Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Ahmed2010 R

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Desalination 261 (2010) 3–18

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

Heterogeneous photocatalytic degradation of phenols in wastewater: A review on


current status and developments
Saber Ahmed a,⁎, M.G. Rasul a, Wayde N. Martens b, R. Brown c, M.A. Hashib d
a
Faculty of Science, Engineering and Health, CQ University, QLD 4702, Australia
b
Queensland University of Technology, Discipline of Chemistry, Faculty of Science and Technology, Australia
c
Queensland University of Technology, School of Engineering Systems, Australia
d
Department of Ecological Engineering, Toyohashi University of Technology, Japan

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

Article history: In recent years, there has been an enormous amount of research and development in the area of
Received 18 December 2009 heterogeneous photocatalytic water purification process due to its effectiveness in degrading and
Received in revised form 30 March 2010 mineralising the recalcitrant organic compounds as well as the possibility of utilising the solar UV and
Accepted 29 April 2010
visible spectrum. One hundred and twenty recently published papers are reviewed and summarised here
Available online 9 June 2010
with the focus being on the photocatalytic oxidation of phenols and their derivatives, predominant in waste
water effluent. In this review, the effects of various operating parameters on the photocatalytic degradation
Keywords:
Phenols
of phenols and substituted phenols are presented. Recent findings suggested that different parameters, such
Substituted phenols as type of photocatalyst and composition, light intensity, initial substrate concentration, amount of catalyst,
Photocatalysis pH of the reaction medium, ionic components in water, solvent types, oxidising agents/electron acceptors,
Water purification mode of catalyst application, and calcination temperatures can play an important role on the photocatalytic
degradation of phenolic compounds in wastewater. Extensive research has focused on the enhancement of
photocatalysis by modification of TiO2 employing metal, non-metal and ion doping. Recent developments in
TiO2 photocatalysis for the degradation of various phenols and substituted phenols are also reviewed.
© 2010 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3
2. Photocatalytic oxidation process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1. Principle of photocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2. Phenol, chlorophenol and nitrophenol in photocatalytic medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3. Influence of parameters on the photocatalytic degradation of phenols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1. Comparison of catalyst and substituted group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2. Light intensity and wavelength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.3. Initial concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.4. Catalyst loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.5. Medium pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.6. Co-occurring substances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.7. Oxidants/electron acceptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.8. Calcination temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.9. Doping and mixed semiconductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4. Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

⁎ Corresponding author. Fax: +61 7 4930 9382.


E-mail address: s.ahmed@cqu.edu.au (S. Ahmed).

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.04.062
4 S. Ahmed et al. / Desalination 261 (2010) 3–18

1. Introduction catalyst deactivation [18]. Information from various investigations


suggests that the photocatalytic degradation of phenol and its
Recycling of wastewater effluent is recognised to be a strategic derivatives is mainly dependent on the solution pH, catalysts and
approach in a sustainable water management portfolio by water their composition, organic substrate type and concentration, light
utilities both in Australia and in other parts of the world to minimise intensity, catalyst loading, ionic composition of waste water, types
the growing water demand in a water-scarce environment [1–3]. of solvent, oxidant concentration, and calcination temperatures.
The widespread occurrence of phenols in waste water and associated Understanding the impacts of various parameters on the photo-
environmental hazards has heightened concern over public health catalytic degradation efficiency is of paramount importance from
[4]. Phenols and their derivatives are well known for their bio- the design and operational points of view when choosing a
recalcitrant and acute toxicity. Phenols are being introduced sustainable, efficient technique for the treatment of waste water.
continuously into the aquatic environment through various anthro- This paper aims to review and summarise the role of important
pogenic inputs. The presence of toxic organic compounds in storm operating parameters on the photocatalytic degradation of phenols
and waste water effluent is reported to be a major impediment to the in wastewater together with recent achievements. Recent research
widespread acceptance of water recycling [5,6]. Further, their on the modified TiO2 photocatalysis under visible and/or solar light
variety, toxicity and persistence can directly impact the health of aimed at improving phenols degradation in wastewater by means of
ecosystems and present a threat to humans through contamination metal, non-metal and ion doping is also highlighted in this review.
of drinking water supplies e.g., surface and ground water [3,4]. In The existing limitation and future research needs associated with
response it has become a challenge to achieve the effective removal the treatment technology pertaining to the contaminant of interest
of persistent organic pollutants from waste water effluent to are also discussed.
minimise the risk of pollution problems from such toxic chemicals
and to enable its reuse. Consequently, considerable efforts have been
devoted to developing a suitable purification method that can easily 2. Photocatalytic oxidation process
destroy these bio-recalcitrant organic contaminants. Due to their
incomplete removal during wastewater treatment, they are ubiqui- 2.1. Principle of photocatalysis
tous in secondary wastewater effluents, rivers and lakes at low
concentration. Despite their low concentration, these contaminants In the photocatalytic oxidation process, organic pollutants are
are a major health concern because of their extremely high destroyed in the presence of semiconductor photocatalysts (e.g., TiO2
endocrine disrupting potency and genotoxicity [7]. These findings and ZnO), an energetic light source, and an oxidising agent such as
enunciate the necessity for further research on the removal of trace oxygen or air. Fig. 1 illustrates the mechanism of TiO2 photocatalysis.
contaminants to minimise their accumulation, particularly prior to Only photons with energies greater than the band-gap energy (ΔE)
indirect or direct reuse of reclaimed water. Conventional wastewater can result in the excitation of valence band (VB) electrons which then
purification systems such as activated carbon adsorption, membrane promote the possible reactions. The absorption of photons with
filtration, chemical coagulation, ion exchange on synthetic adsor- energy lower than ΔE or longer wavelengths usually causes energy
bent resins, etc., also generate wastes during the treatment of dissipation in the forms of heat. The illumination of the photocatalytic
contaminated water, which requires additional steps and cost. In surface with sufficient energy leads to the formation of a positive hole

recent years, the heterogeneous photocatalytic oxidation (HPO) (h+
v ) in the valence band and an electron (e ) in the conduction band
process employing TiO2 and UV light has emerged as a promising (CB). The positive hole oxidizes either pollutant directly or water to
new route for the degradation of persistent organic pollutants, and produce ·OH radicals, whereas the electron in the conduction band
produces more biologically degradable and less toxic substances reduces the oxygen adsorbed on the catalyst (TiO2).
[8,9]. 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 ones into relatively innocuous end products
such as CO2 and H2O. In the HPO process, destruction of recalcitrant
organics is governed by the combined actions of a semiconductor
photocatalyst, an energetic radiation source and an oxidizing agent.
Moreover, the process can be driven by solar UV or visible light. Near
the earth's surface, the sun produces 0.2–0.3 mol photons m− 2 h− 1
in the range of 300–400 nm with a typical UV flux of 20–30 W m2.
This suggests using sunlight as an economically and ecologically
sensible light source [10]. As a result, development of an efficient
photocatalytic water purification process for large scale applications
has received considerable attention.
Although heterogeneous photocatalysis appears in numerous
forms, the photocatalytic degradation of phenol and substituted
phenols in wastewater has recently been most widely investigated.
Several reviews [11,12] focusing briefly on the available technol-
ogies for the abatement of phenolic compounds in wastewater have
recently been published. In light of the basic and applied research
reviewed, the photocatalytic oxidation method appears to be a
promising route for the treatment of wastewater contaminated
with phenol and phenolic compounds. However, the major draw-
backs still encountered are low quantum efficiency due to
inefficient visible light harvesting [13], the design of the photo-
reactor [14], the recovery and reuse of titanium dioxide [15], the
generation of toxic intermediates [16,17], and concern about Fig. 1. Schematic diagram illustrating the principle of TiO2 photocatalysis.
S. Ahmed et al. / Desalination 261 (2010) 3–18 5

