You are on page 1of 15

Desalination 276 (2011) 13–27

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / d e s a l

An overview on the photocatalytic degradation of azo dyes in the presence of TiO2


doped with selective transition metals
M.A. Rauf ⁎, M.A. Meetani, S. Hisaindee
Chemistry Department, P.O. Box 17551, UAE University, Al-Ain, UAE

a r t i c l e i n f o a b s t r a c t

Article history: This article presents an overview on the degradation of azo dyes using transition metal doped TiO2 as
Received 12 January 2011 photocatalysts in aqueous solutions. Such dopants reduce the recombination of e−cb and h+vb and decrease
Received in revised form 24 March 2011 the band gap or create intra-band gap states, which result in more light absorption. Moreover, the addition of
Accepted 26 March 2011
these dopants can alter the surface properties of TiO2 catalyst such as surface acidity and surface area.
Available online 20 April 2011
Therefore, the photocatalysis on modified TiO2 can be promoted using visible light. An important role in this
Keywords:
process is played by molecular oxygen and other active species such as O•− • •
2 , HO2, H2O2 and OH which are

Azo dyes generated in a sequence of reactions. Besides this, the degradation of dyes depends on several other
Doping parameters such as pH, catalyst concentration, substrate concentration and the nature of the doping
Free radicals substances. Depending on the structure of the azo dye, the major identified intermediates are hydroxylated
Heterogeneous catalysis derivatives, aromatic amines, naphthoquinone, phenolic compounds and several organic acids. This review
Intermediates also presents the literature findings on the available pathways and mechanisms of degradation of some azo
Photocatalysis dyes in the presence of metal doped TiO2 catalysts.
Transition metals
© 2011 Elsevier B.V. All rights reserved.
TiO2

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.1. Classification of dyes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2. Experimental techniques used for studying dye degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3. Principles of photocatalysis and mechanistic pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.1. Direct photocatalytic pathway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2. Indirect photocatalytic mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4. Operational factors influencing the photocatalytic degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.1. Effect of dye concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.2. Effect of catalyst amount . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.3. Effect of pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.4. Effect of light intensity and irradiation time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.5. Dissolved oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5. Effect of dopants on dye degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.1. Doping TiO2 with Cr ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.2. Doping TiO2 with Cu ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5.3. Doping with Fe ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5.4. Doping with Mn ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5.5. Doping with Zn ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5.6. Doping with V ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.7. Doping with Ag ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.8. Doping with W ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

⁎ Corresponding author. Tel.: + 971 3 7136122; fax: + 971 3 7136944.


E-mail address: raufmapk@yahoo.com (M.A. Rauf).

0011-9164/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2011.03.071
14 M.A. Rauf et al. / Desalination 276 (2011) 13–27

1. Introduction [18–20], ozonation process [21], sonolysis [22–25], photocatalytic


approach [26–30], biodegradation [31–33] and the radiation induced
Synthetic dyes comprise an important part of industrial water degradation of dyes [34–37].
effluents, as they are discharged in abundance by many manufactur- Various physico-chemical techniques are available for the elimi-
ing industries. The impact of these dyes on the environment is a major nation of dyes from wastewater and in particular photocatalysis is a
concern because of the potentially carcinogenic properties of these more promising tool. Moreover, photocatalysis can be used to cause
chemicals [1]. Besides this, some dyes can undergo anaerobic redox transformations and decompose a dye molecule. The use of
decoloration to form potential carcinogens [2,3]. The wastewater photosensitive semiconductors such as TiO2, ZnO, Fe2O3, CdS, ZnS and
which is colored in the presence of these dyes can block both sunlight V2O5 has been reported in the literature for their use in reducing color
penetration and oxygen dissolution, which are essential for aquatic of the dye solutions owing to their environmental-friendly benefits in
life. Thus there is a considerable need to treat these colored effluents the saving of resources such as water, energy, chemicals, and other
before discharging them to various water bodies. Various approaches cleaning materials [38–42]. Titanium dioxide (TiO2) mediated based
on handling and decontaminating such effluents have been reported photodegradation has attracted extensive interest owing to its great
in the literature. Typical techniques include the classical methods advantages in the complete removal of organic pollutants from
such as adsorption [4–6], coagulation [7,8], ion flotation [9] and wastewater [43]. This is mainly because of its various merits, such as
sedimentation [10]. All these techniques are versatile and useful, but optical-electronic properties, low-cost, chemical stability, and non-
they all end up in producing a secondary waste product which needs toxicity [44]. The main aim of this paper is to produce a condensed and
to be processed further. Another set of techniques which are relatively coherent overview on the photocatalytic degradation of azo dyes in
newer, more powerful, and very promising is called Advanced the presence of TiO2 doped with selective transition metals, which can
Oxidation Processes (AOPs) which has been developed and employed serve as a ready reference for future scientific endeavors in this area.
to treat dye-contaminated wastewater effluents [11]. This methodol-
ogy would normally utilize a strong oxidizing species such as •OH 1.1. Classification of dyes
radicals produced in situ, which causes a sequence of reactions
thereafter to break down the macromolecules into smaller and less Dyes are mainly classified based on their structure, source, color and
harmful substances. In many cases the macromolecule is completely method of application in color index (C.I.), which has been continuously
mineralized into water and carbon dioxide. The AOP technique has edited since 1924. Depending on the chromophores, dyes can be
drawn considerable attention from various quarters of scientific classified in various classes. These include the acridine dyes, azo dyes,
community as it is easy to handle and produces significantly less arylmethane dyes, anthroqinone dyes, nitro dyes, xanthenes dyes and
residuals as compared to the classical approaches. Amongst the many quinine–amine dyes etc. Table 1 summarizes the structures of various
techniques employed in the AOP approach are the UV photolytic classes of dyes along with an example of each class. The investigation of
technique [12–14], Fenton process [15–17], photo-Fenton process radiation induced decomposition of dyes and their derivatives due to

Table 1
Classification of dyes based on molecular structure.

Class Structure Representative dye Structure


H H
Acridine Acridine O
N (H3C)2N N N(CH3)2

C C
H H
NaO3S SO3Na

N N
Azo N N Amido B N N
OH NH2
NO2

NH 2
Diarylmethane C Auramine O (H 3C)2N C N(CH 3)2

HO

O O

OH OH O CH 3
Anthraquinone Carmine OH CO 2H
OH
O
HO OH
OH O

C N(CH3)2
C
Triarylmethane Malachite green

N(CH3)2

ONa
OH
HO3S NO2
NO2
Nitro Naphthol

NO2
(C2H5)2N O N(C2H5)2

O
Xanthene Rhodamine B
CO2H

H 3C N CH3

N NH
H 2N N NH2
Quinone-imine Safranin O
N
M.A. Rauf et al. / Desalination 276 (2011) 13–27 15

Table 2
Representative examples of various types of azo dyes.

Class Representative dye Structure

Monoazo Orange II HO

NaO3S N N

NH2 H2N

Diazo Congo Red N N N N

NaO3S SO3Na

NaO3S HO

N N N N N N
Triazo Direct Blue 71
NaO3S
SO3Na
SO3Na

their environmental hazard is a rather popular subject and many papers then calculated from the decrease of absorbance of the dye solution at
have been published in this field [12–37]. its maximum absorption wavelength as follows [53]:
Azo dyes are the largest group of synthetic colorants (60–70%) and  
are being used in industry for applications such as textiles, papers, A
τ = 1− i × 100 ð1Þ
leathers, gasoline, additives, foodstuffs, cosmetics, laser materials, A0
xerography, laser printing, etc. and the resulted by-products contain
both dyes and metal ions. Since only 45–47% dyestuffs has been where, A0 and Ai are the absorbance values of the dye solution before
reported biodegradable, and residual color is mostly due to insoluble and after irradiation respectively.
dyes which have low biodegradability [33]. Because of the toxicity and A typical laboratory-scale experimental set up includes a vigor-
persistence of azo dyes, their removal from the wastewater has ously stirred batch photochemical cell, cylindrical flasks or vessels and
become an important issue of interest during the last few years. glass dishes [29,54,55], whereas, upflow type, membrane based and
Azo dyes are characterized by nitrogen to nitrogen double bonds coated surfaces have also been reported as photocatalytic reactors for
(–N=N–) that are usually attached to two moieties of which at least one dye degradation studies [56–59]. Fig. 1 depicts the standard
but usually both are aromatic groups (benzene or naphthalene rings). photochemical reactor used in this type of work.
Some representative examples of azo dyes are Methyl Orange, Acid
Orange 7, Acid Orange 20, Orange II, Methyl Red, Reactive Red 2, 3. Principles of photocatalysis and mechanistic pathways
Reactive Orange 16, Reactive Black 5, Congo Red, Solvent Red 1, Direct
Blue 160, Basic Yellow 15, Basic Blue 41, Disperse Orange 1, Disperse Red Among many known AOPs, photocatalytic degradation has proven
1, Amido Black, Remazol Brilliant Orange 3R, Amaranth, etc. Depending to be a promising technology for degrading organic compounds
upon the number of –N=N–groups in the molecule, the azo dyes can be [60,61]. The technique is more effective as compared to other AOPs
classified as monoazo (Acid orange 7, Orange G, Methyl red, etc.), diazo
(Congo Red), triazo (Reactive red 120, Naphthol blue black, etc.) and
triazine containing azo dyes (Reactive brilliant red K-2G, Procion red sampling syringe
MX-5B, etc.). Table 2 shows the representative example of each type of
azo dye. The color of azo dyes is determined by the azo bonds and their
associated chromophores and auxochromes. Azo bonds are the most
active bonds in azo dye molecules and can be oxidized by positive hole
or hydroxyl radical or reduced by electron in the conduction band. The
cleavage of –N=N– bonds leads to the decoloration of dyes.

