. (2007). The influence of inorganic oxidants and metal
ions on semiconductor sensitized photodegradation of 4flurophenol. Chemical Engineering Journal, 128, 5157.
Shakthivel, S., Janczarek, M., & Kisch, H. (2004). Visible light
activity and photoelectrochemical properties of nitrogendoped TiO2. The Journal of Physical Chemistry B, 108
(50), 1938419387.
Sheng, Y., Xu, Y., Jiang, D., Liang, L., Wu, D., & Sun Y.
(2008). Hydrothermal preparation of visible-light-driven
N-Br-Codoped TiO2 photocatalysts. International Journal
of Photoenergy, Article ID 563949
Shifu, C., & Gengyu, C. (2005). Photocatalytic degradation of
pesticides using floating photocatalyst TiO2.SiO2 beads by
sunlight. Solar Energy, 79, 19.
Shifu, C., & Yunzhang, L. (2007). Study on the photocatalytic
degradation of glyphosate by TiO2 photocatalyst. Chemosphere, 67, 10101017.
Shohrabi, M. R., Davallo, M., & Miri, M. (2009). Influence of
operating parameters on eliminating Azo dyes from

wastewater by advanced oxidation technology. International Journal of ChemTech Research, 1(3), 446451.
Silva, A. M. T., Silva, C. G., Draic, G., & Faria, J. L. (2009).
Ce-doped TiO2 for photocatalytic degradation of chlorophenol. Catalysis Today, 144, 1318.
Silveyra, R., Torre Senz, L. D., Flores, W. A., Martinez, V. C.,
& Elguzabal, A. A. (2005). Doping of TiO2 with nitrogen
to modify the interval of photocatalytic activation towards
visible radiation. Catalysis Today, 107108, 602605.
Sobana, N., & Swaminathan, M. (2007). The effect of
operational parameters on the photocatalytic degradation
of acid red 18 by ZnO. Separation and Purification
Technology, 56, 101107.
Sonawane, R. S., & Dongare, M. K. (2006). Solgel synthesis
of Au/TiO2 thin films for photocatalytic degradation of
phenol in sunlight. Journal of Molecular catalysis A:
Chemical, 243, 6879.
Song, X.-M., Wu, J.-M., & Yan, M. (2009). Photocatalytic
degradation of selected dyes by titania thin films with
various nanostructures. Thin Solid Films, 517, 4341
4347.
Soutsas, K., Karayannis, V., Poulios, I., Riga, A., Ntampegliotis, K., & Spiliotis, X. (2009). Decolorization and
degradation of reactive azo dyes via heterogeneous photocatalytic processes. doi:10.1016/j.desal.2009.09.054
Su, Y., Deng, L., Zhang, N., Wang, X., & Zhu, X. (2008).
Photocatalytic degradation of C.I. Acid Blue 80 in
aqueous suspensions of titanium dioxide under sunlight.
Reaction Kinetics and Catalysis Letters. doi:10.1007/
s11144- 009-0059-4.
Subramanian, M., Vijayalakshmi, S., Venkataraj, S., & Jayavel,
R. (2008). Effect of cobalt doping on the structural and
optical properties of TiO2 films prepared by solgel
process. Thin Solid Films, 516, 37763782.
Sun, L., & Bolton, J. R. (1996). Determination of the quantum
yield for the photochemical generation of hydroxyl
radicals in TiO2 suspensions. Journal of Physical Chemistry, 100, 41274134.
Talebian, N., & Nilforoushan, M. R. (2009). Comparative study
of the structural, optical and photocatalytic properties of
semiconductor metal oxides toward degradation of Methylene Blue. Thin Solid Films, 2009, doi.10.1016/j.
tsf.2009.07.135
Tariq, M. A., Faisal, M., & Muneer, M. (2005). Semiconductormediated photocatalysed degradation of two selected azo dye
derivatives, amaranth and Bismarck brown in aqueous
suspension. Journal of Hazardous Materials, B127, 172179.
Tariq, M. A., Faisal, M., Saquib, M., & Muneer, M. (2008).
Heterogeneous photocatalytic degradation of an anthraquinone and a triphenylmethane dye derivative in aqueous
suspensions of semiconductor. Dyes and Pigments, 76,
358365.
Terzian, R., & Serpone, N. (1991). Heterogeneous photocatalysed oxidation of creosote components; mineralization of xylenols by illuminated TiO2 in oxygenated
aqueous media. Journal of Photochemistry and Photobiology A: Chemistry, 89, 163175.
Tian, B., Li, C., Gu, F., Jiang, H., Hu, Y., & Zhang, J. (2009).
Flame sprayed V-doped TiO2 nanoparticles with enhanced
photocatalytic activity under visible light irradiation.
Chemical Engineering Journal, 151, 220227.