The activation of TiO2 by UV light can be represented by the converted to acetylene, maleic acid, carbon monoxide and carbon
following steps: dioxide [20].
Chlorophenols are water pollutants of moderately toxicity and are
− þ
TiO2 þ hν →e þ h ð2:1Þ suspected of carcinogenic properties. They originate from the natural
degradation of chlorinated herbicides, the chlorination of phenolic
− · substances and disinfection of water. Nitrophenols are water
e þ O2 →O2 − ð2:2Þ
pollutants of high toxicity and they are released into the water
In this reaction, h+ and e− are powerful oxidizing and reductive environment due to the synthesis and use of organo-phosphorous
agents respectively. The oxidative and reductive reaction steps are pesticides, azo dyes, and some medical goods. The main byproducts
expressed as: detected during their photocatalytic degradation are 4-nitrocatechol,
Oxidative reaction: benzoquinone, hydroquinone and a number of organic acids. Fig. 2
illustrates the photocatalytic degradation of phenol and the formation
þ
h þ Organic→CO2 ð2:3Þ of intermediates. The •OH radical attacks the phenyl ring of the phenol
yielding catechol, resorcinol, 1,2,3-benzenetriol and hydroquinone,
þ · þ then the phenyl rings in these compounds disintegrate to give
h þ H2 O→ OH þ H ð2:4Þ
malonic acid, then short-chain organic acids such as, maleic, oxalic,
acetic, formic 3-hydroxy propyl carboxylic acid, 2-hydroxy propanal,
Reductive reaction: 2-hydroxy-ethanoic acid glycol acid, finally CO2 and H2O. H• produced
· during the attack of bonds by •OH is reported to be an important
OH þ Organic→CO2 ð2:5Þ
active free radical in the degradation process. During the process, H+
Hydroxyl radical generation by the photocatalytic oxidation or H• is scavenged by oxygen to form HO2• radicals, which eventually
process is shown in the above steps. In the degradation of organic convert to •OH radicals. Therefore, the principal reaction leading to
pollutants, the hydroxyl radical, which is generated from the organics decomposition would be the one with •OH radicals.
oxidation of adsorbed water where it is adsorbed as OH−, is the
primary oxidant; and the presence of oxygen prevents the recombi- 3. Influence of parameters on the photocatalytic degradation
nation of an electron–hole pair. In the photocatalytic degradation of of phenols
pollutants, when the reduction process of oxygen and the oxidation of
pollutants do not advance simultaneously, there is an electron 3.1. Comparison of catalyst and substituted group
accumulation in the CB, thereby causing an increase in the rate of
recombination of e− and h+. Thus it is of paramount importance to The photocatalytic activity of TiO2 is dependent on surface and
prevent electron accumulation in efficient photocatalytic oxidation. In structural properties of the semiconductor such as crystal composi-
photocatalysis, TiO2 has been studied extensively because of its high tion, surface area, particle size distribution, porosity, band gap and
activity, desirable physical and chemical properties, low cost, and surface hydroxyl density. Particle size is of primary importance in
availability. Of three common TiO2 crystalline forms, anatase and heterogeneous catalysis, because it is directly related to the efficiency
rutile forms have been investigated extensively as photocatalysts. of a catalyst through the definition of its specific surface area. A
Anatase has been reported to be more active as a photocatalyst than number of commercially available catalysts have been investigated for
rutile. Different light sources such as UV lamps and solar radiation the photocatalytic degradation of phenolic compounds and dyes in
have been used in previous investigations into the photocatalysis of aqueous environments. Table 2 shows the specification and char-
various pesticide and herbicide derivatives dominant in storm water acteristics of some commercial TiO2 samples. The photocatalyst
and wastewater effluent. titanium dioxide Degussa P-25 has been widely used in most of the
experimental conditions; other catalyst powders, namely, Hombikat
UV100, PC 500 and TTP were also used for degradation of toxic organic
2.2. Phenol, chlorophenol and nitrophenol in photocatalytic medium compounds. P-25 contains 75% anatase and 25% rutile with a specific
BET surface area of 50m2/g and a primary particle size of 20 nm [25].
Phenol along with its chlorophenol and nitrophenol derivatives Hombikat UV 100 consists of 100% pure and smaller anatase with a
is highly soluble in water. The solubility of phenol, o-chlorophenol, specific BET surface area of 250 m2/g and a primary particle size of
m-chlorophenol, p-chlorophenol, o-nitrophenol, m-nitrophenol and 5 nm [25]. The photocatalyst PC 500 has a BET surface area of 287 m2/
p-nitrophenol is shown in Table 1. The main reaction site for their g with 100% anatase and a primary particle size of 5–10 nm [25], and
destruction during photocatalysis is the bulk liquid, where the attack TiO2 obtained from TTP, India has a BET surface area of 9.82 m2/g. It
of hydroxyl radicals on the ring carbons results in various oxidation has been demonstrated that the degradation rate of dyes proceeds
intermediates. Hydroquinone, catechol, and p-benzoquinone are much more rapidly in the presence of Degussa P-25 as compared to
reported to be the major intermediates formed during the photo- other photocatalysts. The efficiency of photocatalysts was shown to
catalytic degradation of phenol [19]. Reaction intermediates such as follow the order: P25 N UV100 N PC500 N TTP for the degradation of
chloro-hydroquinone, 4-chlorocatechol and resorcinol are eventually various dyes [26–32]. The photocatalytic activity of Degussa P25 was
reported to be higher due to slow recombination between electrons
and holes whereas Hombikat UV 100 is reported to have a high
Table 1 photoreactivity due to fast interfacial electron transfer rate. The
Solubility of phenol and its chloro and nitro-derivatives in water. differences in the photocatalytic activity are related to the variations
Compound Solubility (g/100 g) Ref.
in the BET surface, impurities, existence of structural defects in the
crystalline framework or density of hydroxyl groups on the catalyst
Phenol 9.48 [21]
surface.
o-chlorophenol 2.04 [22]
m-chlorophenol 2.25 [22] These factors could influence the adsorption behaviour of a
p-chlorophenol 2.77 [22] pollutant or degradation intermediates and the life time and
o-nitrophenol 0.21 [23] recombination rate of electron–hole pairs [27]. Some of the illustra-
m-nitrophenol 2.19 [23]
tive works in recent years is depicted in Table 3. Priya and Madras
p-nitrophenol 1.34 [23]
[33] compared the photocatalytic degradation of nitrophenols using
6 S. Ahmed et al. / Desalination 261 (2010) 3–18

Fig. 2. Photocatalytic degradation of phenol by nanomaterial TiO2 in wastewater [24].

combustion synthesised nano-TiO2 and Degussa P-25. The photo- on the adsorption over catalyst, the electronic nature and number of
degradation kinetics was first order. The photocatalytic degradation the substituents and their position in the aromatic ring. Lachheb et al.
rates were considerably higher in combustion synthesized TiO2 [38] reported that in comparison to PC 500, P-25 was more efficient
compared to that of P-25. For both catalysts, the degradation rate for the degradation of phenols and poly nitrophenols (4-NP, 2,4-DNP
was shown to follow the order 4-Nitrophenol N 2-Nitrophenol N and 2,4,6-TNP) in the presence of either artificial or solar light. The
3-Nitrophenol N 2,4-Ditrophenol. The position of substitution is degradation followed first order kinetics. The photocatalytic degrada-
reported to affect the rate of degradation. Shukla et al. [34] tions of the tested compounds were shown to be in the following
compared the photocatalytic degradation efficiency of phenol, order: 2,4,6-TNPN 2,4-DNPN 4-NPN Phenol. However, for PC 500 sup-
dichlorophenol (DP) and trichlorophenol (TCP) at 0.266 mM ported on Ahlstrom paper 1048, the order is the opposite — PhenolN
solution concentration in the presence of UV–vis/ZnO/persulfate.
Degradation of the tested compounds under simulated solar light
Table 3
alone was demonstrated to be negligible due to poor absorbance of Some representative work on the photocatalytic degradation of organic contaminants.
light in the available wavelength. In comparison to phenol,
chlorophenols degradation was significantly lower under the con- Pollutant Experimental Degradation rate Ref.
conditions
ditions examined. The order of the degradation rate was found to
be as phenol N DCP N TCP. The photocatalytic activity of ZnO was Phenol C0 phenol = 50ppm, 92% degradation [35]
TiO2 = 2g/l, under UV achieved in 6 h
observed to be higher than TiO2 for the degradation of phenolic
irradiation
contaminants under UV–vis light. Salah et al. [35] compared the (λmax = 366 nm)
efficiency of three commercial catalysts Degussa P25 (20 nm), Tri-substituted TiO2 = 0.5 g/l in 200 ml, 100―56% TCP removal [37]
TiO2-A1 (160 nm) and TiO2-A2 (330 nm) available in anatase form, phenol; TCP, TBP, TMP Co = 0.−0.5 mM obtained in 1 h. 70―44%
and ZnO for the degradation of phenol. After 5 h reactions, the TBP removal achieved in
1 h. 60―22% TMP
order of the efficiency was shown to follow ZnO N P25 N TiO2-A1 N removal in 1 h.
TiO2-A2. The observed variation in the efficiency is related to the Phenol, DNP, 4-NP, TNP Co = 30 mg/l, I = 125W Complete degradation of [38]
structure, diameter of particles and electrical properties of the MP Hg, λ N 290 nm, all nitro phenols occurs
photocatalyst. Peiro et al. [36] observed that the photocatalytic TiO2 = 0.5 g/l, between 2 and 7 h
Phenol, DCP, TCP C0Phenol = 25ppm, 85% degradation of TNP [34]
degradation of phenol and ortho-substituted phenolic compounds
C0DCP = 44 ppm occurs in 150 min. 90%
were shown to follow the order: Guaiacol N 2-chlorophenol = Phenol N C0TCP = 52 ppm, degradation of DNP
Catechol. Kusvuran et al. [37] studied the photocatalytic degradation ZnO = 0.4 g/l, occurs in 150 min and
of 2,4,6-trimethylphenol (TMP), 2,4,6-trichlorophenol (TCP), 2,4,6- Persulfate = 2 g/l 93% degradation of
tribromophenol (TBP), 2,4-dimethylphenol (DMP), 2,4-dichlorophenol phenol occurs in
150 min.
(DCP) and 2,4-dibromophenol (DBP). Under the conditions tested,
Chrysoidine R 1 , Acid C0 = 0.22–0.25 mM, 92% degradation and [26]
degradation of these tri-substituted phenols was found to be in the Red 292 125 W Mp Hg, 77% mineralisation of
following order: TCP N TBP N TMP and the degradation of di-substituted TiO2 = 1 g/l, pH= 4.3– dye1 takes place after
phenols was shown to be in the order: DCP N DBPN DMP. The degradation 5.2 1 h irradiation. 98%
degradation and 88%
rate of the tested compounds was observed to be strongly dependent
mineralisation of dye2
occurs in 2 h
Amaranth1, Bismarck2 C0 = 0.4–0.5 mM, Nearly complete [26]
125 W Mp Hg, degradation occurs in
TiO2 = 1 g/l, 75 min for dye1 and
pH = 3.25–5.68 40 min for dye2.
Table 2 Acid blue 451, Xelenol C0 = 0.3–0.4 mM, About 51 and 73% [30]
Specification and characteristics of commercial TiO2 samples. orange2 125 W Mp Hg, degradation of dye1 and
TiO2 = 1 g/l, pH = not dye2 takes place in
TiO2 sample BET surface area Particle size (nm) Composition
available 80 min irradiation time.
(m2/g)
Bromothymol blue C0 = 0.25 mM, Nearly complete TOC [32]
Degussa P-25 50 21 75% Anatase, 25% Rutile TiO2 = 1 g/l, 125W Mp removal achieved in
Millennium PC 287 5–10 Anatase Hg, pH = not available 90 min.
500 Acridine Orange1, C0 = 1.0 mM, Complete degradation of [31]
2
Hombikat UV 100 250 5 Anatase Ethidium bromide TiO2 = 1 g/l, 125 W Mp dye1 and dye2 takes
TTP 9.82 N/A N/A Hg, pH = not available place in 75 and 195 min
respectively.
N/A = Not available.
S. Ahmed et al. / Desalination 261 (2010) 3–18 7