2. Experimental techniques used for studying dye degradation


water in
water out
Various experimental techniques have been proposed for both
qualitative and quantitative analysis of various dyes concerning their
degradation. These methods generally involve instrumental analysis UV lamp
such as UV–vis spectrophotometry [45,46], LC/MS [47], HPLC [48], ion
chromatography [49], capillary electrophoresis [50], FTIR [51] and
NMR [52]. In many of these cases, prior procedures of extraction of the
aqueous sample with an organic solvent or filtration are adopted
when a heterogeneous catalyst or solid reactant is employed.
Since dyes are colored in nature, it is relatively easy to monitor
their color change during the course of the experiment. In photo-
catalytic studies, the sample is normally exposed to UV irradiation for magnetic
a given time period and the changes in solution property (such as stirrer
color) are observed by using a spectrophotometer. The solution after
irradiation is then transferred to a cuvette and subjected to absorption
studies on a spectrophotometer. The degree of decoloration (τ) is Fig. 1. Schematic diagram of a photoreactor.
16 M.A. Rauf et al. / Desalination 276 (2011) 13–27

because semiconductors are inexpensive and can easily mineralize


various organic compounds [62].
The basic steps for heterogeneous photocatalysis consist of initially
transferring of the reactants in the liquid phase onto the catalyst
surface followed by adsorption of the reactant on the catalyst surface,
reaction in the adsorbed phase, desorption of the final product and
finally removal of the final products in the liquid phase.
The photocatalytic degradation of an organic compound such as
dye is believed to take place according to the following mechanisms:
When a catalyst is exposed to UV radiation, electrons are promoted
from the valence band to the conduction band. As a result of this, an
electron–hole pair is produced [63]
− þ Fig. 2. Mechanism of generation of oxidative species in a photocatalytic study.
Catalyst þ hν→ecb þ hvb ð2Þ

where, e− +
cb and h vb are the electrons in the conduction band and the
where r is the reaction rate for the oxidation of reactant
electron vacancy in the valence band, respectively. Both these entities can
(mg/L min), kr is the specific reaction rate constant for the
migrate to the catalyst surface, where they can enter in a redox reaction
oxidation of the reactant (mg/L min), ka is the equilibrium
with other species present on the surface. In most cases h+ vb can react
constant of the reactant (L/mg) and C is the dye concentration.
easily with surface bound H2O to produce •OH radicals, whereas, e− cb can
When the chemical concentration Co is a millimolar solution (Co
react with O2 to produce superoxide radical anion of oxygen [59]
is small) the equation can be simplified to an apparent first-order
þ •
H2 O þ hvb → OH þ H
þ
ð3Þ equation:

− −    
O2 þ ecb →O2 ð4Þ C0
ln = kapp t or C = Co exp −kapp t ð10Þ
C
This reaction prevents the combination of the electron and the
hole which are produced in the first step. The •OH and O−2 produced in
the above manner can then react with the dye to form other species A plot of ln Co/C versus time represents a straight line, the slope of
and is thus responsible for the discoloration of the dye. which upon linear regression equals the apparent first-order rate
constant kapp. Generally first-order kinetics is appropriate for the
−• þ •
O2 þ H →HO2 ð5Þ entire concentration range up to few ppm and several studies

were reasonably well fitted by this kinetic model. The L–H model
H2 O2 →2 OH ð6Þ was established to describe the dependence of the observed
• 9 10 –1 –1 reaction rate on the initial solute concentrations. Representative
OH þ dye→dyeox ðk ¼ 10 −10 M s Þ ð7Þ
examples of dye degradation adhering to such a process are

Dye þ ecb →reduction products ð8Þ reported in the literature [65,66].
(ii) The Eley–Rideal process
In this process, the free charged carriers are initially photo-
It may be noted that all these reactions in photocatalysis are
fragmented followed by subsequent trapping of the holes by
possible due to the presence of both dissolved oxygen and water
surface defects. The surface active centers (S) can then react
molecules. Without the presence of water molecules, the highly
with the dye (chemisorption) to form an adduct species such as
reactive hydroxyl radicals (•OH) could not be formed and inhibit the
(S-dye)+ which can further decompose to produce products or
photodegradation of liquid phase organic molecules.
can recombine with electrons. The reaction scheme is outlined
A schematic presentation of the mechanisms of generation of
below [67]:
oxidative species in a photocatalytic study is shown in Fig. 2. Direct and
indirect photocatalytic pathways are the two suggested mechanisms for − þ
Catalyst þ hν→e þ h ðphotogeneration of free carriersÞ ð11Þ
a given photocatalytic reaction. These are briefly discussed below.
þ þ
S þ h →S ðhole trapping by surface defectsÞ ð12Þ
3.1. Direct photocatalytic pathway
þ −
S þ e →Sðphysical decay of active centersÞ ð13Þ
Two different approaches have been suggested for this type of
mechanism: þ þ
S þ dye→ðS  CPÞ ðchemisorptionÞ ð14Þ
(i) Heterogeneous photocatalysis—the Langmuir–Hinshelwood
þ −
process ðS  dyeÞ þ e →S þ products ð15Þ
Heterogeneous photocatalysis can be generally explained on
the basis of the Langmuir–Hinshelwood process, wherein the 3.2. Indirect photocatalytic mechanism
production of electrons and holes produced by the photoexci-
tation of the catalyst are taken into account. The hole produced In this process, electron–hole pairs are photogenerated on the
is trapped by the adsorbed dye molecule on the catalyst surface surface of the catalyst. The hole is then trapped by water molecules
to form a reactive radical state which can decay as a result of leading to the formation of •OH radicals and H+. These •OH radicals can
recombination with an electron. The catalyst is regenerated in combine to form H2O2 or they can attack the dye to produce
the process. The Langmuir–Hinshelwood (L–H) equation can intermediates and end products. Moreover, the electrons can be trapped
be expressed as follows [64]: by molecular oxygen to form superoxide radical which can enter into a
chain reaction to produce HO•2 and H2O2. Finally, the radicals formed
1 1 1 during this mechanism are responsible for the oxidation of the organic
= + ð9Þ
r kr kr ka C molecule producing intermediates and end products [54].
M.A. Rauf et al. / Desalination 276 (2011) 13–27 17

The stepwise mechanism is illustrated below: 4.3. Effect of pH


− þ
TiO2 þ hν→e þ h ð16Þ Photodegradation efficiency of dyes is affected by the pH of the
þ • þ solution. The variation of solution pH changes the surface charge of
h þ H2 OðadsÞ→ OHðadsÞ þ H ðadsÞ ð17Þ
TiO2 particles and shifts the potentials of catalytic reactions. As a
O2 þ e
− −•
→O2 ðadsÞ ð18Þ result, the adsorption of dye on the surface is altered thereby causing a
change in the reaction rate. Under acidic or alkaline condition the
−• þ • surface of titania can be protonated or deprotonated respectively
O2 ðadsÞ þ H ↔HO2 ðadsÞ ð19Þ
according to the following reactions [63]:

HO2 ðadsÞ→H2 O2 ðadsÞ þ O2 ð20Þ
þ þ
TiOH þ H →TiOH2 ð23Þ

H2 O2 ðadsÞ→2HO ðadsÞ ð21Þ
− −

TiOH þ OH →TiO þ H2 O ð24Þ
OH þ dye→intermediates→CO2 þ H2 O ð22Þ
Thus titania surface will remain positively charged in acidic
medium and negatively charged in alkaline medium. Titanium dioxide
4. Operational factors influencing the photocatalytic degradation
is reported to have higher oxidizing activity at lower pH, but excess H+
can decrease reaction rate. TiO2 behaves as a strong Lewis acid due to
The oxidation rates and efficiency of the photocatalytic system are
the surface positive charge. The anionic dye on the other hand is a
highly dependent on a number of operational parameters that govern
fused polynuclear aromatic compound with an extensive π electron
the photodegradation of the organic molecule [68]. This section will
conjugation which can easily form a stable complex by donating
briefly discuss the significance of each operational parameter.
electrons to the vacant d orbital of titanium. In other words, the anionic
dye acts as a strong Lewis base and can easily adsorb on the positively
charged catalyst surface. This favors the adsorption of the dye under
4.1. Effect of dye concentration
acidic conditions, while in the alkaline conditions this complexation
process is not favored presumably because of competitive adsorption
The quantity of the dye adsorbed on the surface of the photo-
by hydroxyl groups and the dye molecule in addition to the Coulombic
catalyst is of foremost importance since only this amount contributes
repulsion due to the negatively charged catalyst with the dye molecule
to photocatalytic process and not the one in the bulk of the solution.
[71].
The extent of dye adsorption depends on the initial dye
The extent of dye adsorption depends on the initial dye
concentration. The initial concentration of dye in a given photo-
concentration, nature of the dye, surface area of photocatalyst and
catalytic reaction is an important factor which needs to be taken into
pH of the solution. The pH determines the surface charge of the
account. Generally speaking the percentage degradation decreases
photocatalyst. Adsorption of the dye is minimum when the pH of the
with increasing amount of dye concentration, while keeping a fixed
solution is at the isoelectric point (point of zero charge). The surface of
amount of catalyst [69].
the photocatalyst is positively charged below isoelectric point and
This can be rationalized on the basis that as dye concentration
carries a negative charge above it. Depending on the nature of dye that
increases, more organic substances are adsorbed on the surface of
needs to be adsorbed on the surface of a photocatalyst, the adsorption
TiO2, whereas less number of photons are available to reach the
can be low or high in acidic and basic media.
catalyst surface and therefore less •OH are formed, thus resulting in
At low pH, reduction by electrons in conduction band plays a very
less degradation percentage.
important role in the degradation of dyes due to the reductive
cleavage of azo bonds [72–74]. Acid Orange 7 (an anionic dye) has
shown to degrade more at pH 3 [75]. The rationale behind this is that
4.2. Effect of catalyst amount
at low pH values, more H+ are available for adsorption to mask the
surface of the catalyst thus preventing the photoexcitation of
Dye degradation is also influenced by the amount of the
semiconductor particles, thereby reducing the generation of free
photocatalyst and aggregation of catalyst particles in high amounts
radicals. In alkaline solutions, •OH radicals are easier to be generated
of catalyst. The dye degradation increases with increasing catalyst
by oxidizing more hydroxide ions available on TiO2 surface, thus the
concentration, which is characteristic of heterogeneous photocata-
efficiency of the process is increased [76]. Similar results have been
lysis [11]. The increase in catalyst amount actually increases the
reported in the photocatalysed degradation of diazo and triazo dyes
number of active sites on the photocatalyst surface thus causing an
[77,78]. Fig. 3 shows the trend in % degradation of azo dye at different
increase in the number of •OH radicals which can take part in actual
pH values.
discoloration of dye solution. Beyond a certain limit of catalyst
amount, the solution becomes turbid and thus blocks UV radiation for
the reaction to proceed and therefore percentage degradation starts 4.4. Effect of light intensity and irradiation time
decreasing [70]. A representative example showing the effect of
catalyst amount on % decoloration is presented in Table 3. Both light intensity and time of irradiation affect the dye
degradation [42,79]. It has been shown that at low light intensities
(0–20 mW/cm2), the rate would increase linearly with increasing
Table 3
Effect of V2O5/TiO2 on % decoloration of dye solution in the presence of UV( concentration light intensity (first order), whereas at intermediate light intensities
of TB = SO= 80 μM, CV= 40 μM, irradiation time = 20 min, pH= 6.1) [93]. (25 mW/cm2) the rate would depend on the square root of the light
intensity [80]. At high light intensities the rate is independent of light
Dye Catalyst amount (mg/20 mL)
intensity, because at low light intensity reactions involving electron–
10 20 25 30 40
hole formation are predominant and electron–hole recombination is
Toludine Blue (TB) 15 40 47 41 27 negligible. On the other hand, when light intensity is increased, the
Safranin Orange (SO) 8 11 21 12 – electron–hole pair separation competes with recombination, thereby
Crystal Violet (CV) 6.5 12 – 17 15
causing lower effect on the reaction rate. In many literature studies, it
18 M.A. Rauf et al. / Desalination 276 (2011) 13–27

100
known to induce the cleavage mechanism for aromatic rings in
A07 organic pollutants that are present in the water matrices [39].
80 MO
Degradation %