Water Air Soil Pollut


Tsai, W.-T., Lee, M.-K., Su, T.-Y., & Chang, Y.-M. (2009).
Photodegradation of bisphenol-A in a batch TiO2 suspension
reactor. Journal of Hazardous Materials, 168, 269275.
Turchi, C. S., & Ollis, D. F. (1990). Photocatalytic degradation of
organic water contaminant mechanisms involving hydroxyl
radical attack. Journal of Catalysis, 122, 178192.
Venkatachalam, N., Palanichamy, M., & Murugesan, V.
(2007a). Solgel preparation and characterization of
alkaline earth metal doped nano TiO2: Efficient photocatalytic degradation of 4-chlorophenol. Journal of Molecular Catalysis A: Chemical, 273, 177185.
Venkatachalm, N., Palanichamy, M., Arabindo, B., & Murugesan,
V. (2007b). Enhanced photocatalytic degradation of 4chlorophenol by Zr4+ doped nano TiO2. Journal of
Molecular Catalysis A: Chemical, 266, 158165.
Vijayabalan, A., Selvam, K., Velmurugan, R., & Swaminathan,
M. (2009). Photocatalytic activity of surface fluorinated
TiO2-P25 in the degradation of Reactive Orange 4.
Journal of Hazardous Materials, 172, 914921.
Wang, H., Niu, J., Long, X., & He, Y. (2008). Sonophotocatalytic degradation of methyl orange by nano-sized Ag/
TiO2 particles in aqueous solutions. Ultrasonics Sonochemistry, 15, 386392.
Wawrzyniak, B., Morawski, A. W., & Tryba, B. (2006).
Preparation of TiO2-nitrogen-doped photocatalyst active
under visible light. International Journal of Photoenergy,
Article ID 68248, Pages 18 doi:10.1155/IJP/2006/68248
Wen, C., Zhu, Y.-J., Kanbara, T., Zhu, H.-Z., & Xiao, C.-F.
(2009). Effects of I and F codoped TiO2 on the photocatalytic degradation of Methylene Blue. Desalination,
249, 621625.
Wenhua, L., Hang, L., Suwan, C., Jianqing, Z., & Chanan, C. (2000).
Kinetics of photocatalytic degradation of aniline in water over
TiO2 supported on porous nickel. Journal of Photochemistry
and Photobiology A: Chemistry, 131, 125132.
Wong, M.-S., Hsu, S.-W., Koteswara, K., & Praveen Kumar, C.
(2008). Influence of crystallinity and carbon content on
visible light photocatalysis of carbon doped titania thin films.
Journal of Molecular Catalysis A: Chemical, 279, 2026.
Wu, C.-H., & Yu, C.-H. (2009). Effects of TiO2 dosage, pH and
temperature on decolorisation of C.I. Reactive Red 2 in a
UV/US/TiO2 system. Journal of Hazardous Materials,
169, 11791183.
Wu, C.-H., Kuo, C.-Y., & Chang, C.-L. (2008). Decolourization
of C.I Reactive Red 2 by catalytic ozonation processes.
Journal of Hazardous Materials, 153, 1521058.
Xiao, Q., & Ouyang, L. (2009). Photocatalytic activity and
hydroxyl radical formation of carbon-doped TiO2 nano-

crystalline: effect of calcinations temperature. Chemical


Engineering Journal, 148, 248253.
Xie, Y., & Zhao, X. (2008). The effects of synthesis
temperature on the structure and visible-light-induced
catalytic activity of FN-codoped and SN-codoped
titania. Journal of Molecular Catalysis A: Chemical, 285,
142149.
Yang, Y., Li, X.-J., Chen, J.-T., & Wang, L.-Y. (2004). Effect of
doping mode on the photocatalytic activities of Mo/TiO2.
Journal of Photochemistry and Photobiology A: Chemistry, 163, 517522.
Yi, Z., Ke-long, H., Zhi-ping, Z., & Chang-bin, X. (2007).
Eu2+/Gd3+-codoped nanocrystalline titania catalyst and its
photocatalytic activity under natural light. Transaction of
Nonferrous Metals Society of China, 17, 11121116.
Yuan, Z., Jia, J.-H., & Zhang, L.-D. (2002). Influence of codoping of Zn(II) and Fe(III) on the photocatalytic activity
of TiO2 for phenol degradation. Materials Chemistry and
Physics, 73, 323326.
Zhang, M., An, T., Hu, X., Wang, C., Sheng, G., & Fu, J.
(2004). Preparation and photocatalytic properties of a
nanometer ZnOSnO2 coupled oxide. Applied Catalysis
A: General, 260, 215222.
Zhang, H., Quan, X., Chen, S., Zhao, H., & Zhao, Y. (2006).
Fabrication of photocatalytic membrane and evaluation of
its efficiency in removal of organic pollutants from water.
Separation and Purification Technology, 50, 147155.
Zheng, W., Bingru, Z., & Fengting, L. I. (2007). A simple and
cheap method for preparation of coupled ZrO2/ZnO with
high photocatalytic activities. Front. Environ. Sci. Eng.
China, 1(4), 454458. doi:10.1007/s11783-007-0072-7.
Zhou, Y., Lu, S. X., & Xu, W. G. (2009). Photocatalytuc
activity of Nd-doped ZnO for the degradation of C.I.
Reactive Blue 4 in aqueous suspension. Environmental
progress and Sustainable Energy 2892), doi:10.1002/ep
Zhu, J., Zheng, W., He, B., Zhang, J., & Anpo, M. (2004).
Characterization of FeTiO2 photocatalysts synthesized by
hydrothermal method and their photocatalytic reactivity
for photodegradation of XRG dye diluted in water.
Journal of Molecular Catalysis A: Chemical, 216, 3543.
Zhu, X., Yuan, C., Bao, Y., Yang, J., & Wu, Y. (2005).
Photocatalytic degradation of pesticide pyridaben on TiO2
particles. Journal of Molecular Catalysis: Chemical, 229,
95105.
Zou, L., & Zhu, B. (2008). The synergistic effect of ozonation
and photocatalysis on color removal from reused water.
Journal of Photochemistry and Photobiology A: Chemistry, 196, 2432.

You might also like