4-NP N 2,4-DNPN 2,4,6-TNP. The difference in poly nitrophenols disap- 3.2. Light intensity and wavelength
pearance rates was related to the variation in adsorption behaviour.
For both supported Degussa P25 and Millennium PC-500 photocata- The light intensity determines the extent of light absorption by the
lysts, the maximum quantities of adsorbed phenolic compounds semiconductor catalyst at a given wavelength. The rate of initiation
increase in the order: Phenolb 4-NP b DNPb TNP. Gulliard et al. [39] for photocatalysis electron–hole formation in the photochemical
compared the photocatalytic efficiency of different catalysts with reaction is strongly dependent on the light intensity [46]. Light
various surface areas, crystallite and particle sizes and their chemical intensity distribution within the reactor invariably determines the
of surface for the degradation of 4-chlorophenol. The observed order overall pollutant conversion and degradation efficiency [47]. Conse-
of efficiency based on the rate of degradation upon solar exposure was quently the dependency of pollutant degradation rate on the light
shown to be: TiONA PC 10 (9 m2/g)N P-25 (50 m2/g)N TiL COM HC intensity has been studied in numerous investigations for various
120 (120 m2/g)N Hombikat UV 100 (250 m2/g). The TiONA PC 10 organic pollutants [48,49]. While in some cases the rate of reaction
photocatalyst has an activity about two times higher than Hombikat exhibited a square root dependency on the light intensity, others
UV100 and TiL COM HC 120, and 1.4 times higher than that of P-25. observed a linear relationship between the two variables [48–50].
All these catalysts have mainly the anatase structure. In the presence Ollis et al. [49] reviewed the effect of light intensity on the organic
of P-25, Sevlam et al. [40] indicated that complete degradation of pollutant degradation rate. It has been reported [48–50] that the rate
4-flurophenol using P25 (BET=55 m2/g, particle size=30 nm) was is proportional to the radiant flux Ф for Ф b 25 mW/cm2, and above
achieved in 60 min while complete degradation occurred in 90 min in the 25 mW/cm2 the rate has been shown to be varied as Ф1/2, indicating a
presence of ZnO (BET=10 m2/g, particle size=0.1–4 µm). The higher too high value of the flux and an increase of the electron–hole
efficiency of P-25 is due to its larger surface area compared to the ZnO. recombination rate. At high intensity, the reaction rate was indepen-
The oxidizing power of hydroxyl radicals produced by these catalysts is dent of light intensity. This is likely because, at low intensity, reactions
strong enough to break C–C, C–F and C=C bonds of 4-FP adsorbed on involving electron–hole formation are predominant and electron–
their surfaces leading to the formation of CO2 and mineral acids. hole recombination is insignificant. Puma and Yue [51] examined the
Sobana and Swaminathan [41] compared the efficiencies of effect of light wavelength on the photocatalytic degradation of
various catalysts (ZnO, TiO2 anatase, ZnS, SnO2, Fe2O3 and CdS for 2-Chlorophenol (2-CP) using UV-A alone and individually with
the degradation of Acid red 18 under UV irradiation. The highest simultaneous UV-A, B, and-C radiation. The rate of degradation and
degradation was achieved with ZnO over TiO2 anatase catalyst due mineralization of 2-CP were significantly improved with UV-ABC
to the higher surface area of ZnO (10 m2/g) over TiO2 anatase radiation compared to UV-A radiation. The initial rate of 2-CP deg-
(8.9 m2/g). SnO2, Fe2O3, CdS and ZnS have negligible activity on Acid radation was found to be 1.8 times faster under UV-ABC radiation
Red 18 decolourisation due to their smaller band gaps which permit and the degradation of 2-CP was 98% in 20 min. The enhanced
rapid recombination of hole and electron. Talebian and Nilfor- degradation rate with UV-ABC was attributed to the photon flux with
oushan [42] compared the photocatalytic degradation efficiency of UV-ABC, which was reported to be 1.56 times higher compared to
methylene blue (MB) under UV irradiation. The observed order of UV-A radiation. The rate improvement was related to the combined
photocatalytic efficiency was shown to be SnO2 b ZnO b TiO2 b In2O3. use of photolysis, photocatalysis, and synergistic effects due to
The same trend was also observed at different pH values for the associated photolysis and photocatalysis. Kaneco et al. [52] investi-
photocatalytic activities of different catalysts due to their various gated the effect of light intensity on the solar photocatalytic deg-
micro structures. Zhang et al. [43] compared the efficiency of radation of bisphenol A in water with TiO2 on sunny and cloudy
coupled oxide ZnO–SnO2 with ZnO or SnO2 for the degradation of days. The degradation efficiency is shown to increase rapidly with
methylene orange (MO). The photocatalytic efficiency of ZnO–SnO2 increase in the light intensity up to 0.35 mW/cm2, and then the
was significantly higher compared to ZnO or SnO2 alone. Under the efficiency increased gradually. Venkatachalam et al. [53] compared
experimental conditions, the degradation of MO follows the first- the mineralisation efficiency of 4-CP using lamps of wavelength
order reactions during the first 15 min of irradiation, and the 365 nm and 254 nm over TiO2. The mineralisation rate at 365 nm is
apparent rate constants for ZnO–SnO2, Degussa P25 TiO2, ZnO– reported to be slightly higher than at 254 nm. Since the band gap
SnO2/NaOH, ZnO and SnO2 are 0.0865, 0.0822, 0.0339, 0.0335 and excitation of electrons in TiO2 with 254 nm can promote electrons
0.0124 min− 1, respectively. The higher activity of ZnO–SnO2 is to the conduction band with high kinetic energy, they can reach the
attributed to its smaller size and the charge separation in coupled solid–liquid interface easily, suppressing the electron–hole recom-
oxide. Cun et al. [44] compared the efficacy of ZS, pure ZnO and pure bination in comparison to 365 nm. Chiou et al. [54] studied the effect
SnO2 for the photocatalytic degradation of MO. The degradation of UV light intensity (20–400W) on the phenol degradation. All the
efficiency was reported to follow the order of ZS N ZnO N SnO2. The reactions followed the first order kinetics. In the UV/TiO2 system,
photocatalytic reaction by Z2S is 40.2 and 66.1% faster than those by the degradation rate constants with a light intensity of 20, 100, and
ZS and ZnO, respectively. The higher activity of ZS was related to the 400 W are 8.3 × 10− 3, 0.012 and 0.031 min− 1 respectively. Under the
larger BET of ZS than ZnO, and the suppression of electron–hole conditions tested, an acceptably good linear correlation exists
pairs' recombination. Ksibi et al. [45] studied the photocatalytic between the apparent first order rate constant and light intensity.
degradation of six substituted phenols (2,4,6-TNP, Hydroquinone, Increasing light intensity enhances the formation of hydroxyl radical
2,4-DNP, Resorcinol, phenol and 4-NP) over UV irradiated TiO2. The thereby improving the degradation rate. Shukla et al. [34] studied
adsorption constants were shown to decrease in the order: 2,4- the effect of light intensity (160–330 W) on the photocatalytic deg-
DNP N Hydro N 4-NP N phenol N resorcinol N 2,4,6-TNP. The higher radation of phenol using ZnO and artificial solar light. Under the
value of 2,4-DNP (31 L/mol) was attributed to the ortho-position conditions tested (C0 = 25 ppm, ZnO = 0.4 g/l, Persulfate = 2 g/l),
of nitro (–NO2) group in the aromatic ring. In contrast the lowest the rate of degradation is shown to be favoured by the increase in
adsorption constant of 2,4,6-TNP (1.45 L/mol) is related to the big lamp power with 95% conversion being obtained within 3 h using
geometric hindrance due to the large number of nitro groups on the the lamp power of 330 W as opposed to 4 h while using the lamp
aromatic ring. The influence of the substituents on the photo- power of 160 W. Han et al. [55] studied the photocatalytic degrada-
catalytic degradation of phenols was reported to be Resorcinol N tion and mineralisation of 4-CP, hydroquinone and 4-nitrophenol
phenol.4-NP N 2,4-DNP N Hydro N 2,4,6-TNP. The variation in photo- (4-NP) using UV lamp emitting at 254 nm and VUV lamp emitting
reactivity of these compounds is ascribed to the electronic character at both 254 and 185 nm. VUV irradiation was reported to be effi-
(electro-withdrawing or electro-donor) and to the position of the cient for the degradation of organics. This effect was attributed to the
substituent groups in the aromatic ring. additional photons of 185 nm and the higher energy of illumination
8 S. Ahmed et al. / Desalination 261 (2010) 3–18

with VUV lamps than with UV lamps. The order of TOC removal rate TBP and TMP, the degradation rate constant was shown to increase
of each phenol was shown to be VUVPCD N VUVPD N UVPCD N UVPD, from 0.0252 to 0.0071 min− 1 and 0.0153 to 0.0034 min− 1 respective-
except for 4-NP under VUV irradiation which had no significant dif- ly. TCP with the strongest electronegativity character of the com-
ference due to TiO2 catalyst. The degradation rates in both processes pounds studied exhibits lower affinity to be adsorbed on TiO2 catalyst.
are reported to follow the order: 4-CP N hydroquinoneN 4-NP. As a result of that, the number of reactive sites on TiO2 increased and
higher degradation levels were observed when initial TCP concentra-
3.3. Initial concentration tion was low. Chiou et al. [62] tested the effect of initial concentration
(12.5–200 mg/l) on the photocatalytic degradation of phenol using
Successful application of the photocatalytic oxidation system 0.072 mol% Pr-doped TiO2. The highest degradation was 100% after
requires the investigation of the dependence of photocatalytic 60 min irradiation for a range of concentration 12.5–25 mg/l. The rate
degradation rate on the substrate concentration. Hong et al. [56] of degradation was shown to decrease as the phenol concentration
examined the effect of initial concentration (50–400 ppm) on the increases from 50 to 200 mg/l. Chiou et al. [63] studied the effect of
degradation of phenol over TiO2 when the ratio of H2O/titanium initial concentration (15.3–83.5 mg/l) on the photocatalytic degrada-
isopropoxide is 75. Highest degradation was observed at 50ppm tion of phenol and m-nitrophenol by UV/TiO2 process. More than 90%
initial concentration. Parida and Parija [57] studied the effect of initial of phenol is degraded within 60 min irradiation when the initial
substrate concentration (2–25 g/l) on the photocatalytic degradation concentration is 15.3 mg/l. However, 42% phenol is removed in
of phenol under sunlight, visible and UV light, respectively. With the 180 min when initial concentration increases up to 83.5 mg/l. Under
increase in the substrate concentration, the degradation efficiency the identical conditions, 83% and 43% of m-NP is degraded at an initial
decreased from 100% to 60% under solar irradiation. Under UV light, concentration of 15.3 and 83.5 mg/l, respectively. The limited number
the degradation was found to decrease from 94% to 52% with of surface sites on TiO2 particles may control the photodegradation. As
increasing initial concentration. The degradation was reported to indicated in several investigations, as the concentration of target
decrease from 95% to 50% under visible light. Shukla et al. [34] studied pollutant increase, more and more molecules of the compound get
the effect of initial concentration (12.5–37.5 ppm) on the degradation adsorbed on the surface of the photocatalyst. Therefore, the require-
of phenol using UV–vis/ZnO/persulfate. Under the conditions tested ment of reactive species (·OH and ·O− 2 ) needed for the degradation
(ZnO = 0.4 g/l, persulfate = 2 g/l and power = 330W), the degra- of pollutant also increases. However, the formation of ·OH and ·O− 2
dation was highest at 12.5 ppm. Parida et al. [58] examined the on the catalyst surface remains constant for a given light intensity,
effect of substrate concentration on the degradation of 4-nitrophenol catalyst amount and duration of irradiation. Hence, the available
(20–100 mg/l). The amount of degradation was found to decrease OH radicals are inadequate for the pollutant degradation at higher
from 100 to 40.9% as the initial substrate concentration increases from concentrations. Consequently the degradation rate of pollutant
20 to 100 mg/l. This is due to the absorption of light by the substrate at decreases as the concentration increases [25]. In addition, an increase
high concentration for the given catalyst loading. Priya and Madras in substrate concentration can lead to the generation of intermediates
[33] examined the effect of initial concentration (10–76 ppm) on the which may adsorb on the surface of the catalyst. Slow diffusion of the
photocatalytic degradation of 2,4-dinitrophenol in the presence of generated intermediates from the catalyst surface can result in the
combustion synthesised TiO2 and P25. Maximum degradation was deactivation of active sites of the photocatalyst, and consequently
observed at 76 ppm. Pardeshi and Patil [59] reported that the degra- result in a reduction in the degradation rate. In contrast, at low
dation efficiency of phenol was shown to decrease as the phenol concentration the number of catalytic sites will not be a limiting factor,
concentration increases from 25 to 300 mg/l in the presence of ZnO and the rate of degradation is proportional to the substrate concen-
and sunlight. When the concentrations of phenol are high, more and tration in accordance with apparent first-order kinetics [48,49]. Several
more phenol molecules are adsorbed on the surface of ZnO. In investigations adequately described the dependence of photocatalytic
contrast, the relative number of ·OH and ·O2− radicals attacking the degradation rates on the concentration of various phenols and dyes by
phenol molecules decreases due to constant reaction conditions. As a the Langmuir–Hinshelwood (L–H) kinetics model [64–66]. The L–H
result the photocatalytic degradation decreases. Venkatachalam et al. model is used to describe the dependence of the observed reaction rate
[53] examined the effect of initial concentration (50–300 mg/l) on the on the initial solute concentrations [67], although the model
photocatalytic degradation of 4-CP using nano TiO2. The degradation parameters will be strongly dependent on the composition of the
rate was found to increase up to an initial concentration of 250 mg/l and effluent and other reactor operating conditions.
then decreased. The degradation follows pseudo-first order kinetics
at low 4-CP. This is ascribed to the screening effect at concentrations 3.4. Catalyst loading
greater than 250 mg/l, and hence degradation efficiency de-
creased. Lathasree et al. [60] studied the effect of initial concentration Several studies have indicated that the photocatalytic rate initially
(40–100 ppm) on the photocatalytic degradation of phenol using increases with catalyst loading and then decreases at high values
ZnO as catalyst. The initial rates of photodegradation were high at the because of light scattering and screening effects. The tendency toward
lower concentration range and it decreased with increase in agglomeration (particle–particle interaction) also increases at high
concentration. The degradation was found to follow the first order solids concentration, resulting in a reduction in surface area available
kinetics. For chlorophenols, the initial rate was observed to increase for light absorption and hence a drop in photocatalytic degradation
with increase in the initial concentration range 40–60 ppm and rate. Although the number of active sites in solution will increase with
decreased with further increase in concentration. Barakat et al. [61] catalyst loading, a point appears to be reached where light
tested the effect of initial concentration (12.5–75 ppm) on the penetration is compromised because of excessive particle concentra-
photocatalytic degradation of 2-CP using 0.036 mol% Co-doped TiO2. tion. The trade off between these two opposing phenomena results in
At a concentration of 12.5 ppm, 2-CP can be totally decomposed an optimum catalyst loading for the photocatalytic reaction [15]. A
within 3 h. The degradation process follows a pseudo first-order further increase in catalyst loading beyond the optimum will result in
reaction. The observed rate constant was shown to vary from 0.020 to non-uniform light intensity distribution, so that the reaction rate
0.008 min− 1 as the concentration increases from 12.5 to 75 mg/l. would indeed be lower with increased catalyst dosage. Table 4
Kusvuran et al. [37] examined the effect of initial concentration presents the influence of catalyst loading on the photocatalytic
(0.1–0.5 mM) for the photocatalytic degradation of di and tri-substituted degradation of phenols. Parida et al. [58] studied the effect of ZnO
phenols using UV/TiO2. The degradation rate constant of TCP was concentration (0.2 to 2.0 g/l) on the photocatalytic degradation of
increased from 0.0372 to 0.0136 min− 1 under the conditions tested. For 4-nitrophenol under solar irradiation. At high concentration, the
S. Ahmed et al. / Desalination 261 (2010) 3–18 9