5. Effect of dopants on dye degradation


60
Heterogeneous photocatalysis involving titanium dioxide (TiO2)
40 appears to be the most promising technology for organic dyes
degradation. However one of the major problems in using TiO2 as a
20 catalyst is the low photo-quantum efficiency which arises from the fast
recombination of photogenerated electrons and holes. Moreover, TiO2
0 is inactive under visible light due to its wide band gap (3.03 eV for
1 2 3 4 5 7 9 rutile and 3.18 for anatase form). This inherently causes the inability to
pH make use of the vast potential of solar photocatalysis. Various
techniques have been employed to make TiO2 absorb photons of
Fig. 3. Comparison of photocatalytic degradation of Acid Orange 7 and Methyl Orange
lower energy as well. These techniques include surface modification
by 4.0% WOx/TiO2 composite at different pH values [133].
via organic materials and semiconductor coupling, band gap modifi-
cation by creating oxygen vacancies and oxygen sub-stoichiometry, by
nonmetals including co-doping of nonmetals and metal doping.
Dopants, such as transitional metals have been added to the TiO2
has been shown that the dye decolorization initially increases as the catalyst to improve its response and also reduce the recombination of
light intensity is increased [81,82]. The reaction rate decreases with photogenerated electrons and photogenerated holes [84–86]. The
irradiation time as it follows the pseudo first-order kinetics and main objective of doping is to induce a batho-chromic shift, i.e., a
additionally a competition for degradation may occur between the decrease of the band gap or introduction of intra-band gap states,
reactant and the intermediate products. The slow kinetics of dye which results in more visible light absorption. The effect of metal ion
degradation after certain time limit is mainly attributed to the dopants on the photocatalytic activity is a complex problem. The total
difficulty in the reaction of short chain aliphatics with •OH radicals, induced alteration of the photocatalytic activity is made up from the
and the short lifetime of photocatalyst because of active sites sum of changes which occur in the light absorption capability of the
deactivation by strong by-products deposition [83]. Fig. 4 shows the TiO2 photocatalyst, adsorption capacity of the substrate molecules at
effect of time on % dye degradation. the catalyst's surface and interfacial charge transfer rate. Using
suitable transitional metals as dopants improves the performances
of TiO2 [87–90]. This results in the overlap of the conduction band due
4.5. Dissolved oxygen to Ti (3d) with d levels of the transition metals causing red shift of the
band edge of TiO2. It can thus also allow the light absorption to be
Oxygen dissolved in solution is commonly employed as an widened into the visible region to various extents, depending on the
electron acceptor in photocatalysis reaction to assure sufficient type of the dopant and its concentration. Therefore the photocatalysis
electron scavengers present to trap the excited conduction band on modified TiO2 can be promoted using visible light.
electron from recombination [39]. The oxygen does not affect the Various transition metals, such as Ni2+, Zn2+, Cr3+ and Fe3+ can be
adsorption on the TiO2 catalyst surface as the reduction reaction takes easily incorporated into the crystal lattice of TiO2 because of their similar
place at a different location from where oxidation occurs. Dissolved ionic radii (The ionic radii of Ni2+ = 0.72 Å, Zn2+ = 0.74 Å, Cr3+ =
oxygen involves in the stabilization of radical intermediates, miner- 0.76 Å and Fe3+ = 0.69 Å are quite similar to that of host Ti4+ = 0.75 Å
alization and direct photocatalytic reactions. Its presence is also ions) [91]. The degradation of dyes is usually faster in mixed systems as
compared to single systems, because the oxidation of dyes consumes
photo-excited holes promptly and efficiently, thus attenuating electron–
hole recombination. In mixed systems this might be because the
recombination is increased by crystallite defects [92–94].

5.1. Doping TiO2 with Cr ions

The incorporation of transition metal ions in TiO2 is interesting for


inducing a batho-chromic shift of the band gap. The band gap energy is
shifted by 2.00 eV in the case of chromium doping [95]. This can be
explained by the excitation of an electron of Cr3+ into the conduction
band of TiO2. For any photocatalytic reaction, the lifetimes of electrons
and holes must be long enough to allow them to reach the surface of
the photocatalyst. The addition of transition metal ions produces new
trapping sites which affect the lifetime of the charge carriers. A lifetime
of 89.3 ms has been reported for pure titania, whereas doping reduces
this lifetime to about 30 ms in the case of Cr3+ [95]. When Cr3+ ions are
present in the matrix, they are initially oxidized by holes (h+):

3þ þ 4þ
Cr þ h →Cr ð25Þ

The beneficial effect of Cr3+ under UV irradiation can be explained


by considering the efficient separation of photo-excited electrons and
Fig. 4. Effect of irradiation time on % degradation of an azo dye. holes. Cr3+ can act as photogenerated hole trapper, due to the energy
M.A. Rauf et al. / Desalination 276 (2011) 13–27 19

level for Cr3+/Cr4+ above the valence band edge of TiO2 [95]. The Rhodamine B undergoes 25% more degradation in the presence of
trapped holes in Cr4+ can migrate to the surface absorbed hydroxyl 10% Cr3+ doped titania as compared to in neat titania [95]. At chromium
ion to produce hydroxyl radicals: concentrations below 1%, both the lifetime and photocatalytical activity
were reported to decrease drastically to a nearly constant low level. This
Cr
4þ −
þ OH ðadsÞ→Cr
3þ •
þ OHðadsÞ ð26Þ can be explained on the basis that is that at low concentrations of Cr3+,
these ions can occupy regular lattice sites which act as traps.
The reaction scheme is well supported by the photocatalytic At higher ion concentrations of Cr3+ ions, the interstitial sites are
chromium cycle which has been reported in the literature [96]. The re- filled up and thus do not contribute to the recombination. In another
oxidation of photoreduced chromium species takes place with study, higher concentration of chromium ions (10%) was used so that
molecular oxygen. In the photocatalytic degradation process, organic they are completely dispersed into the titania lattice [26]. The 10% of
matter, molecular oxygen and UV light are consumed in the chromium was selected for two reasons: (a) 10% Cr3+ could be well
chromium cycle, which enables the atmospheric oxygen to oxidize dispersed in the titania lattice without the formation of any crystalline
and thus degrade organic pollutant [97]. chromium oxide, (b) 10% chromium resulted in higher surface area

Fig. 5. Schematic diagram of various Intermediate formations of Orange G dye as a result of photocatalytic degradation [99].
20 M.A. Rauf et al. / Desalination 276 (2011) 13–27

than other compositions and also that the anatase phase is retained It was shown that the Cu2+ doped TiO2 catalyst increased the
which is known to be the most catalytically active phase of titania photocatalytic degradation of two typical azo dyes used in dyeing
[98]. Chromium–titanium mixed oxide was used to photodegrade industry namely Acid Orange 7 (AO7) and Tartrazine (Tart) [104]. The
Orange G dye (azo dye) and the degradation followed the pseudo behaviour of Cu2+ was explained on the basis that Cu2+ may scavenge
first-order kinetics with a degradation efficiency of 70% in 300 min electrons resulting in the formation of Cu+, which in turn enhances
[99]. The degraded products were analyzed by using the liquid the oxidation of the substrate.
chromatography tandem mass spectrometry. Copper doped TiO2 nanocatalysts have been synthesized by photo-
The major intermediates were substituted phenols, and aromatic deposition and sol–gel methods and their activities in the discolor-
hydroxyl amines in addition to nitroso compounds and dicarboxylic ation and mineralization of AO7 were evaluated [105]. The results
aromatic compounds. The degradation pathway scheme (Fig. 5) indicated that the Cu-doped TiO2 nanocatalysts with a low copper
suggests that these products were produced by desulfonation and concentration prepared by the photo-deposition method showed
hydroxylation of the aromatic ring together with oxidative cleavage of enhanced photocatalytic activity; TiO2 nanocatalyst doped with 1% Cu
the azo bond and the destruction of the aromaticity and ring cleavage. showed the best performance. Complete color removal and 99% of
Desulfonation was explained by the attack of •OH radicals onto total organic carbon (TOC) removal were achieved after 150 min of
sulphur followed by its elimination in the form of H2SO4 or NaHSO4. reaction. TiO2 nanocatalysts doped with more than 1% Cu concentra-
The remaining aromatic radical species were quenched to produce tion blocks the TiO2 from absorbing the incoming photons thereby
phenols and hydroxyl amines. The findings also showed that π bonds showing decreased photocatalytic activities.
in the azo group are the preferable site for •OH radical attack which Several heterojunctions such as Cu2O/TiO2, Cu/Cu2O/TiO2 and
can happen on either of the nitrogen atoms of the azo group thus Cu/Cu2O/CuO/TiO2 were prepared and studied for their potential
leading to a variety of degradation products [26,100]. In another application as photocatalysts. They were found to show high perfor-
study, chromium doped titania electrodes were used to degrade Acid mance under visible light [106]. Orange II was used as a representative
Red G (azo dye) [101]. The photocatalytic efficiency was remarkably dye molecule in this study. The highest photocatalytic activity was
enhanced by the incorporation of Cr3+. The enhanced photocatalytic observed for the Cu/Cu2O/CuO/TiO2 system under visible light. It is well
degradation rate was attributed to the increase of the charge known that Cu cannot act as a photosensitizer, but it can mediate the
separation in these systems. electron injection from Orange II to the conduction band of TiO2 making
the generation of radical O2•− possible.

5.2. Doping TiO2 with Cu ions


5.3. Doping with Fe ions
In solution, Cu(I) produced from Cu(II) by photoreduction
undergoes relatively fast re-oxidation mostly in reaction with TiO2 particles can be easily doped by iron, forming mixed oxides or
molecular oxygen [96]: mixtures of simple/mixed oxides. Iron cations occupy substitutional
positions because of the similar radius of Fe3+ and Ti4+ which form
þ 2þ •−
Cu þ O2 →Cu þ O2 ð27Þ solid solutions with titania at low concentrations (b1 at Fe%). The
presence of iron catalyzes the anatase to rutile transformation, with
Consequently Cu(I) disappearance in oxic environments is very fast
rutile being detected even at high temperature. The formation of solid
(in milliseconds). In suboxic conditions, the Cu(I) concentration varies
solutions and the existence of solubility limits for Fe ions in TiO2 lead
with solar UV irradiation and can even reach 50% of the dissolved copper
to formation of α-Fe2O3 at higher concentrations.
[102]. Oxygen dissolved in water reduces the Cu(I) concentration and
In the presence of low doping levels of this metal, photo generated
thus O•2− and H2O2 are produced instead. These play an important role in
holes and electrons are well separated thereby increasing the
the Cu cycle as they can act either as an oxidant or a reducing agent with
efficiency of the catalyst. Fe3+ traps photo generated holes and
copper. The oxidation can take place by the following steps [96]:
forms Fe4+ which reacts with the surface adsorbed hydroxyl ions to
þ 2þ • − produce hydroxyl radicals and O2− in the surface lattice of doped TiO2
Cu þ H2 O2 →Cu þ OH þ OH ð28Þ
and tends to trap photogenerated holes and produce hydroxyl
þ
Cu þ OH→Cu
• 2þ
þ OH

ð29Þ radicals. Alternatively, Fe4+ can also react with photogenerated
electrons thereby affecting the photocatalytic activity. The reaction
whereas, reduction can occur by the following reactions: mechanism is summarized below [107]:
2þ − þ •− þ − þ
Cu þ HO2 →Cu þ O2 þ H ð30Þ TiO2 þ hν→e þ h ð32Þ