Table 4
Effect of catalyst loading on the photocatalytic degradation of various phenols.

Pollutant Light source Photocatalyst Tested catalyst loading (g/l) Optimum loading (g/l) Ref.

4-Nitrophenol Solar ZnO 0.2–2.0 0.6 [58]


Phenol Solar ZnO 0.5–3.5 2.5 [59]
4-Chlorophenol UV TiO2 1.0–5.0 2.0 [53]
Phenol UV ZnO 1.0–3.0 2.0 [60]
2-Chlorophenol UV Co-TiO2 0..005–0.03 0.01 [61]
Phenol UV TiO2 1.0–4.0 2.0 [56]
Phenol UV Pr–TiO2 0.–1.2 1.0 [62]

degradation rate was observed to have levelled off. Under the of 2.0 g/l for the degradation of 4-CP. At high catalyst loading, the
conditions tested, the degradation of 4-NP was found to increase reported low degradation rate is attributed to the deactivation of
with increase in ZnO loading up to 0.6 g/l. An increase in the amount activated molecules through collision with ground state titania
of catalyst provides an increased number of active sites for adsorption; molecules. Lathasree et al. [60] examined the effect of ZnO on the
however, the simultaneous increase in solution opacity causes a photocatalytic degradation of phenols in the range of 1.0–3.0 g/l. The
decrease in the penetration of the photon flux. This suggests that the optimum degradation was shown to be at 2.0 g/l ZnO under the
amount of photocatalyst to be used should maintain a balance experimental conditions. The decrease in the degradation rate beyond
between these two opposing effects. Selvam et al. [40] indicated catalyst loading of 2.0 g/l was attributed to the screening effect of
that the degradation rate constant of 4-fluorophenol increases from excess catalyst particles in the solution. Barakat et al. [61] studied the
0.0152 to 0.0358 min− 1 as the concentration of TiO2 increases from effect of 0.036 mol% Co-doped TiO2 concentrations (5–30 mg/L) on
50 to 150 mg in the presence TiO2 P25. Further increase in catalyst the photocatalytic degradation of 2-CP. The degradation reached a
loading (150 to 250 mg) results in a decrease in the rate constant from maximum value of 93.4% with catalyst dosage of 10 mg/L. However, a
0.0358 to 0.0296 min− 1. In the case of ZnO, a similar observation has further increase in the catalyst dosage slightly decreased the
been made. This was related to the aggregation of free catalyst degradation efficiency. Hong et al. [56] investigated the effect of
particles and the screening effect resulting from the excessive opacity catalyst dosage (1.0–4.0 g/l) on the photocatalytic degradation of
of the solution. The optimum concentrations were found to be 150 mg phenol. Under the conditions tested, optimum degradation was
TiO2 and 200 mg ZnO for efficient removal of 4-fluorophenol. The achieved at 2.0 g/l. Chiou et al. [62] studied the effect of 0.072 mol%
photodegradation efficiency of BPA was shown to increase when the Pr-doped TiO2 concentration (0.2–1.2 g/L) on the photocatalytic
dosage increased up to 100 mg/200 cm3, and then the efficiency was degradation of 50 mg/l phenol. The degradation increases with
nearly 100% in the TiO2 dosage range 100–600/200 cm3 [68]. increasing dosage of 0.072 mol% Pr-doped TiO2, and reaches a
Pardeshi and Patil [59] studied the effect of ZnO concentration (0.5 maximum of 96.5% at a dosage of 1.0 g/L. However, a further increase
to 3.5 g/l) on the solar photocatalytic degradation of phenol. As the in catalyst dosage to 1.2 g/L, slightly reduces the photodegradation
amount of ZnO increases, the number of phenol molecules adsorbed efficiency due to the decrease in light penetration by the suspension.
was also increased due to an increase in the number of ZnO particles. In order to ensure uniform light intensity in the photocatalytic reactor,
Thus the photocatalytic degradation was enhanced. Further increase optimum catalyst loading must be determined.
in the amount of catalyst showed an adverse effect. Under the
conditions tested, the highest degradation was observed at 2.5 g/l of 3.5. Medium pH
ZnO. The decrease in degradation beyond 2.5 g/l is attributed to the
screening effect of excess ZnO particles in the solution and scatte- Organic compounds in wastewater differ greatly in several
ring of light. Shukla et al. [34] observed an optimum concentration of parameters, particularly in their speciation behaviour, solubility in
0.8 g/l of ZnO for the solar photocatalytic degradation of phenol under water and hydrophobicity. While some compounds are uncharged at
a UV–vis/ZnO/persulfate system. For ZnO loading of 0.2 g/L, the common pH conditions typical of natural water or wastewater, other
degradation efficiency of phenol is shown to be 95% at 4 h while a compounds exhibit a wide variation in speciation (or charge) and
complete degradation is reached within 3 h for 0.4 g/L or within 2 h at physico-chemical properties. At a pH below its pKa value, an organic
higher amounts of catalyst loading. In contrast to accelerating the rate compound exists as a neutral species. Above this pKa value, an organic
of reaction resulting from the addition of excess catalyst, it can compound attains a negative charge. Some compounds can exist in
possibly cause a negative effect by reducing the light transmittivity positive, neutral, as well as negative forms in aqueous solution. This
due to the formation of an opaque solution. Venkatachalam et al. [53] variation can also significantly influence their photocatalytic degra-
tested the effect of catalyst loading (1–5 g/l) on the photocatalytic dation behaviour. The pH of the wastewater can vary significantly. The
degradation of 4-CP and observed an optimum catalyst concentration pH of an aquatic environment plays an important role on the

Table 5
Influence of pH on the photocatalytic degradation of phenol and substitute phenols.

Pollutant Light source Photocatalyst Tested pH range Optimum pH Ref.

4-Nitrophenol Solar ZnO 2.0–7.0 6.0 [58]


Phenol Solar ZnO 3.0–11.0 5.0–7.0 [59]
4-Chlorophenol UV TiO2 4.0–9.0 5.0 [53]
2-phenyl phenol UV ZnO 6.6–12.0 12.0 [73]
2-Chlorophenol UV Co–TiO2 4.0–12.0 9.0 [61]
4-flurophenol UV ZnO/TiO2 4.0–9.0 7.0 [40]
Phenol UV TiO2 3.0–8.0 5.0 [74]
m-Nitrophenol UV TiO2 4.1–12.7 8.9 [63]
4-Chlorophenol UV N–TiO2 2.0–5.0 3.0a [75]
a
Precursor solution pH.
10 S. Ahmed et al. / Desalination 261 (2010) 3–18

photocatalytic degradation of organic contaminants since it deter- Khodja et al. [73] examined the effect of pH (6.6–12) on the
mines the surface charge of the photocatalyst and the size of photocatalytic degradation of 2-phenyl phenol using ZnO. The
aggregates it forms [69,70]. The surface charge of a photocatalyst degradation efficiency was shown to be greater at higher pH due to
and ionization or speciation (pKa) of an organic pollutant can be the abundance of hydroxyl radicals or to the anionic form of the tested
profoundly affected by the solution pH. Electrostatic interaction compound to be oxidized compared to its molecular form. Barakat et al.
between semiconductor surface, solvent molecules, substrate and [61] investigated the effect of pH (4–12) on the photocatalytic
charged radicals formed during photocatalytic oxidation is strongly degradation of 2-CP using 0.036 mol% Co-doped TiO2. At lower pH
dependent on the pH of the solution. In addition, protonation and (4–7), the degradation efficiency was significantly less as compared to
deprotonation of the organic pollutants can take place depending on higher pH, with a maximum value of 93.4% at pH 9. They reported
the solution pH. Sometimes protonated products are more stable that higher concentrations of hydroxyl ions are available at higher pH
under UV-radiation than their main structures [71]. Therefore the pH values to react with the holes to form hydroxyl radicals (·OH), thereby
of the solution can play a key role in the adsorption and photocatalytic enhancing the photodegradation of 2-CP. Further increase in pH
oxidation of pollutants. The ionization state of the surface of the indicated only a small increase in the photodegradation efficiency.
photocatalyst can also be protonated and deprotonated under acidic Selvam et al. [40] studied the effect of pH (4–7) on the photocatalytic
and alkaline conditions respectively as shown in the following degradation of 4-FP. After 30 min irradiation, the pseudo-first order rate
reactions: constants were 0.0233, 0.0370 and 0.028 min− 1 for P25 and 0.0213,
0.0325 and 0.0231 min− 1 for ZnO at pH 4, 7 and 9 respectively. Since
þ þ
pHbPzc TiOH þ H →TiOH2 ð5:1Þ the pKa of 4-flurophenol is 9.89, it exists as molecular form for the
tested pH range. As a result, higher adsorption and consequently higher
− − degradation are reported to be observed at neutral pH for both catalysts.
pHNPzc TiOH þ OH →TiO þ H2 O ð5:2Þ In the presence of P25 and ZnO, the effect of pH on the photocatalytic
degradation of 4-fluorophenol was observed to be pH 7 N pH 9 N pH 4.
The point of zero charge (Pzc) of the TiO2 (Degussa P25) is widely Akbal and Onar [74] examined the effect of pH (3–8) on the photo-
investigated/reported at pH ∼ 6.25 [25]. While under acidic conditions, catalytic degradation of phenol. At low pH degradation of phenol was
the positive charge of the TiO2 surface increases as the pH decreases rapid, and the highest degradation (70.3%) was achieved at pH 5. The
(Eq. (5.1)); above pH 6.25 the negative charge at the surface of the addition of H2O2 to the UV/TiO2 system was observed to enhance the
TiO2 increases with increasing pH. Table 5 shows the influence of pH phenol degradation to 99.2%. Chiou et al. [63] studied the effect of pH
on the photocatalytic degradation of various phenols. At pH 5–9, the (4.1–12.7) on the photocatalytic degradation of phenol and m-NP
Zn2+ and Fe3+ co-doped TiO2 is found to show higher photoactivity (C0 = 0.51 mM, TiO2 = 1 g/l). The degradation efficiency is above 65%
for the degradation of phenol, whereas an optimal pH range of 3–4 within 180 min with the pH lower than 11; however, it reaches 53% at
was reported for the pure TiO2 [72]. Venkatachalam et al. [53] studied pH 12.7. The degradation is over 85% within 180 min for m-NP, and the
the effect of pH (4–9) on the photocatalytic degradation of 4-CP degradation was more effective in neutral and weakly alkaline media.
(250 mg/l) using nano TiO2. The rate of degradation in the acidic pH The degradation decreases in the order pH 8.9 N pH 7 N pH 11. In neutral
range (at 5) was found to be higher than the alkaline pH. The high to weak alkaline media, TiO– is the predominant group on TiO2 particles,
degradation rate in the acidic pH is due to enhanced adsorption of 4- and phenols are primarily present in the molecular forms. Both of them
CP on the surface of nano TiO2 that carries positive charge. combined by hydrogen bonding simply, and thus increased the amount
In the acidic pH, minimisation of electron–hole recombination is of adsorption and enhanced the photodegradation rate. Yu et al. [75]
also an important factor for the enhanced degradation of 4-CP. In tested the influence of a precursor solution pH (2.0–5.0) on the
alkaline pH (at 9), the degradation rate was found to be less due to the photocatalytic activity of N–TiO2 calcined at 500 °C for the degradation
negative charge of nano TiO2 and 4-CP. Parida et al. [58] examined the of 4-CP. The efficiency of N-doped TiO2 was found to increase as the pH
effect of pH (2–7) on the photocatalytic degradation of 4-NP. The decreased from 5.0 to 3.0. Further decrease in pH to 2.0 affected the
degradation rate was found to increase 92% at pH 6.0 and thereafter photocatalytic activity of the catalyst negatively. Hence, the optimum
decreased with increasing pH of the solution. At pH 6, the surface pH for that particular catalyst's preparation was 3.0. A possible reason
charge of ZnO can attract the largest amount of negatively charged for the adverse effect on the photocatalyst performance at low pH is
anions of 4-nitrophenol. At pH above 7, negative charges predominate that the increased of H+ concentration may restrain hydrolysation of
on the surface of ZnO which hinders the degradation of 4-NP. Pardeshi Ti(OBu)4 and thereby reduce the crystal size of the prepared N-doped
and Patil [59] investigated the effect of pH (3–11) on the solar TiO2 nanoparticles. In sum, different phenols have different activity in
photocatalytic degradation of phenol using ZnO suspension. Under photocatalytic reaction. Some are degraded faster in acidic pH, while
the conditions tested, the highest degradation was achieved at pH others degrade faster at higher pH. All these effects can be attributed
5–7. At pH 5–7, most of the phenol remains undissociated, hence the to the nature of the pollutant to be degraded. Thus, it is important to
maximum number of phenol molecules are adsorbed on the ZnO study the nature of the pollutants to be degraded, and determine the
surface and consequently result in enhanced photodegradation. In optimum pH.
an alkaline medium, the ZnO surface is negatively charged and
phenolate intermediates may be repelled away from the ZnO surface 3.6. Co-occurring substances
which opposes the adsorption of substrate molecules on the surface
of the catalyst. As a result, phenol degradation decreases in alkaline The amount of UV absorption is influenced by water transmittance
medium. Using ZnO, Lathasree et al. [60] studied the effect of pH over the spectral UV range of interest. Some common constituents
(3.5–9) on the photocatalytic degradation of phenol, o-chlorophenol that affect water transmittance are dissolved organic matter, nitrate
and p-chlorophenol. The phenol degradation was reported to be and ferrous/ferric ions. The presence of “spectator's components” in
favourable in mild acidic and neutral solutions. However, o-chlor- water can affect adversely the contaminant degradation rates.
ophenol and p-chlorophenol was found to undergo degradation at a Inorganic anions, such as phosphate, sulphate, nitrate, and chloride,
faster rate at lower pH values. This behaviour is attributed to the have been reported to limit the performance of solar based
undissociated state of the chlorophenol species in acidic conditions photocatalysis [76]. Bicarbonate in particular is detrimental to reactor
which favours its adsorption on the ZnO surface. At high pH, performance as it acts as a hydroxyl radical scavenger [76]. Long time
chlorophenols exist as negatively charged anions which weaken its experience with photocatalytic oxidation systems showed that humic
adsorption due to electrostatic repulsion between catalyst and anions. substances in contaminated water can strongly adsorb titanium
S. Ahmed et al. / Desalination 261 (2010) 3–18 11

dioxide particles and reduce activity toward the target substances catalytic oxidation of 4-Chlorophenol. The degradation rate of 4-
[77]. Lin and Lin [78] observed that the presence of humic acid Chlorophenol was shown to decrease from 1.0436 h− 1 to 0.0525 h− 1
caused a significant retardation on the photocatalytic degradation of as the acetonitrile content increases in the reaction mixture. A similar
4-chlorophenol. The observed retardations of humic acids were retardation effect was also observed when 5% isopropanol was added
related to the inhibition (surface deactivation), competition and into the acetonitrile: water (1:1) reaction mixture. Daneshvar et al.
light attenuation effects. Moreover, the presence of humic acid in the [82] studied the effect of ethanol (0–9% v/v) addition on the
reaction mixture has been reported to significantly reduce light photodegradation of Acid Orange 7 in the presence of ZnO. Under
transmission, and therefore the photooxidation rate. Humic acid can the conditions examined, the addition of ethanol inhibited the
also compete with organohalides for the active sites on the TiO2 degradation of AO7 due to ·OH competitive reactions with AO7 and
surface. ethanol. Thus, the incorporation of a pre-treatment process has been
Popadam et al. [79] examined the effect of the water matrix on the indicated to have a beneficial effect on the performance of
decolourisation of acid orange 20. The presence of 5 g/l NaCl and photocatalytic oxidation.
Na2SO4 led to decreased decolourisation which was attributed to the
trapping of photo generated valence band holes and the hydroxyl
radicals by the respective anions. Aarthi and Madras [80] investigat- 3.7. Oxidants/electron acceptor
ed the effect of metal ions (Cu2+,Fe3+,Zn2+, and Al3+) on the photo-
catalytic degradation of Rhodamine B which was shown to follow The electron/hole recombination is one of the main drawbacks in
the order of Cu2+ N Al3 N Zn2 N Fe3+. In the presence of Cu2+ and Zn2+, the application of TiO2 photocatalysis as it causes a waste of energy. In
the degradation of Malachite Green was reduced by 76 and 66% the absence of suitable electron acceptor or donor, recombination step
respectively. In the presence of metal ions, the surface of TiO2 is modified is predominant and thus it limits the quantum yield. Thus it is crucial
due to the adsorption of cations or anions. The effect of metal ions on the to prevent electron–hole recombination to ensure efficient photo-
photocatalytic degradation rates is attributed to the suppression of OH catalysis. Molecular oxygen is generally used as an electron acceptor
radicals due to trapping of the conduction band electrons by the in heterogeneous photocatalytic reactions. Addition of external
adsorbed metal ions. Khodja et al. [73] examined the effect of oxidant/electron acceptors into a semiconductor suspension has
sodium chloride, sulphate and nitrate concentration on the photo- been shown to improve the photocatalytic degradation of organic
catalytic degradation of 2-phenyl phenol. A large amount of these contaminants by (1) removing the electron–hole recombination by
anions is shown to affect negatively the degradation rate. This accepting the conduction band electron; (2) increasing the hydroxyl
inhibition can be attributed to a competitive adsorption on the ZnO radical concentration and oxidation rate of intermediate compounds;
of these anions and molecules of substrate. Selvam et al. [40] and (3) generating more radicals and other oxidizing species to
examined the effect of transition metal ions on the photocatalytic accelerate the degradation efficiency of intermediate compounds
degradation of 4-flurophenol which was shown to be in the order of [25–32]. Since hydroxyl radicals appear to play an important role in
Mg2+ N Fe2+ N Fe3+ N Cu2+ in an acidic medium, whereas the trend the photocatalytic degradation, several researchers have investigated
was observed to be Mg2+ N Fe3+ N Cu2+ N Fe2+ in a basic medium. The the effect of the addition of electron acceptors such as H2O2, KBrO3,
enhancement of degradation in the presence of these ions except Cu2+ is and (NH4)2S2O8 on the photocatalytic degradation of various organic
due to the increase of charge separation by accepting the conduction compounds [26–32] to enhance the formation of hydroxyl radicals as
band electrons. The inhibition of inorganic anions on the degradation of well as to inhibit the electron/hole (e−/h+) pair recombination. In
4-FP was demonstrated to be CO2− − − − 2− all cases, the addition of oxidants has resulted in higher pollutant
3 N HCO3 N Cl N NO3 N SO4 . The main
inhibition from these anions is attributed to their adsorption on the TiO2 degradation rates compared to molecular oxygen. In most of the cases,
surface. Addition of CO2− − − the order of enhancement is reported to be UV/TiO2/H2O2 N UV/TiO2/
3 , HCO3 and Cl ions decreases degradation of
4-flurophenol due to the hydroxyl radical quenching by these ions. The BrO− 2−
3 N UV/TiO2/S2O8 . The enhancement of degradation rate is due