2þ •− þ 3þ þ 4þ
Cu þ O2 →Cu þ O2 ð31Þ Fe þ h →Fe ð33Þ
Thus peroxide and superoxide species are not only formed in the 3þ − 2þ
Fe þ e →Fe ð34Þ
copper cycle, but they also had an influence on the Cu(I)/Cu(II)
interconversion. 2þ 3þ −
Fe þ O2 ðadsÞ→Fe þ O2 ð35Þ
It is now well known that copper can participate in photoredox
cycles and its role cannot be ignored. The most important part of the 2þ 4þ 3þ 3þ
Fe þ Ti →Fe þ Ti ð36Þ
cycle is photoreduction of Cu(II) to Cu(I) induced by solar light and
oxidation of copper bound pollutants to the environmentally benign 3þ 4þ −
Ti þ O2 ðadsÞ→Ti þ O2 ð37Þ
forms. Therefore, the Cu(I) is oxidized easily to Cu(II), which may
4þ − 3þ
coordinate with another molecule. Thus the Cu photocatalytic cycle Fe þ e →Fe ð38Þ
contributes to the environmental cleaning [96].
4þ − 3þ •
Photooxidation of pollutants with participation of Cu(II) complexes Fe þ OH ðadsÞ→Fe þ OHðadsÞ ð39Þ
may also occur in heterogeneous systems on semiconductor surfaces.
Since this system requires presence of other substances such as TiO2 or The presence of high concentration of Fe3+ ions can act as
WO3, it seems to be more useful for wastewater cleaning from recombination centers for the photogenerated electrons and holes,
hazardous species rather than for environmental self-cleaning [103]. resulting in the decrease of photocatalytic activity.
M.A. Rauf et al. / Desalination 276 (2011) 13–27 21

The excitation behavior of Fe–TiO2 (under visible irradiation) arises The photogenerated electrons preferentially go to Mn ion rather
from electronic transition from the dopant energy level (Fe3+/Fe4+) to than oxygen as the transfer time to oxygen is reported to be of the
the conduction band of TiO2. Due to the fact that the t2g level of 3d orbital order of milliseconds [109]. Since Mn3+ and Mn+ ions are relatively
of Fe3+ ion is above the valence band of TiO2, Fe3+ ion can absorb a unstable as compared to Mn2+ ions, (d5), therefore there is tendency
photon with a wavelength exceeding 400 nm to produce a Fe4+ ion and for the transfer for the trapped charges from Mn3+ and Mn+ to the
a TiO2 conductive band electron [107]. The conductive band electron interface to initiate the following reactions [110].
further reacts with adsorbed O2 to form O2 −, while Fe4+ reacts with
þ 2þ •−
surface hydroxyl group to produce hydroxyl radical. Mn þ O2 →Mn þ O2 ðelectron releaseÞ ð53Þ

3þ 4þ −
Fe þ hν→Fe þe ð40Þ 3þ − 2þ •
Mn þ OH → Mn þ OHðhole releaseÞ ð54Þ

e þ

O2 ðadsÞ→O2 ð41Þ The photogenerated superoxide ion (O•−
2)
and hydroxyl radical
(•OH) are highly reactive and can effectively degrade the dye.
The photocatalytic activity of Fe–TiO2 catalysts was evaluated from In the case of Mn2+ doping, the rates of detrapping of the charge
the photodegradation of Methyl orange (azo dye) aqueous solution carriers are high which accounts for its higher activity. Additionally,
both under UV and visible light irradiation [107]. Fe3+ doping the induced oxygen vacancies in Mn2+–TiO2 samples can efficiently
effectively improved the photocatalytic activity under both UV light capture the photo generated electrons thereby reducing the recom-
irradiation and visible light irradiation with an optimal doping bination rate. These oxygen vacancies enhance the process of
concentration of 0.1 and 0.2% respectively. The results were explained adsorption of oxygen and increase the interaction between photoin-
on the basis that appropriate Fe3+ doping can efficiently separate the duced electrons of oxygen vacancies and adsorbed oxygen favoring
photogenerated electrons and holes and consequently improves the the photo catalytic reactions which increases the adsorption of dye
photocatalytic activity, however, no product formation was reported molecules [111].
in this work. In the presence of visible light irradiation, Fe3+ doping TiO2 was doped with Mn2+ and its photocatalytic activity was
introduces a new energy level (Fe3+/Fe4+) above the valence band, studied under solar light for the degradation of Methyl Orange and
resulting in the enhancement of light absorption in visible light Amaranth azo dye [60,112]. Mn2+–TiO2 (0.06%) showed highest
regions and the improvement of photocatalytic activity. photocatalytic activity due to the synergetic effect observed in the
In another study using the iron modified TiO2, 0.1 wt.% was found mixed phase and also due to its unique half filled electronic
to be the optimum dopant concentration for the maximum removal of configuration.
Reactive Brilliant X-3B azo dye [108]. It was found that if excessive Methyl Orange (azo dye) has been degraded by using Mn-doped
iron was deposited on TiO2, more Fe(OH)2+ would be formed. The TiO2/SiO2 matrix [113]. It was shown that the photocatalytic activity
light energy falling on the surface of TiO2 decreased due to greater of TiO2/SiO2 enhanced after manganese doping. The photocatalytic
absorption of the incidence light in the range of 290–400 nm by Fe activity of Mn-doped TiO2/SiO2 increased with Mn content and then
(OH)2+. As a result, the removal efficiency of Reactive Brilliant X-3B diminished when the Mn-doped content was increased to 0.5%. When
was found to decrease. Mn-doped content reached 0.8%, the photocatalytic reactivity was
similar to that of TiO2/SiO2. Compared to TiO2/SiO2, the high
5.4. Doping with Mn ions photocatalytic activity of Mn-doped TiO2/SiO2 is due to the variable
valences of manganese. Since Mn2+ and Mn3+ coexist on the surface
Photodegradation on Mn-doped TiO2 surface can be represented of Mn/TiO2/SiO2 photocatalyst, metal ions are the acceptors of
by the following mechanism: electrons, and Mn3+ can trap electrons in TiO2 conduction band.
Thus Mn3+ should trap electrons and prohibit electron–hole
þ −
TiO2 þ hν→hvb þ ecb ð42Þ recombination. Moreover, the electrons trapped in Mn3+ site are
subsequently transferred to the adsorbed O2, and decrease the
4þ − 3þ
Ti þ ecb → Ti ð43Þ recombination rate of the electron–hole pair.
2− − −
O þ hν→O þ e ð44Þ
5.5. Doping with Zn ions
2+
In the presence of Mn ion, the following reactions are possible
[109]: The ionic radii of Zn2+ (0.74 Å) is similar to that of host Ti4+
2þ − þ
(0.75 Å) ions. Hence this ion can easily substitute Ti4+ ion in TiO2
Mn þ ecb →Mn ð45Þ lattice without distorting the crystal structure, thereby stabilizing the
þ 2þ •−
anatase phase over a range of dopant concentrations.
Mn þ O2 ðadsÞ→Mn þ O2 ð46Þ The presence of Zn2+ as a dopant can act as follows [91]:
2þ þ 3þ
Mn þ hvb →Mn ð47Þ
2þ − þ
Zn þ e →Zn ð55Þ
3þ − 2þ •
Mn þ OH →Mn þ OH ð48Þ þ 2þ −•
Zn þ O2 ðadsÞ→Zn þ O2 ð56Þ
Alternatively, the Mn 3+ can also trap conduction band electron or þ þ 2þ
Mn+ can trap valence band hole to retain half filled electronic Zn þ h →Zn ð57Þ
structure of Mn2+ 2þ þ 3þ
Zn þ h →Zn ð58Þ
þ þ 2þ
Mn þ hvb →Mn ð49Þ −
þ• OH
3þ 2þ
Zn þ OH →Zn ð59Þ
3þ − 2þ
Mn þ ecb →Mn ð50Þ 3þ − 2þ
Zn þ e →Zn ð60Þ
2þ 3þ þ 4þ
Mn þ Ti →Mn þ Ti ðelectron trapÞ ð51Þ
The inclusion of Zn2+ ions as dopants in TiO2 can thus generate O−•
2

2þ − 3þ 2−
and •OH radicals which can then take part in subsequent reactions to
Mn þ O →Mn þ O ðhole trapÞ ð52Þ effectively increase the dye degradation.
22 M.A. Rauf et al. / Desalination 276 (2011) 13–27