SO2−
·− to the reaction between BrO3– and conduction band electrons. This
4 ions react with hydroxyl radicals to form SO4 ions. The sulphate
radical anion can accelerate the reaction, and no quenching was reaction reduces the recombination of electron–hole pairs.
observed for the NO− 3 ion. Naeem and Feng [81] studied the inhibition − − þ −
BrO3 þ 6eCB þ 6H →Br þ 3H2 O ð7:1Þ
effect of anions on the photocatalytic degradation of phenol. The order of
effects was demonstrated to be Cl− N SO2− − 2−
4 N NO3 N CO3 . This is due to S2O2− .
8 can generate sulphate radical anion (SO4−) both thermally
the fact that the Cl− ions act as radical scavengers, and can react with and photolytically in aqueous solution. SO.−
4 then reacts with H2O to
hydroxyl radicals to form ·Cl2 or ·ClOH− which may retard the produce .OH radicals.
photocatalytic oxidation reaction dramatically. But the inhibitive effect
2− :−
of SO2− 2−
4 can be attributed to the competitive adsorption of SO4 and the S2 O8 →2SO4 ð7:2Þ
target organic compound on the catalyst surface. At certain concentra-
tions, Fe3+ is reported to improve the degradation rate whereas Ca2+, :− :
SO4 þ H2 O→ OH þ SO4 þ H
2− þ
ð7:3Þ
Cu2+, and Mg2+ hinder the degradation process. The improvement in
degradation with Fe3+ addition is due to the electron scavenging effect
that prevents the recombination of electrons and holes which favours With the addition of H2O2, the enhancement of degradation is due
the formation of ·OH and O2− on the TiO2 surface. The decrease in
·
to the increase in the hydroxyl radical concentration as shown by
degradation in the presence of other metal ions is attributed to low Eqs. (7.4) and (7.5);
reduction potential. The effect of surfactants (0.05 mM/L) on the − : −
degradation rate of phenol was reported to be sodium dodecyl benzene H2 O2 þ eCB → OH þ OH ð7:4Þ
sulphate (SDBS) N sodium dodecyl sulfonate (AS) N sodium dodecyl
: :
sulphate (SDS). The presence of anions is reported to alter the ionic H2 O2 → OH þ OH ð7:5Þ
strength of the solution, therefore influencing the catalytic activity and
hence the photocatalytic degradation. The degradation efficiency of the UV/TiO2/oxidant process is
Organic solvents are mostly present in many industrial waste- slightly more in an acidic medium than in a basic medium. Selvam
waters. As a result, their effects on the photocatalytic degradation et al. [40] studied the effect of oxidants on the photocatalytic
system have been investigated in several studies. Lin and Lin [78] degradation of 4-flurophenol which was shown to be in the order of
examined the effects of acetonitrile or isopropanol on the photo- IO− − − −
4 N BrO3 N S2O82 N H2O2 N ClO3 . After 15 min irradiation at pH 4, the
12 S. Ahmed et al. / Desalination 261 (2010) 3–18

degradations of 4-flourophenol were observed to be 79.86%, 66.89%, range 3–9. The degradation rate constant was shown to increase from
60.88%, 53.85% and 46.35% in UV/TiO2/IO− −
4 , UV/TiO2/BrO3 , UV/TiO2/ 6.9 × 103 to 7.9 × 10− 3 min− 1 and 7.0 × 10− 3 to 10.1 × 10− 3 min− 1
H2O2, UV/TiO2/S2O8 and UV/TiO2/ClO3 processes respectively. IO−
2− − 4 respectively for pH values 3 and 5 as the calcination temperatures
was found to be the most efficient oxidant due to the generation of a increased from 300 to 500 °C. Nonetheless, the rate constant was
number of highly reactive intermediate radicals such as IO.3, OH. and higher at 300 °C and 500 °C compared that for 400 °C at pH 9.
IO−
4 . The same trend is also observed in both acidic and basic mediums Increasing calcination temperature results in reduced surface area
for the UV/ZnO/oxidant process. However, the UV/TiO2/oxidant from 175 to 67 m2/g at pH 3 and 170 to 68 m2/g at pH 5, and increased
process is slightly more efficient than the UV/ZnO/oxidant process. crystallinity and particle size from 3.3 to 8.7 nm at pH 3 and 4.0 to
The higher efficiency of BrO− 3 than S2O8
2−
is due to the electron 9.8 nm at pH 9. No significant changes in the anatase and rutile

scavenging effect of BrO3 . Under visible light irradiation, the influence proportion are observed as the calcination temperature increases.
of oxidants on the degradation of Acid Red 88 by Ag–TiO2 was found Gorska et al. [87] studied the effect of calcination temperatures
to be peroxomonosulfate N peroxodisulfate N hydrogen peroxide [83]. (350 °C–750 °C) on the photocatalytic degradation of 0.21 mM phenol
The addition of 50 ml H2O2 to a 1.00 mM p-dihydroxybenzene under UV (25 b λ b 400 nm) and visible (λ N 400 nm) light irradiation.
solution was reported to increase the removal efficiency from 37 to After 60 min irradiation, the maximum amount of phenol degradation
80% after 200 min reaction time due to the high oxidising capacity of was about 80% when P25 TiO2 was calcinated at 350 °C in the visible
hydroxyl radicals [84]. Chiou et al. [54] investigated the effect of H2O2 light. In addition to the largest BET surface area (205.8 m2/g) and the
addition from 1.77 to 88.2 mM on the photocatalytic degradation of smallest crystallite size (8.4 nm) and average pore diameter (8.3 nm)
phenol. The addition of H2O2 from 1.77 to 8.82 mM leads to an of TiO2, this effect was attributed to the efficient absorption of light vis
increase in removal efficiency from 58 to 84% within 3 h. In contrast, region due to the existence of the largest amount of carbon in
phenol is completely degraded within 2.5 and 1 h when the level of aromatic C–C bonds (10.1 at.%). However, the photocatalyst calci-
H2O2 increases to 44.1 and 88.2 mM respectively. nated at 450 °C was found to have the highest activity in UV light
irradiation. Silveyra et al. [88] tested the effect of calcination
3.8. Calcination temperature temperatures (600–800 °C) on the photocatalytic degradation of
phenol using N-doped TiO2 P-25. Highest degradation was obtained at
Structure and size of TiO2 crystallites are significantly dependent 600 °C and no variation was observed at 700 °C and 800 °C. No
on calcination temperatures. Thermal treatment of TiO2 gels at higher discernible change in anatase to rutile ratio is noticed for N2-doped
temperatures promotes phase transformation from thermodynami- TiO2 at 600 °C whereas this ratio is beginning to be noticeable at 700
cally metastable anatase to the more stable and condensed rutile to 800 °C. Liqiang et al. [89] observed that the rutile phase began to
phase. As the dehydration process occurs during heat treatment, appear when the pure TiO2 samples calcinated at 600 °C, and the
crystallites grow to dimensions larger than those of the original anatase phase disappeared when the calcination temperature was at
particles. Table 6 shows the influence of calcination temperatures on 700 °C. In contrast a little rutile phase appeared in the La-doped TiO2
the photocatalytic degradation of various phenolic compounds. samples when the sample was calcinated at 700 °C, and even most
Nanocrystalline titanium oxide powder for catalytic purposes can be was shown to be anatase phase at a calcination temperature of 800 °C.
applied by a variety of techniques including sol–gel method, solvo- A certain temperature, 600 °C was reported to be the optimum
thermal process, reverse micellar, hydrothermal method and electro- calcination temperature for the highest photocatalytic degradation of
chemical methods. An et al. [85] studied the effect of calcination phenol using 5 mol% La-doped TiO2. Wawrzyniak et al. [90] studied
temperatures (350–650 °C) on the photocatalytic degradation of the effect of calcination temperatures (100–800 °C) on the photo-
2,4,6-tribromophenol (TBP) using mesoporous-TiO2 prepared by catalytic degradation of phenol and two azo dyes (reactive Red, direct
hydrothermal assistant sol–gel method. The TBP degradation rate Green) by N doped-TiO2 under visible light irradiation. The highest
constant was found to increase from 0.010 to 0.018 min− 1 when effectiveness of phenol degradation was achieved at a calcination
calcination temperature was increased from 350 to 550 °C. A sharp temperature of 700 °C. In the case of degradation of both of the azo
decrease in the rate constant to 0.004 min− 1 was observed when the dyes, the highest degradation was observed at 500 °C and 600 °C.
sample was thermally treated at 650 °C. As the treatment temperature Cheng et al. [95] studied the effect of calcination times (0.5–3 h) on
increases from 350 to 550 °C, the resulting TiO2 increases its crystallite the photocatalytic degradation of 4-CP using carbon modified TiO2
size from 9.8 to 18.4 nm, while BET surface area deceased from 142 to calcinated at 350 °C under visible light irradiation. The surface area of
49 m2/g. The increased photocatalytic degradation efficiency was synthesised TiO2 decreases from 257 to 119 m2/g and the crystallite
attributed to the increased crystallinity. The decrease in photocata- size increases from 4.35 to 5.58 nm as the calcination times increase
lytic degradation efficiency at 650 °C was ascribed to the completely from 0.5 to 3.0 h. The catalyst calcinated for 1 h is shown to have
collapsed mesoporous structure of TiO2. Pecchi et al. [86] investigated higher photocatalytic activity than the catalyst calcinated for 0.5 h.
the effect of calcination temperatures on the photocatalytic degrada- The higher activity of the catalyst calcinated for 1 h is due to its
tion of pentachlorophenol using the sol–gel method under the pH enhanced crystallization in anatase TiO 2 which acts as the

Table 6
Influence of calcinations temperature on the photocatalytic degradation of phenols.

Pollutant Light source Photocatalyst Range of calcination temperature Optimum calcination temperature Ref.