The 0.1% Zn–TiO2 nanocatalysts exhibited the best photodegradation accepting a hole. Both O•2− and •OH can further degrade the dye
of Acid Orange 52 [41]. The appearance of an optimal dopant molecule. As a result, the introduction of appropriate amount of V4+
concentration was figured out to be a result of the delicate balance of ions in TiO2 lattice can restrain the recombination rate of photo-
an increase in trap sites leading to efficient trapping and fewer trapped generated electrons and holes, enhancing the photocatalytic activity
carriers leading to longer lifetimes for interfacial charge transfer. of TiO2.
The photocatalytic degradation of Congo Red (CR) has also been In the case of vanadium doped catalyst, it was found that Congo
reported in the presence of Zn doped TiO2 [114]. It was found that Zn2+– Red degraded to almost 53% in the presence of V5+–TiO2 against to
TiO2 showed 100% dye degradation of Congo Red when Zn2+ was 0.06%. only 6% in neat TiO2 [114]. Vanadium doped TiO2 nanocrystalline was
The higher activity in the case of Zn2+ (0.06%) –TiO2 was attributed to prepared and used to degrade Methyl Orange (azo dye) in solution
the generation of excess hydroxyl radicals due to the prolonged [121]. The dye degraded 55% more by using the doped catalyst as
separation of charge carriers. At this optimum concentration of Zn2+, compared to the neat TiO2. The photocatalytic activity of vanadium
the surface barrier becomes higher and the space charge region gets doped TiO2 film was evaluated through degradation of Reactive
extended, leading to the efficient separation of electrons and holes. Brilliant Red(azo dye) under visible light, and was compared to pure
Although literature findings on the degradation of azo dyes using titania [122]. Results showed that the photocatalytic degradation of
zinc doped titania are limited, there are quite a few studies on using the dye was 60% in the case of doped catalyst as to only 7% in pure TiO2
the ZnO/TiO2 composites as potential catalysts for degrading azo dyes. film. This was explained in terms of narrowing the band gap which
The coupling of different semiconductor oxides seems useful in order causes the red shift of the absorption band edge by vanadium doping
to achieve a more efficient electron–hole pair separation under as shown in other studies [123,124].
irradiation and, consequently, a higher photocatalytic activity.
In the TiO2/ZnO composite, the electron transfer occurs from the 5.7. Doping with Ag ions
conduction band of light-activated ZnO to the conduction band of
light-activated TiO2 and, conversely, hole transfer can take place from Molecular oxygen is the only electron accepting species in
the valence band of TiO2 to the valence band of ZnO [115]. This undoped titania. However, in Ag doped titania, two additional
efficient charge separation increases the photocatalytic activity of electron accepting species are introduced, namely silver ions (Ag+)
TiO2/ZnO composite. The nanosized TiO2/ZnO powders were found to and metallic silver (Ag0). Since the Fermi level of titania is higher, the
display high photocatalytic activity towards the decolorization of electrons would migrate from the titania film to the metallic Ag
Basic Blue 41(azo dye) in water under solar irradiation and a Ti/Zn particles resulting in a space charge layer at the boundaries between
molar ratio of 1:1, and approximately 100% color removal was Ag and titania. The electric field would eventually drive the electrons
achieved in 1 h [116]. Likewise Acid Red 14 and Methyl Orange (both to the interior and abstracts holes to the interfacial region of the
azo dyes) were effectively degraded using a nanosized powder of titania film. In this situation, silver particles act as electron traps [125].
TiO2/ZnO composite [117,118]. This results in the efficiency of the charge carrier separation thereby
Treatment of Acid Orange 7 (azo dye) with UV/ZnO, UV/[(TiO2 + resulting in the inhibition of recombination.
ZnO)] led to complete decolorization [119]. The catalytic effect of The most significant advantage of the Ag doping process is
UV/(TiO2 + ZnO) was found to be higher than UV/ ZnO and UV/TiO2. improving the charge separation efficiency of the titania. Due to
improved separation efficiency, the lifetime of charge carriers will
increase and higher number of holes will reach the interfacial region
5.6. Doping with V ions of the titania film and produce more superactive radicals such as HO•2,
OH• and H2O2 [125].
Most catalysts based on vanadium oxide consist of a vanadium oxide Another explanation for the role of metal silver on the catalyst
phase deposited on the surface of an oxide support, such as SiO2, Al2O3, surface is that the noble metal improves the quantum yields by
TiO2 and ZrO2. Supporting a metal oxide on the surface of another oxide accelerating the removal and the transfer of electrons from catalyst to
improves the catalytic activity of the active metal oxide phase due to a molecular oxygen [126]. Therefore, molecular oxygen can trap
gain in surface area and mechanical strength. Supported vanadium photogenerated electrons to form superoxide radicals and conse-
oxides show chemical and electronic properties, which are entirely quently enhance the oxidation of organic molecule such as Methyl
different from those found for unsupported vanadium pentoxide Orange (azo dye) by as much as 30% [127].
(V2O5), in aqueous media and in the solid state. The most important In another study, the photocatalytic degradation of Direct Red 23
oxidation states in aqueous solution are V5+ and V4+. The specific (azo dye) by nano-Ag doped TiO2 has also been investigated [128].
vanadium oxide species that can exist depend on the solution pH and Photodeposition of 1.5% Ag on the surface of TiO2 enhanced its
the vanadium oxide concentration. photoactivity in Direct Red 23 degradation. The higher photoactivity
V5+ is known to undergo the following mechanism [114]: of Ag–TiO2 than TiO2 under UV irradiation was ascribed to the effect of
5þ − 4þ silver deposits acting as electron traps on the Ag–TiO2 surface.
V þ e →V ð61Þ
Likewise, Methyl Red (azo dye) was found to degrade by 99% in the
4þ 5þ •− presence of Ag doped TiO2 as against to 85% in neat TiO2 [129].
V þ O2 ðadsÞ→V þ O2 ð62Þ
Methyl Orange (azo dye) was subjected to photodegradation in
5þ þ 6þ
V þ h →V ð63Þ the presence of Ag/TiO2 [130].The results showed that the final
V

þ OH →V
− 5þ
þ OH

ð64Þ concentration of the dye solution in the presence of Ag/TiO2 powder
was much less than that with TiO2. This was explained on the basis
5þ 4þ •
V þ dyeðadsÞ→V þ dye ð65Þ that Ag on the surface of TiO2 acts as photogenerated electron trap,
which accelerates the rate of electron transfer to molecular oxygen,
3þ 4þ •−
V þ O2 ðadsÞ→V þ O2 ð66Þ and thus inhibits the recombination of photogenerated electrons and
holes [131]. On one hand, photogenerated holes produce •OH radicals
The V4+ traps holes and electron and forms V5+ and V3+ which oxidize the dye molecular adsorbed on the surface of Ag/TiO2 ,
respectively [120]. Subsequently, the trapped electrons and holes while on the other hand, photogenerated electrons accumulate on the
are released and migrate to the surface of TiO2. By accepting an Ag particles by conduction band, and react with the sorbed oxygen
electron, the adsorbed O2 on the surface of TiO2 is reduced to O•2−, molecular to produce free oxygen radicals, such as O•2−, HO•2− , •OH,
while surface hydroxyl group translates into hydroxyl radical (•OH) by etc. [132]. Moreover, the decrease of the band-gap energy contributed
M.A. Rauf et al. / Desalination 276 (2011) 13–27 23

2-hydroxy-1,4- phenol
naphthoquinone O
OH

OH

maleic acid
NaO3S
O
HO2C CO2H
phthalide O benzoic acid
CO2H

O
N
N
2-hydroxy
propanoic acid
OH
2-naphthol
2,5-cyclohexadiene
OH CO2H
-1,4-dione
O HO

Acid Orange 2-formyl-benzoic


acid

O
O

CO2H

Fig. 6. Degradation mechanism proposed on the basis of GC/MS analysis during photo degradation of Acid Orange 7 [66].

by the Ag loaded on the TiO2 induces red shift of the photo-absorption of azo dyes [66,133–137]. A combination of WO3/TiO2 can induce
band, and expands the responding photo-spectra to the longer spectra effective charge separation by trapping photogenerated electrons.
region. Moreover, the addition of WO3 to TiO2 will increase the Lewis surface
acidity of the catalyst. Thus the WO3/TiO2 particles can adsorb more

5.8. Doping with W ions OH or H2O on the surface and generate a greater amount of •OH
radicals. At the same time, WOx/TiO2 particles have a higher
Complementary to metal/TiO2 photocatalysts, TiO2 has been adsorption affinity toward the reactant molecules [133]. In W/Ti
coupled with other semiconductors such as WO3 for the degradation catalysts, the presence of tungsten in the TiO2 lattice produces local

benzene sulfonic N-(Dimethylamino) phenol


acid
N(CH3)2 SO3H N(CH3)2

(CH3)2N NH2

OH

N
N
NH2 NHCH3

HO3S NH2

SO3Na
aniline N-methylaniline
Methyl orange

Fig. 7. Degradation mechanism proposed on the basis of GC/MS analysis during photo degradation of Methyl Orange [66].
24 M.A. Rauf et al. / Desalination 276 (2011) 13–27

levels in the forbidden band inducing the formation of recombination The above oxidation reduction reactions are in competition with
centers for e−–h+ pairs [134]. The recombination occurs according to the redox processes that could occur at the solid/liquid interface and
the following reactions could play a significant role for the lighter doped samples, negatively
− 6þ 5þ
influencing their photoactivity.
eðTiO2Þcb þ WðbulkÞ →WðbulkÞ ð67Þ By increasing the tungsten content, tungsten oxide species with an
þ 5þ 6þ octahedrally coordination could be produced on the particle surfaces,
hðTiO2Þcb þ WðbulkÞ →WðbulkÞ ð68Þ some interface acceptor states mainly formed by tungsten species can

NH2 NH2

N N N N

(M=700.0)
SO3Na SO3Na
OH
NH2 NH2

N N N N + SO42- + Na+

OH

NO2 NO2

N N N N

(detected (M+H) = 553.9)

OH
H NO2 NO2
C +
O N N N N

(detected (M+H) = 331.4)

OH

OH NO2
C
O N N N N

(detected (M+H) = 347.4)

-COO- OH

NO2
HO
N N N N + CO2

OH
NO2
HO + NO2- + NH4+
NO2 O2N NO2 O2N

NO3-, N2
OH

OH

Cn acids CH3COOH + HCOOH CO2 + H2O


HO
(n<6)
OH

Fig. 8. Photodegradation pathway of CR in the presence of the PW11/TiO2 film [137].