Phenol UV TiO2 P-25 350–750 °C 350 °C [87]


Visible TiO2 P-25 350–750 °C 450 °C [87]
Phenol Visible N2–TiO2 600–800 °C 700 °C [88]
Phenol UV La– TiO2 600–800 °C 600 °C [89]
Phenol Solar Au–TiO2 275–500 °C 275 °C [91]
p-chlorophenol UV TiO2 300–500 °C 500 °C [86]
4-CP UV Zr4+–TiO2 105–1100 °C 900 °C [92]
Phenol UV/Visible I–TiO2 400–600 °C 400 °C [93]
Phenol Visible N2–TiO2 350–500 °C 400 °C [94]
2,4,6-TBP UV TiO2 350–650 °C 550 °C [85]
Phenol UV Pr–TiO2 100–800 °C 600 °C [62]
S. Ahmed et al. / Desalination 261 (2010) 3–18 13

recombination centres for photogenerated electrons and holes. and visible light irradiation (λ N 400 nm) using I-doped TiO2 with
Nonetheless, calcination time of more than 1 h results in reduced anatase phase. Of the temperatures tested, I-doped TiO2 calcinated at
photocatalytic activity due to the decrease of carbonaceous residues 400 °C shows significantly higher photocatalytic activity compared to
and surface area. Shaktivel et al. [96] studied the effect of calcination undoped TiO2 nanoparticles (P-25 and pure TiO2). Wang et al. [94]
temperatures (400–600 °C) on the photocatalytic degradation of 4-CP tested the effect of calcination temperatures (350–500 °C) on the
under visible light irradiation (λ ≥ 455 nm) using a nitrogen-doped photocatalytic activity of N2-doped TiO2 with anatase phase for
TiO2 prepared by TiCl4 hydrolysis in tetrabutylammonium hydroxide. the degradation of phenol with visible light (λ N 400 nm). Under the
A sample calcinated at 400 °C for 1 h is reported to show 70% TOC conditions tested, mineralisation of phenol was significantly higher at
reduction of 0.25 mM 4-CP. Porter et al. [97] examined the micro- 400 °C due to its smaller particle size and higher adsorption area
structural changes in Degussa P-25 due to heat treatment. The TiO2 toward the organic substrate. Hong et al. [56] examined the influence
powder was annealed from 600 to 1000 °C. Under UV irradiation, the of calcination temperatures (200–700 °C) on the photocatalytic
apparent crystallite size and rutile content of the catalyst increased degradation of phenol using TiO2 prepared by the sol–gel method.
with increasing calcination temperatures, whereas the specific surface Under the experimental conditions, the maximum degradation was
area and the rate of phenol degradation was reported to decrease. observed at a calcination temperature of 400 °C with anatase phase
Sonawane and Dongare [91] tested the effect of calcination tempera- TiO2. The degradation of phenol decreases drastically above 600 °C
tures (275–500 °C) on the photocatalytic activity of 2%-Au/TiO2 for which was attributed to increased crystallite size and rutile phase of
the degradation of phenol under solar light. The catalyst calcinated at titania. Lv and Lu [98] studied the effect of calcination temperatures
275 °C exhibited the highest activity compared to the samples (200–800 °C) on the photocatalytic degradation of phenol using pure
calcinated at higher temperatures. This effect is attributed to the anatase and rutile TiO2 in aqueous suspension, in the absence and
formation of the most active pure anatase phase at this temperature presence of NaF. Under the experimental conditions (Anatase
and the particle size. At lower temperatures the particle growth will TiO2 = 1.0 g/L, pH = 3), the maximum photocatalytic degradation
be limited to the surface area and the number of active sites will be rate was shown to be at 600 °C under UV irradiation (λ N 320 nm).
more, hence the photocatalytic activity obtained is more than for However, the degradation rate was retarded in the presence and
prepared Au/TiO2 and catalyst calcinated at higher temperatures. absence of NaF when rutile TiO2 was used. XRD analysis revealed that
Zhang et al. [43] studied the effect of calcination temperatures the rutile phase appears at calcination temperatures up to 700 °C, and
(300–900 °C) on the photocatalytic degradation of MO using ZnO– almost all the anatase crystallizes to rutile at 800 °C. As the calcination
SnO2. At 350 °C, the higher photocatalytic activity of coupled oxides temperature increased, the particle sizes of both anatase and rutile
was reported due to variation in phase composition and particle size. increased according to the diffraction intensity.
Using Z2S as the catalyst, an optimum temperature of 600 °C was Chiou et al. [62] examined the efficiency of 0.072 mol% Pr-doped
found for the degradation of MO [44]. Lukac et al. [92] studied the TiO2 for the photocatalytic degradation of phenol over a range of
effect of annealing temperature (105–1100 °C) on the photocatalytic calcination temperatures from 100 to 800 °C. XRD patterns indicated
activity of Zr-doped TiO2 for the degradation of 4-chlorophenol. The that the samples annealed at 100–400 °C become crystalline with
catalysts annealed between 800 and 900 °C are reported to be more dominantly anatase structure. Upon annealing at 500–600 °C, the
efficient compared to the standard Degussa P25. The optimum size of samples possess both anatase and rutile structures, whereas the
the Zr-doped anatase crystallites is 69.2 and 86.5 nm obtained at 875 samples annealed at 800 °C are completely transformed to the rutile
and 900 °C respectively. The photocatalyst annealed at 900 °C was structure. The highest degradation is achieved with samples annealed
reported to have a 1.5 times higher degradation rate than the standard at 600 °C (BET = 40 m2/g) compared to the samples annealed at
TiO2 P25. The most photoactive catalyst annealed at 900 °C was 800 °C. There have been several investigations on photocatalytic
shown to have 87 wt.% of anatase and 13 wt.% of rutile. Barakat et al. degradation of various dyes by titania nanoparticles, however,
[61] studied the effect of annealing temperatures (100–800 °C) on the information on the titania film fabrication routes on the photocata-
efficiency of 0.036 mol% co-doped TiO2 for the photocatalytic lytic degradation of dyes is still desirable. Song et al. [99] investigated
degradation of 2-CP. The sample annealed at 800 °C is found to have the effect of titania films with different nanostructures of nanorods
rutile structure and shows lower degradation efficiency compared to (NR), sol–gel film (SG), nanotubes (NT) and nanoparticle aggregates
the one annealed at 600 °C with anatase phase. The enhanced (DP) on the photocatalytic degradation of rodamine B (RB),
photoreactivity is attributed to the small particle size of TiO2 (about methylene blue (MB) and methyl orange (MO). In the case of RB
25 nm) with a large surface area. The band gap energy measured for degradation, the observed order was reported to be NR N DP N SG N NT,
this anatase sample was 3.2 eV, which is identical to that of the bulk and the order was shown to be NR N SG N DP N NT for the degradation of
TiO2. Hong et al. [93] examined the effect of calcination temperatures MB. An order of NT N DP N SG N NR was observed for the degradation of
(400–600 °C) on the photocatalytic degradation of phenol under UV MO. The loss of TiO2 nanoparticle surface area throughout the growth

Table 7
Influence of doping concentration on the photocatalytic degradation of various phenols.

Pollutant Light source Photocatalyst Doping (%) Optimum doping (%) Ref.

Phenol Visible Fe3+–TiO2 0.4–5.1 1.0 [102]


4-Chlorophenol Visible Ce–TiO2 0–1.0 0.6 [103]
2,4-Dichlorophenol Visible V–TiO2 0.25–5.0 1.0 [104]
p-chlorophenol UV Ag–TiO2 0.1–0.5 0.5 [101]
p-chlorophenol UV Pd–TiO2 0.1–0.5 0.5 [101]
Phenol UV Fe3+–TiO2 0–3.0 0.5 [105]
4-Chlorophenol UV Zr4+–TiO2 0.5–5.0 3.0 [106]
2-Chlorophenol UV Co–TiO2 0.004–0.14 0.036 [61]
Phenol UV La–TiO2 1.0–5.0 1.0 [89]
Phenol UV Pr–TiO2 0–0.22 0.072 [62]
2,4,6-Trichlorophenol UV-A Ag–TiO2 0–2.0% 0.5 [107]
14 S. Ahmed et al. / Desalination 261 (2010) 3–18