M.A. Rauf et al. / Desalination 276 (2011) 13–27 25

be produced on the surface and for higher amounts the so-called [7] A.L. Ahmad, S.W. Puasa, Reactive dyes decolourization from an aqueous solution
by combined coagulation/micellar-enhanced ultrafiltration process, Chemical
isotypic heterojunctions can appear [133]. The presence of interfer- Engineering Journal 132 (2007) 257–265.
ence acceptor states and/or localized heterojunctions allows electron [8] M. Riera-Torres, C. Gutiérrez-Bouzán, M. Crespi, Combination of coagulation–
transfer from the conduction band of TiO2 to W5+(surface) and/or to the flocculation and nanofiltration techniques for dye removal and water reuse in
textile effluents, Desalination 252 (2010) 53–59.
conduction band of WO3 islands. Consequently, the separation of the [9] K. Shakir, A.F. Elkafrawy, H.F. Ghoneimy, S.G. Elrab Beheir, M. Refaat, Removal of
photoproduced pairs would improve and the rate of electron transfer rhodamine B (a basic dye) and thoron (an acidic dye) from dilute aqueous solutions
to O2 (which is used as a bubbling gas) would increase. and wastewater simulants by ion flotation, Water Research 44 (2010) 1449–1461.
[10] S. Zodi, O. Potier, F. Lapicque, J.-P. Leclerc, Treatment of the industrial
Tungsten (4%) doped titania was used to photodegrade Acid wastewaters by electrocoagulation: optimization of coupled electrochemical
Orange 7 and Methyl Orange dyes (azo dyes) and the degradation and sedimentation processes, Desalination 261 (2010) 186–190.
followed the pseudo first-order kinetics with a degradation efficiency [11] M.A. Rauf, S.S. Ashraf, Application of Advanced Oxidation Processes (AOP) to dye
degradation—an overview, in: Arnold R. Lang (Ed.), Dyes and Pigments: New
of 80% in 300 min [66]. The degraded products were analyzed by using
Research, Nova Science Publishers, Inc, 2009.
the liquid chromatography tandem mass spectrometry and degrada- [12] T.M. Elmorsi, Y.M. Riyad, Z.H. Mohamed, H.M.H. Abd El Bary, Decolorization of
tion mechanism was proposed as shown in Figs. 6 and 7. Mordant red 73 azo dye in water using H2O2/UV and photo-Fenton treatment,
Another example of the photocatalytic degradation of an azo dye Journal of Hazardous Materials 174 (2010) 352–358.
[13] S. Gül, Ö. Özcan-YildIrIm, Degradation of Reactive Red 194 and Reactive Yellow
using a tungsten/ titanium oxide composite film was reported by Li et 145 azo dyes by O3 and H2O2/UV-C processes, Chemical Engineering Journal 155
al. [137]. Intermediates and final products generated during the Congo (2009) 684–690.
Red degradation were analyzed by mass spectrometry and ion [14] F.H. AlHamedi, M.A. Rauf, S.S. Ashraf, Degradation studies of Rhodamine B in the
presence of UV/H2O2, Desalination 239 (2009) 159–166.
chromatography. Fig. 8 shows the successive degradative pathway [15] A. Masarwa, S. Rachmilovich-Calis, N. Meyerstein, D. Meyerstein, Oxidation of
as a result of OH radical attack on the diazo dye compound. The organic substrates in aerated aqueous solutions by the Fenton reagent,
degradation events included cleavage of C–S bonds, aromatic ring Coordination Chemistry Reviews 249 (2005) 1937–1943.
[16] E. Chamarro, A. Marco, S. Esplugas, Use of Fenton reagent to improve organic
opening, cleavage of –N=N– double bonds, cleavage of various C–N chemical biodegradability, Water Research 35 (2001) 1047–1051.
and C–C bonds in addition to the decarboxylation. [17] C. Bouasla, M.E.-H. Samar, F. Ismail, Degradation of methyl violet 6B dye by the
Fenton process, Desalination 254 (2010) 35–41.
[18] J.M. Monteagudo, A. Durán, I.S. Martín, M. Aguirre, Catalytic degradation of
Orange II in a ferrioxalate-assisted photo-Fenton process using a combined UV-
6. Conclusion A/C-solar pilot-plant system, Applied Catalysis B: Environmental 95 (2010)
120–129.
Photodegradation of azo dyes using transition metal doped TiO2 has [19] A.K. Abdessalem, N. Bellakhal, N. Oturan, M. Dachraoui, M.A. Oturan, Treatment
of a mixture of three pesticides by photo- and electro-Fenton processes,
been found to be very effective for the remedial of dye-contaminated Desalination 250 (2010) 450–455.
solutions. The basis of the reaction is a photoredox process. An [20] N. Modirshahla, M.A. Behnajady, F. Ghanbary, Decolorization and mineralization
important role in this process is played by molecular oxygen and of C.I. Acid Yellow 23 by Fenton and photo-Fenton processes, Dyes and Pigments
73 (2007) 305–310.
other active species, such as O•2−, HO•2, H2O2, •OH and HO•2−, which are [21] A.R. Tehrani-Bagha, N.M. Mahmoodi, F.M. Menger, Degradation of a persistent
generated in a sequence of reactions. They make the photocatalytic organic dye from colored textile wastewater by ozonation, Desalination 260
processes more efficient resulting in enhanced dye degradation via the (2010) 34–38.
[22] S. Song, H. Ying, Z. He, J. Chen, Mechanism of decolorization and degradation of CI
formation of intermediates such as aromatic amines, phenolic com- Direct Red 23 by ozonation combined with sonolysis, Chemosphere 66 (2007)
pounds and several organic acids. 1782–1788.
Although azo dyes are one of the most commonly used class of [23] H. Ghodbane, O. Hamdaoui, Intensification of sonochemical decolorization of
anthraquinonic dye Acid Blue 25 using carbon tetrachloride, Ultrasonics
dyes in industrial applications, limited literature is available on their
Sonochemistry 16 (2009) 455–461.
remediation from effluents using TiO2 doped with transition metals. [24] S. Merouani, O. Hamdaoui, F. Saoudi, M. Chiha, Sonochemical degradation of
Different papers citing the use of the doped semiconductors seems to Rhodamine B in aqueous phase: effects of additives, Chemical Engineering
have no connectivity between them as to the amount of catalyst, Journal 158 (2010) 550–557.
[25] X. Wang, Z. Yao, J. Wang, W. Guo, G. Li, Degradation of reactive brilliant red in
reactor set up, reaction time and type, concentration of substrates, and aqueous solution by ultrasonic cavitation, Ultrasonics Sonochemistry 15 (2008)
intensity of UV/Vis radiation. It might be critical at this point to 43–48.
standardize the conditions of reactions instead of reporting data on [26] M.A. Rauf, M.A. Meetani, A. Khaleel, A. Ahmed, Photocatalytic degradation of
Methylene Blue using a mixed catalyst and product analysis by LC/MS, Chemical
different azo dyes. Another limitation which was observed during this Engineering Journal 157 (2010) 373–378.
survey was the lack of suggestive degradation mechanisms of various [27] R. Xu, J. Li, J. Wang, X. Wang, B. Liu, B. Wang, X. Luan, X. Zhang, Photocatalytic
azo dyes. This aspect should not be overlooked while reporting any degradation of organic dyes under solar light irradiation combined with Er3+:
YAlO3/Fe- and Co-doped TiO2 coated composites, Solar Energy Materials and
future work. A demonstrated ability to use transition metal doped Solar Cells 94 (2010) 1157–1165.
titanium oxide systems at a pilot scale for effluent purification [28] W. Zhang, J. Zhang, Z. Chen, T. Wang, Photocatalytic degradation of methylene
processes would certainly benefit the environment. blue by ZnGa2O4 thin films, Catalysis Communications 10 (2009) 1781–1785.
[29] M.H. Habibi, N. Talebian, Photocatalytic degradation of an azo dye X6G in water:
a comparative study using nanostructured indium tin oxide and titanium oxide
thin films, Dyes and Pigments 73 (2007) 186–194.
References [30] S.B. Bukallah, M.A. Rauf, S.S. Ashraf, Photocatalytic decoloration of Coomassie
Brilliant Blue with titanium oxide, Dyes and Pigments 72 (2007) 353–356.
[1] S. Parsons, Advanced Oxidation Processes for Water and Wastewater, IWA [31] L. Ayed, K. Chaieb, A. Cheref, A. Bakhrouf, Biodegradation and decolorization of
Publishing, London, UK, 2004. triphenylmethane dyes by Staphylococcus epidermidis, Desalination 260 (2010)
[2] M. Zubair Alam, S. Ahmad, A. Malik, M. Ahmad, Mutagenicity and genotoxicity of 137–146.
tannery effluents used for irrigation at Kanpur, India, Ecotoxicology and [32] X. Zhao, I.R. Hardin, H.-M. Hwang, Biodegradation of a model azo disperse dye by
Environmental Safety 73 (2010) 1620–1628. the white rot fungus Pleurotus ostreatus, International Biodeterioration &
[3] C. O'Neill, A. Lopez, S. Esteves, F.R. Hawkes, D.L. Hawkes, S. Wilcox, Azo-dye Biodegradation 57 (2006) 1–6.
degradation in an anaerobic-aerobic treatment system operating on simulated [33] Ö. ÇInar, S. Yasar, M. Kertmen, K. Demiröz, N.Ö. Yigit, M. Kitis, Effect of cycle time
textile effluent, Applied Microbiology and Biotechnology 53 (1999) 249–254. on biodegradation of azo dye in sequencing batch reactor, Process Safety and
[4] N. Nasuha, B.H. Hameed, A.T.M. Din, Rejected tea as a potential low-cost Environmental Protection 86 (2008) 455–460.
adsorbent for the removal of methylene blue, Journal of Hazardous Materials 175 [34] A. Vahdat, S.H. Bahrami, M. Arami, A. Motahari, Decomposition and decoloration
(2010) 126–132. of a direct dye by electron beam radiation, Radiation Physics and Chemistry 79
[5] M.J. Martin, A. Artola, M.D. Balaguer, M. Rigola, Activated carbons developed (2010) 33–35.
from surplus sewage sludge for the removal of dyes from dilute aqueous [35] K.A. Mohamed, A.A. Basfar, A.A. Al-Shahrani, Gamma-ray induced degradation of
solutions, Chemical Engineering Journal 94 (2003) 231–239. diazinon and atrazine in natural groundwaters, Journal of Hazardous Materials
[6] M.A. Rauf, S.M. Qadri, S. Ashraf, K.M. Al-Mansoori, Adsorption studies of 166 (2009) 810–814.
Toluidine Blue from aqueous solutions onto gypsum, Chemical Engineering [36] Y.-P. Chen, S.-Y. Liu, H.-Q. Yu, H. Yin, Q.-R. Li, Radiation-induced degradation of
Journal 150 (2009) 90–95. methyl orange in aqueous solutions, Chemosphere 72 (2008) 532–536.
26 M.A. Rauf et al. / Desalination 276 (2011) 13–27

[37] K. Dajka, E. Takács, D. Solpan, L. Wojnárovits, O. Güven, High-energy irradiation efficient WOx/TiO2 photocatalyst, Journal of Hazardous Materials 177 (2010)
treatment of aqueous solutions of C.I. Reactive Black 5 azo dye: pulse radiolysis 781–791.
experiments, Radiation Physics and Chemistry 67 (2003) 535–538. [67] M. Pera-Titus, V. García-Molina, M.A. Baños, J. Giménez, S. Esplugas, Degradation
[38] L. Andronic, A. Enesca, C. Vladuta, A. Duta, Photocatalytic activity of cadmium of chlorophenols by means of advanced oxidation processes: a general review,
doped TiO2 films for photocatalytic degradation of dyes, Chemical Engineering Applied Catalysis B: Environmental 47 (2004) 219–256.
Journal 152 (2009) 64–71. [68] M.N. Chong, B. Jin, C.W.K. Chow, C. Saint, Recent developments in photocatalytic
[39] A. Fujishima, X. Zhang, D.A. Tryk, TiO2 photocatalysis and related surface water treatment technology: a review, Water Research 44 (2010) 2997–3027.
phenomena, Surface Science Reports 63 (2008) 515–582. [69] C.-C. Wang, C.-K. Lee, M.-D. Lyu, L.-C. Juang, Photocatalytic degradation of C.I.
[40] M.M. Mohamed, M.M. Al-Esaimi, Characterization, adsorption and photocataly- Basic Violet 10 using TiO2 catalysts supported by Y zeolite: an investigation of
tic activity of vanadium-doped TiO2 and sulfated TiO2 (rutile) catalysts: the effects of operational parameters, Dyes and Pigments 76 (2008) 817–824.
degradation of methylene blue dye, Journal of Molecular Catalysis A: Chemical [70] J. Sun, L. Qiao, S. Sun, G. Wang, Photocatalytic degradation of Orange G on
255 (2006) 53–61. nitrogen-doped TiO2 catalysts under visible light and sunlight irradiation,
[41] C. Chen, Z. Wang, S. Ruan, B. Zou, M. Zhao, F. Wu, Photocatalytic degradation of C.I. Journal of Hazardous Materials 155 (2008) 312–319.
Acid Orange 52 in the presence of Zn-doped TiO2 prepared by a stearic acid gel [71] H. Lachheb, E. Puzenat, A. Houas, M. Ksibi, E. Elaloui, C. Guillard, J.-M. Herrmann,
method, Dyes and Pigments 77 (2008) 204–209. Photocatalytic degradation of various types of dyes (Alizarin S, Crocein Orange G,
[42] Z.M. El-Bahy, A.A. Ismail, R.M. Mohamed, Enhancement of titania by doping rare Methyl Red, Congo Red, Methylene Blue) in water by UV-irradiated titania,
earth for photodegradation of organic dye (Direct Blue), Journal of Hazardous Applied Catalysis B: Environmental 39 (2002) 75–90.
Materials 166 (2009) 138–143. [72] W. Baran, A. Makowski, W. Wardas, The effect of UV radiation absorption of
[43] U. Diebold, The surface science of titanium dioxide, Surface Science Reports 48 cationic and anionic dye solutions on their photocatalytic degradation in the
(2003) 53–229. presence TiO2, Dyes and Pigments 76 (2008) 226–230.
[44] Y.B. Xie, X.Z. Li, Interactive oxidation of photoelectrocatalysis and electro-Fenton [73] V. Augugliaro, C. Baiocchi, A. Bianco Prevot, E. García-López, V. Loddo, S. Malato,
for azo dye degradation using TiO2–Ti mesh and reticulated vitreous carbon G. Marcí, L. Palmisano, M. Pazzi, E. Pramauro, Azo-dyes photocatalytic
electrodes, Materials Chemistry and Physics 95 (2006) 39–50. degradation in aqueous suspension of TiO2 under solar irradiation, Chemosphere
[45] M. Saquib, M. Abu Tariq, M. Faisal, M. Muneer, Photocatalytic degradation of two 49 (2002) 1223–1230.
selected dye derivatives in aqueous suspensions of titanium dioxide, Desalina- [74] K. Soutsas, V. Karayannis, I. Poulios, A. Riga, K. Ntampegliotis, X. Spiliotis, G.
tion 219 (2008) 301–311. Papapolymerou, Decolorization and degradation of reactive azo dyes via
[46] F.H. Abdullah, M.A. Rauf, S.S. Ashraf, Kinetics and optimization of photolytic heterogeneous photocatalytic processes, Desalination 250 (2010) 345–350.
decoloration of carmine by UV/H2O2, Dyes and Pigments 75 (2007) 194–198. [75] M.A. Behnajady, N. Modirshahla, M. Shokri, Photodestruction of Acid Orange 7
[47] M.A. Meetani, S.M. Hisaindee, F. Abdullah, S.S. Ashraf, M.A. Rauf, Liquid (AO7) in aqueous solutions by UV/H2O2: influence of operational parameters,
chromatography tandem mass spectrometry analysis of photodegradation of a Chemosphere 55 (2004) 129–134.
diazo compound: a mechanistic study, Chemosphere 80 (2010) 422–427. [76] M. Qamar, M. Saquib, M. Muneer, Photocatalytic degradation of two selected dye
[48] A. Bianco Prevot, D. Fabbri, E. Pramauro, C. Baiocchi, C. Medana, High- derivatives, chromotrope 2B and amido black 10B, in aqueous suspensions of
performance liquid chromatography coupled to ultraviolet diode array detection titanium dioxide, Dyes and Pigments 65 (2005) 1–9.
and electrospray ionization mass spectrometry for the analysis of intermediates [77] B. Zielinska, J. Grzechulska, B. Grzmil, A.W. Morawski, Photocatalytic degradation
produced in the initial steps of the photocatalytic degradation of sulfonated azo of Reactive Black 5: a comparison between TiO2-Tytanpol A11 and TiO2-Degussa
dyes, Journal of Chromatography. A 1202 (2008) 145–154. P25 photocatalysts, Applied Catalysis B: Environmental 35 (2001) L1–L7.
[49] J.-M. Herrmann, C. Guillard, J. Disdier, C. Lehaut, S. Malato, J. Blanco, New [78] C. Guillard, H. Lachheb, A. Houas, M. Ksibi, E. Elaloui, J.-M. Herrmann, Influence of
industrial titania photocatalysts for the solar detoxification of water containing chemical structure of dyes, of pH and of inorganic salts on their photocatalytic
various pollutants, Applied Catalysis B: Environmental 35 (2002) 281–294. degradation by TiO2 comparison of the efficiency of powder and supported TiO2,
[50] S.G. Huling, R.G. Arnold, R.A. Sierka, P.K. Jones, D.D. Fine, Contaminant adsorption Journal of Photochemistry and Photobiology A: Chemistry 158 (2003) 27–36.
and oxidation via Fenton reaction, Journal of Environmental Engineering 126 [79] N. Daneshvar, D. Salari, A.R. Khataee, Photocatalytic degradation of azo dye acid
(2000) 595–600. red 14 in water: investigation of the effect of operational parameters, Journal of
[51] D. Solpan, O. Güven, E. Takács, L. Wojnárovits, K. Dajka, High-energy irradiation Photochemistry and Photobiology A: Chemistry 157 (2003) 111–116.
treatment of aqueous solutions of azo dyes: steady-state gamma radiolysis [80] J.-M. Herrmann, Photocatalysis fundamentals revisited to avoid several mis-
experiments, Radiation Physics and Chemistry 67 (2003) 531–534. conceptions, Applied Catalysis B: Environmental 99 (2010) 461–468.
[52] C. Galindo, P. Jacques, A. Kalt, Photochemical and photocatalytic degradation of [81] T. Sauer, G. Cesconeto Neto, H.J. José, R.F.P.M. Moreira, Kinetics of photocatalytic
an indigoid dye: a case study of acid blue 74 (AB74), Journal of Photochemistry degradation of reactive dyes in a TiO2 slurry reactor, Journal of Photochemistry
and Photobiology A: Chemistry 141 (2001) 47–56. and Photobiology A: Chemistry 149 (2002) 147–154.
[53] M.A. Rauf, S. Ashraf, S.N. Alhadrami, Photolytic oxidation of Coomassie Brilliant [82] C.M. So, M.Y. Cheng, J.C. Yu, P.K. Wong, Degradation of azo dye Procion Red MX-
Blue with H2O2, Dyes and Pigments 66 (2005) 197–200. 5B by photocatalytic oxidation, Chemosphere 46 (2002) 905–912.
[54] B. Yue, Y. Zhou, J. Xu, Z. Wu, X. Zhang, Y. Zou, S. Jin, Photocatalytic degradation of [83] Y. Li, S. Sun, M. Ma, Y. Ouyang, W. Yan, Kinetic study and model of the
aqueous 4-chlorophenol by silica-immobilized polyoxometalates, Environmen- photocatalytic degradation of rhodamine B (RhB) by a TiO2-coated activated
tal Science & Technology 36 (2002) 1325–1329. carbon catalyst: effects of initial RhB content, light intensity and TiO2 content in
[55] C.-H. Wu, H.-W. Chang, J.-M. Chern, Basic dye decomposition kinetics in a the catalyst, Chemical Engineering Journal 142 (2008) 147–155.
photocatalytic slurry reactor, Journal of Hazardous Materials 137 (2006) 336–343. [84] M. Asiltürk, F. SayIlkan, E. Arpaç, Effect of Fe3+ ion doping to TiO2 on the
[56] J. Saien, A.R. Soleymani, Degradation and mineralization of Direct Blue 71 in a photocatalytic degradation of Malachite Green dye under UV and vis-irradiation,
circulating upflow reactor by UV/TiO2 process and employing a new method in Journal of Photochemistry and Photobiology A: Chemistry 203 (2009) 64–71.
kinetic study, Journal of Hazardous Materials 144 (2007) 506–512. [85] F. SayIlkan, M. Asiltürk, P. Tatar, N. Kiraz, S. Sener, E. Arpaç, H. SayIlkan,
[57] S. Mozia, Photocatalytic membrane reactors (PMRs) in water and wastewater Photocatalytic performance of Sn-doped TiO2 nanostructured thin films for
treatment. A review, Separation and Purification Technology 73 (2010) 71–91. photocatalytic degradation of malachite green dye under UV and VIS-lights,
[58] A. Danion, J. Disdier, C. Guillard, N. Jaffrezic-Renault, Malic acid photocatalytic Materials Research Bulletin 43 (2008) 127–134.
degradation using a TiO2-coated optical fiber reactor, Journal of Photochemistry [86] M. Bettinelli, V. Dallacasa, D. Falcomer, P. Fornasiero, V. Gombac, T. Montini, L.
and Photobiology A: Chemistry 190 (2007) 135–140. Romanò, A. Speghini, Photocatalytic activity of TiO2 doped with boron and
[59] N.M. Mahmoodi, M. Arami, N.Y. Limaee, N.S. Tabrizi, Kinetics of heterogeneous vanadium, Journal of Hazardous Materials 146 (2007) 529–534.
photocatalytic degradation of reactive dyes in an immobilized TiO2 photo- [87] P. Bouras, E. Stathatos, P. Lianos, Pure versus metal-ion-doped nanocrystalline
catalytic reactor, Journal of Colloid and Interface Science 295 (2006) 159–164. titania for photocatalysis, Applied Catalysis B: Environmental 73 (2007) 51–59.
[60] E. Forgacs, T. Cserháti, G. Oros, Removal of synthetic dyes from wastewaters: a [88] T.K. Ghorai, D. Dhak, S.K. Biswas, S. Dalai, P. Pramanik, Photocatalytic oxidation of
review, Environment International 30 (2004) 953–971. organic dyes by nano-sized metal molybdate incorporated titanium dioxide
[61] A. Fujishima, T.N. Rao, D.A. Tryk, Titanium dioxide photocatalysis, Journal of (MxMoxTi1-xO6) (M = Ni, Cu, Zn) photocatalysts, Journal of Molecular Catalysis
Photochemistry and Photobiology C: Photochemistry Reviews 1 (2000) 1–21. A: Chemical 273 (2007) 224–229.
[62] M. Stylidi, D.I. Kondarides, X.E. Verykios, Pathways of solar light-induced [89] X. Yang, F. Ma, K. Li, Y. Guo, J. Hu, W. Li, M. Huo, Y. Guo, Mixed phase titania
photocatalytic degradation of azo dyes in aqueous TiO2 suspensions, Applied nanocomposite codoped with metallic silver and vanadium oxide: new efficient
Catalysis B: Environmental 40 (2003) 271–286. photocatalyst for dye degradation, Journal of Hazardous Materials 175 (2010)
[63] I.K. Konstantinou, T.A. Albanis, TiO2-assisted photocatalytic degradation of azo 429–438.
dyes in aqueous solution: kinetic and mechanistic investigations: a review, [90] S. Rehman, R. Ullah, A.M. Butt, N.D. Gohar, Strategies of making TiO2 and ZnO
Applied Catalysis B: Environmental 49 (2004) 1–14. visible light active, Journal of Hazardous Materials 170 (2009) 560–569.
[64] A.E.H. Machado, J.A. de Miranda, R.F. de Freitas, E.T.F.M. Duarte, L.F. Ferreira, Y.D.T. [91] L.G. Devi, N. Kottam, B.N. Murthy, S.G. Kumar, Enhanced photocatalytic activity
Albuquerque, R. Ruggiero, C. Sattler, L. de Oliveira, Destruction of the organic of transition metal Mn2+, Ni2+ and Zn2+ doped polycrystalline titania for the
matter present in effluent from a cellulose and paper industry using photo- degradation of Aniline Blue under UV/solar light, Journal of Molecular Catalysis
catalysis, Journal of Photochemistry and Photobiology A: Chemistry 155 (2003) A: Chemical 328 (2010) 44–52.
231–241. [92] S. Liu, J.-H. Yang, J.-H. Choy, Microporous SiO2–TiO2 nanosols pillared
[65] C.G. da Silva, J.L. Faria, Photochemical and photocatalytic degradation of an azo montmorillonite for photocatalytic decomposition of methyl orange, Journal of
dye in aqueous solution by UV irradiation, Journal of Photochemistry and Photochemistry and Photobiology A: Chemistry 179 (2006) 75–80.
Photobiology A: Chemistry 155 (2003) 133–143. [93] M.A. Rauf, S.B. Bukallah, A. Hamadi, A. Sulaiman, F. Hammadi, The effect of
[66] A.K.L. Sajjad, S. Shamaila, B. Tian, F. Chen, J. Zhang, Comparative studies of operational parameters on the photoinduced decoloration of dyes using a hybrid
operational parameters of degradation of azo dyes in visible light by highly catalyst V2O5–TiO2, Chemical Engineering Journal 129 (2007) 167–172.
M.A. Rauf et al. / Desalination 276 (2011) 13–27 27

[94] F. Han, V.S.R. Kambala, M. Srinivasan, D. Rajarathnam, R. Naidu, Tailored titanium [116] Y. Jiang, Y. Sun, H. Liu, F. Zhu, H. Yin, Solar photocatalytic decolorization of C.I.
dioxide photocatalysts for the degradation of organic dyes in wastewater Basic Blue 41 in an aqueous suspension of TiO2–ZnO, Dyes and Pigments 78
treatment: a review, Applied Catalysis A: General 359 (2009) 25–40. (2008) 77–83.
[95] K. Wilke, H.D. Breuer, The influence of transition metal doping on the physical [117] S.K. Asl, S.K. Sadrnezhaad, M.K. rad, The seeding effect on the microstructure and
and photocatalytic properties of titania, Journal of Photochemistry and photocatalytic properties of ZnO nano powders, Materials Letters 64 (2010)
Photobiology A: Chemistry 121 (1999) 49–53. 1935–1938.
[96] P. Ciesla, P. Kocot, P. Mytych, Z. Stasicka, Homogeneous photocatalysis by [118] D.L. Liao, C.A. Badour, B.Q. Liao, Preparation of nanosized TiO2/ZnO composite
transition metal complexes in the environment, Journal of Molecular Catalysis A: catalyst and its photocatalytic activity for degradation of methyl orange, Journal
Chemical 224 (2004) 17–33. of Photochemistry and Photobiology A: Chemistry 194 (2008) 11–19.
[97] P. Ciesla, A. Karocki, Z. Stasicka, Photoredox behaviour of the Cr–EDTA complex [119] W.A. Sadik, Decolourization of an azo dye by heterogeneous photocatalysis,
and its environmental aspects, Journal of Photochemistry and Photobiology A: Process Safety and Environmental Protection 85 (2007) 515–520.
Chemistry 162 (2004) 537–544. [120] B. Tian, C. Li, F. Gu, H. Jiang, Y. Hu, J. Zhang, Flame sprayed V-doped TiO2
[98] M. Al-Shamisi, M.Sc Thesis, Environmental Science, United Arab Emirates nanoparticles with enhanced photocatalytic activity under visible light irradi-
University, Al-Ain, 2009. ation, Chemical Engineering Journal 151 (2009) 220–227.
[99] M.A. Meetani, M.A. Rauf, S. Hisaindee, A.A. Khaleel, A. AlZamly, A. Ahmed, [121] H. Li, G. Zhao, Z. Chen, G. Han, B. Song, Low temperature synthesis of visible light-
Mechanistic studies of photoinduced degradation of Orange G using LC/MS, driven vanadium doped titania photocatalyst, Journal of Colloid and Interface
unpublished results. Science 344 (2010) 247–250.
[100] M. Karkmaz, E. Puzenat, C. Guillard, J.M. Herrmann, Photocatalytic degradation of [122] J. Xu, Y. Ao, M. Chen, D. Fu, C. Yuan, Photocatalytic activity of vanadium-doped
the alimentary azo dye amaranth: mineralization of the azo group to nitrogen, titania-activated carbon composite film under visible light, Thin Solid Films 518
Applied Catalysis B: Environmental 51 (2004) 183–194. (2010) 4170–4174.
[101] E.B. Gracien, J. Shen, X. Sun, D. Liu, M. Li, S. Yao, J. Sun, Photocatalytic activity of [123] J.C.S. Wu, C.-H. Chen, A visible-light response vanadium-doped titania
manganese, chromium and cobalt-doped anatase titanium dioxide nanoporous nanocatalyst by sol-gel method, Journal of Photochemistry and Photobiology
electrodes produced by re-anodization method, Thin Solid Films 515 (2007) A: Chemistry 163 (2004) 509–515.
5287–5297. [124] D. Masih, H. Yoshitake, Y. Izumi, Photo-oxidation of ethanol on mesoporous
[102] R. Glazewski, G.M. Morrison, Copper(I)/copper(II) reactions in an urban river, vanadium–titanium oxide catalysts and the relation to vanadium(IV) and (V)
Science of The Total Environment 189–190 (1996) 327–333. sites, Applied Catalysis A: General 325 (2007) 276–282.
[103] D. Chen, A.K. Ray, Removal of toxic metal ions from wastewater by [125] C. He, Y. Yu, X. Hu, A. Larbot, Influence of silver doping on the photocatalytic
semiconductor photocatalysis, Chemical Engineering Science 56 (2001) activity of titania films, Applied Surface Science 200 (2002) 239–247.
1561–1570. [126] H. Gerischer, A. Heller, Photocatalytic oxidation of organic molecules at TiO2
[104] K.V.S. Rao, B. Lavédrine, P. Boule, Influence of metallic species on TiO2 for the particles by sunlight in aerated water, 1992.
photocatalytic degradation of dyes and dye intermediates, Journal of Photo- [127] A. Alem, H. Sarpoolaky, The effect of silver doping on photocatalytic properties of
chemistry and Photobiology A: Chemistry 154 (2003) 189–193. titania multilayer membranes, Solid State Sciences 12 (2010) 1469–1472.
[105] R.S.K. Wong, J. Feng, X. Hu, P.L. Yue, Discoloration and mineralization of non- [128] N. Sobana, K. Selvam, M. Swaminathan, Optimization of photocatalytic
biodegradable azo dye Orange II by copper-doped TiO2 nanocatalysts, Journal of degradation conditions of Direct Red 23 using nano-Ag doped TiO2, Separation
Environmental Science and Health A 39 (2004) 2583–2595. and Purification Technology 62 (2008) 648–653.
[106] N. Helaïli, Y. Bessekhouad, A. Bouguelia, M. Trari, Visible light degradation of [129] C. Sahoo, A.K. Gupta, A. Pal, Photocatalytic degradation of Methyl Red dye in
Orange II using xCuyO2/TiO2 heterojunctions, Journal of Hazardous Materials 168 aqueous solutions under UV irradiation using Ag+doped TiO2, Desalination 181
(2009) 484–492. (2005) 91–100.
[107] T. Tong, J. Zhang, B. Tian, F. Chen, D. He, Preparation of Fe3+-doped TiO2 [130] H. Wang, J. Niu, X. Long, Y. He, Sonophotocatalytic degradation of methyl orange
catalysts by controlled hydrolysis of titanium alkoxide and study on their by nano-sized Ag/TiO2 particles in aqueous solutions, Ultrasonics Sonochemistry
photocatalytic activity for methyl orange degradation, Journal of Hazardous 15 (2008) 386–392.
Materials 155 (2008) 572–579. [131] P.D. Cozzoli, E. Fanizza, R. Comparelli, M.L. Curri, A. Agostiano, D. Laub, Role of
[108] X.-H. Qi, Z.-H. Wang, Y.-Y. Zhuang, Y. Yu, J.-l. Li, Study on the photocatalysis metal nanoparticles in TiO2/Ag nanocomposite-based microheterogeneous
performance and degradation kinetics of X-3B over modified titanium dioxide, photocatalysis, The Journal of Physical Chemistry. B 108 (2004) 9623–9630.
Journal of Hazardous Materials 118 (2005) 219–225. [132] J. Yu, J. Xiong, B. Cheng, S. Liu, Fabrication and characterization of Ag–TiO2
[109] S.D. Sharma, K.K. Saini, C. Kant, C.P. Sharma, S.C. Jain, Photodegradation of dye multiphase nanocomposite thin films with enhanced photocatalytic activity,
pollutant under UV light by nano-catalyst doped titania thin films, Applied Applied Catalysis B: Environmental 60 (2005) 211–221.
Catalysis B: Environmental 84 (2008) 233–240. [133] A.K.L. Sajjad, S. Shamaila, B. Tian, F. Chen, J. Zhang, One step activation of WOx/
[110] C.-y. Wang, C. Bottcher, D.W. Bahnemann, J.K. Dohrmann, A comparative study of TiO2 nanocomposites with enhanced photocatalytic activity, Applied Catalysis B:
nanometer sized Fe(III)-doped TiO2 photocatalysts: synthesis, characterization Environmental 91 (2009) 397–405.
and activity, Journal of Materials Chemistry 13 (2003) 2322–2329. [134] M.A. Saepurahman, F.K. Abdullah, Chong, Preparation and characterization of
[111] L.G. Devi, S.G. Kumar, B.N. Murthy, N. Kottam, Influence of Mn2+ and Mo6+ tungsten-loaded titanium dioxide photocatalyst for enhanced dye degradation,
dopants on the phase transformations of TiO2 lattice and its photo catalytic Journal of Hazardous Materials 176 (2010) 451–458.
activity under solar illumination, Catalysis Communications 10 (2009) 794–798. [135] H. Song, H. Jiang, X. Liu, G. Meng, Efficient degradation of organic pollutant with
[112] Y. Mu, H.-Q. Yu, J.-C. Zheng, S.-J. Zhang, TiO2-mediated photocatalytic WOx modified nano TiO2 under visible irradiation, Journal of Photochemistry
degradation of Orange II with the presence of Mn2+ in solution, Journal of and Photobiology A: Chemistry 181 (2006) 421–428.
Photochemistry and Photobiology A: Chemistry 163 (2004) 311–316. [136] T. Hathway, E.M. Rockafellow, Y.-C. Oh, W.S. Jenks, Photocatalytic degradation
[113] Y. Xu, B. Lei, L. Guo, W. Zhou, Y. Liu, Preparation, characterization and using tungsten-modified TiO2 and visible light: kinetic and mechanistic effects
photocatalytic activity of manganese doped TiO2 immobilized on silica gel, using multiple catalyst doping strategies, Journal of Photochemistry and
Journal of Hazardous Materials 160 (2008) 78–82. Photobiology A: Chemistry 207 (2009) 197–203.
[114] L.G. Devi, B.N. Murthy, S.G. Kumar, Photocatalytic activity of TiO2 doped with [137] D. Li, Y. Guo, C. Hu, C. Jiang, E. Wang, Preparation, characterization and
Zn2+ and V5+ transition metal ions: Influence of crystallite size and dopant photocatalytic property of the PW11O397-/TiO2 composite film towards azo-dye
electronic configuration on photocatalytic activity, Materials Science and degradation, Journal of Molecular Catalysis A: Chemical 207 (2004) 183–193.
Engineering: B 166 (2010) 1–6.
[115] V. Sukharev, R. Kershaw, Concerning the role of oxygen in photocatalytic
decomposition of salicylic acid in water, Journal of Photochemistry and
Photobiology A: Chemistry 98 (1996) 165–169.

You might also like