of nanocrystallites during high temperature calcination (N400 °C) and transfer process. Using the sol–gel method, Venkatachalam et al. [106]
serious aggregation of prepared nanoparticles when dispersed in examined the effect of Zr4+ doping (0.5–5.0 mol% Zr4+) onto TiO2
aqueous solution need to be addressed to achieve further progress. matrix for the photocatalytic degradation of 4-Chlorophenol. The
photocatalytic activity of 3.0 mol% Zr4+-doped TiO2 samples were
3.9. Doping and mixed semiconductor reported to be higher than that of nano TiO2 and Degussa P25.
Enhanced adsorption of 4-CP over the catalyst surface and the
The recombination of photogenerated electrons and holes is the decrease in particle size are indicated to be the reason for high activity
reason behind the low photoactivity of TiO2. Structural imperfections in of the catalyst. The entry of Zr4+ into the lattice of TiO2 creates charge
the TiO2 lattice generate trap sites which may act as recombination compensating anion vacancies in the lattice points of TiO2 which may
centres, leading to a decrease in the levels of electrons and holes. In enhance the adsorption of 4-CP. Zr4+-doped TiO2 contains only
order to enhance the TiO2 photocatalysis as well as the response into the anatase phase whereas pure TiO2 contains both anatase and rutile
visible spectrum of solar light, TiO2 has been doped with certain phase together. The incorporation of Zr4+ into TiO2 led to small grain
transition metals, non-metals and ionic components. Doped ions can size, large surface area and high band gap values. This also led to the
also act as charge trapping sites and thus reduce electron–hole formation of more electron capture traps which contribute to high
recombination. The effect of doping on photocatalytic activity is separation efficiency of photogenerated carriers. Burns et al. [109]
governed by several factors, e.g. the type and concentration of dopant, have studied photodegradation of 2-chlorophenol using sol–gel
preparation method, the structure and the initial concentration of the synthesized Nd-doped TiO2 under UV irradiation. They reported that
pollutants (phenols and its derivatives), and physico-chemical proper- doping TiO2 with Nd3+ reduces the degradation time because of the
ties of the catalyst. Both positive and negative results have been difference in the ionic radii of Nd3+ and Ti4+. Much larger
reported from doping with metal ions. The increase in charge separation substitutional Nd will cause localised charge perturbation and
efficiency will enhance the formation of both free hydroxyl radicals and formation of oxygen vacancies which act as electron traps.
active oxygen species [100]. Table 7 shows the influence of doping on Bellardita et al. [110] investigated the effect of metal loading (W,
the photocatalytic degradation of various phenolic compounds. Kirilov Co and Sm) onto a TiO2 matrix on the photocatalytic degradation of 4-
et al. [101] studied the efficacy of Ag and Pd modified TiO2 prepared by nitrophenol. The photocatalytic activity of the samples with increas-
impregnation method for the photocatalytic degradation of p-chlor- ing metal loading is shown to increase until a maximum value and
ophenol. Under the conditions tested, complete degradation is achieved afterwards decreases. In comparison to Degussa P25, TiO2 loaded with
after 180 min reaction in the case of 0.5%Ag/TiO2 and 0.1%Ag/TiO2. For 1% W and Sm showed a significant improvement in the photocatalytic
0.25%Ag/TiO2, the complete degradation is observed after 240 min. In activity under UV irradiation, whereas Co is reported to be beneficial
contrast, 90% degradation is obtained with unmodified TiO2. Similar only under visible light irradiation. This effect was ascribed to an
trends have also been observed for 0.1%Pd/TiO2 and 0.5%Pd/TiO2. increased charge separation of the photogenerated electron–hole
Comparatively Ag/TiO2 samples showed higher photocatalytic pairs in the presence of dopant ion. Sobana et al. [111] studied the
activity than Pd/TiO2. In comparison to unmodified TiO2, the effect of activated carbon (AC) content onto ZnO for the photocatalytic
improved activity of Ag and Pd modified TiO2 is attributed to better degradation of 4-acetylphenol in the presence of UV irradiation. Bare
charge separation and consequently, to slower recombination. ZnO gives a complete disappearance of 4-AP in about 240 min of UV
Rengaraj and Li [107] investigated the photocatalytic activity of Ag– irradiation. In contrast, 9AC-ZnO removes 4-AP completely from the
TiO2 prepared by the sol–gel method for the degradation and solution within 150 min due to a synergistic effect between ZnO and
mineralisation of 2,4,6-trichlorophenol (TCP) in aqueous solution activated carbon. The distribution of ZnO is found to be more
under UV-A illumination. In Ag-doped TiO2, the Ag content is varied homogeneous in 9AC-ZnO compared to other AC-ZnO catalysts.
from 0.1 to 2.0 wt.%. Degradation of TCP is demonstrated to be more 9AC-ZnO is found to be more efficient in solar light than in UV light.
than 95% within 120 min, and the corresponding TOC reduction is Wang et al. [112] studied the Ag-doped ZnO for the photocatalytic
reported to be 80% in 120 min. The kinetics of TCP degradation follows degradation of MB by varying the doping ratio of Ag/Zn from 0 to 2%.
a pseudo-first-order kinetic. The 0.5%Ag–TiO2 catalyst is shown to Under the conditions tested, Ag doping significantly improved the
have the highest efficiency for the degradation of 2,4,6-TCP. The photocatalytic activity of ZnO due to O vacancies, crystal deficiencies
degradation rate constant is reported to be 0.1046 min− 1 for 0.5%Ag– and increased specific surface area. Barakat et al. [61] studied the
TiO2 compared 0.0682 min− 1 for undoped TiO2. The photocatalytic effect of Co (0.004–0.14 mol%) doping onto TiO2 nanoparticles by a
activity enhancement due to Ag doping is related to better separation sol–gel technique from TiCl4 for the photocatalytic degradation of 2-
of photogenerated charge carriers and improved oxygen reduction Chlorophenol. The photocatalytic degradation efficiency of 0.036 mol
inducing a higher extent of degradation of aromatics. Arabatzis et al. % Co-doped TiO2 was shown to be 75.4% higher compared to undoped
[108] observed a two times faster degradation of methyl orange in the TiO2 under the conditions tested. A further increase in Co concentra-
presence of Au/TiO2 thin films compared to that obtained with the tion to 0.14 mol% resulted in a decrease in the degradation efficiency
original TiO2 material. TiO2 with gold surface coverage of 0.8 μg cm− 2 to 83%.They reported that catalyst samples with 0.036 mol% Co
is shown to be more efficient for photocatalysis. The improvement in particles annealed up to 400 °C were amorphous, and became
the efficiency is reported to be more than 100%. This enhancement is crystalline with dominantly anatase structure upon annealing at
attributed to the presence of Au particles which attract conduction 600 °C. Under irradiation, Co(III) ions in the Co-doped TiO2 work as
band electrons and prevent electron–hole recombination. electron scavengers which may react with the superoxide species and
Sonawane and Dongare [91] tested the effect of Au doping (1–2% prevent the hole–electron recombinations and consequently increase
Au/TiO2) on the photocatalytic degradation of phenol in solar the photo degradation efficiency. Rengifo-Herrera and Pulgarin [113]
light irradiation of 4.5–5 W m2/day. In comparison to undoped TiO2, tested the photocatalytic activity of N,S co-doped and N-doped
TiO2 doped with 1–2%Au showed 2–2.3 times higher photo- commercial anatase TiO2 powders for the degradation of phenol
catalytic activity. The degradation rate constant of pure TiO2 was under simulated solar irradiation. Undoped P25 was shown to have
0.004945 min− 1, whereas the rate constants of 1%Au/TiO2 and 2%Au/ the highest photocatalytic activity. N,S co-doped powders showed
TiO2 were 0.01041 min− 1 and 0.01128 min− 1, respectively. The almost the same photocatalytic activity as undoped TKP 102 (Tayca)
increase in photocatalytic activity of Au-doped TiO2 films was while N-doped TKP 102 was the less active photocatalyst due to N
observed. This increase is attributed to the combined effect of shift impurities on the TiO2 acting recombination centres. Chiou et al. [62]
in absorption wavelength to visible, the shift in Fermi level to more studied the effect of Pr3+ dopant concentration (0.018–0.22 mol%) on
negative potentials and improved efficiency of the interfacial charge the photocatalytic degradation of phenol. Using the undoped sample,
S. Ahmed et al. / Desalination 261 (2010) 3–18 15

42% of 50 mg/l phenol was degraded after 2 h irradiation. The et al. [103] noticed an optimum Ce content of 0.6% (w/w) into the TiO2
degradation efficiency increases with increasing Pr concentration. matrix for enhanced degradation of 4-chlorophenol in the visible
The highest degradation is observed to be 94.4% with the sample light. In addition to the retardation of phase transformation, this
containing 0.072 mol% Pr3+ that was annealed at 600 °C. A further behaviour was related to the shift of TiO2 absorption edge towards
increase in Pr concentration to 0.22 mol% leads to a slight decrease of longer wavelengths, by reducing the band gap of the original material.
degradation efficiency to 91.3%. In the Pr-doped TiO2 particles, Pr+ 3 Wang et al. [117] compared the photocatalytic degradation of
works as an electron scavenger; this may react with superoxide phenol in the presence of N2-doped TiO2 and P25 under visible and
species and prevent the hole–electron recombinations, and thus sunlight irradiation. The photocatalytic activity of the N2-doped TiO2
increase photooxidation efficiency. However, the degradation was with anatase phase is shown to be higher than that of the P25 under
lower than 80% with undoped and other metal-doped TiO2 particles visible light irradiation. P25 exhibits higher photoactivity for the
under similar conditions. Zaleska et al. [114] studied the photo- degradation of phenol compared to N2 doped TiO2. Kim et al. [118]
catalytic activity of TiO2 modified with boric acid for the degradation studied the photocatalytic degradation of 4-CP using anatase, rutile, Ni
of phenol under UV and visible light respectively. The use of boric acid 8 wt.%-doped TiO2 powders (anatase and rutile) under UV and visible
as a boron source resulted in lower photoactivity of the obtained B- light irradiation. TEM and EDS analysis indicated that the added Ni
TiO2 than pure TiO2 under UV light. All tested TiO2 powders with atoms were distributed in the rutile TiO2 lattice which had an average
boric acid series have almost the same photoactivity under grain size of 2–4 nm. BET analysis showed that Ni-doped powder has a
visible light. Naeem and Ouyang [105] studied the effect of Fe3+ larger surface area (234 m2/g) than the other powders (100–150 m2/
doping (0–3.0 mol% Fe3+) on the photocatalytic activity of TiO2 for the g).The UV/VIS-DRS absorption showed that the nano-sized Ni-doped
degradation of phenol under UV light. Fe3+-doped TiO2 was powder had a higher wavelength range (480–500 nm) than the
reported to possess the anatase structure with a range of crystal anatase and rutile powder (380–500 nm). Under both UV and visible
size of 8–11 nm. The highest degradation efficiency was found light irradiation, the effectiveness of Ni 8 wt.%-doped TiO2 (234) for
at 0.5 mol% Fe3+-doped TiO2. Liqiang et al. [89] tested the efficacy of the photocatalytic degradation of 4-CP was much higher than that of
La-doped TiO2 prepared by sol–gel method for the photocatalytic other powders (100–150 m2/g). Adan et al. [102] investigated the
degradation of phenol. After 2 h calcination at 600 °C, the order of effect of Fe3+ doping (0.4–5.1 wt.%) on the photocatalytic degradation
photocatalytic activity of La-doped TiO2 for phenol degradation was of phenol. The photocatalytic activity is shown to be enhanced upon
demonstrated to be 1 N 1.5 N 3 N 0.5 N 5 N 0 mol% La. This is related to the doping with Fe3+ up to ca. 1 wt.%. Doping above ca. 3 wt.% does not
La dopant which could inhibit the crystallite growth and the phase produce any enhancement of the catalytic activity. Nahar et al. [119]
transformation of TiO2 nanoparticles so that the surface content of examined the dependence of the photocatalytic degradation of phenol
oxygen vacancies or defects increased. Venkatachalam et al. [53] on Fe content under UV and visible light irradiation. An increase in the
reported that, under identical and optimal experimental conditions, Fe3+ content (Fe/Ti, x = 0.002 to 0.005) is reported to increase the
1 mol% Mg2+ and Ba2+ doped nano TiO2, Nano TiO2 and P-25 photocatalytic activity, however, the overloaded Fe (x = 0.008, 0.01)
required 300, 360 and 450 min respectively for complete mineralisa- considerably decreased the rate, even below the undoped TiO2. This
tion of 4-Chlorophenol. The photocatalytic degradation of 4-CP over effect was ascribed to the balance of excited electron/hole trapped by
TiO2 and Mg2+ and Ba2+ doped TiO2 indicated the higher activity for the doped Fe3+ and their charges recombination on the doped Fe3+
doped TiO2. Enhanced adsorption of 4-CP on the catalyst surface and level. The molar ratio of Fe/Ti (x = 0.005) was optimum for both the
smaller particle size due to Mg2+ and Ba2+ loadings are indicated to be UV and solar light irradiations. The UV activity was 4.3 times higher
the cause for higher activity of the catalysts. than that of the visible light activity. Gorska et al. [120] compared the
Lee et al. [115] studied the photocatalytic activity of carbon-doped efficiency of phenol degradation for N,C doped-TiO2 prepared by TIP
TiO2 powders synthesized by the sol–gel method for the degradation hydrolysis under visible light to P-25 irradiated by UV light. In the Vis/
of phenol under visible light (λ N 420 nm) irradiation. The particles N,C–TiO2 and UV/P-25 systems, nearly 60 min irradiation resulted in
contain a certain fraction of oxygen vacancies in the anatase structure 69 and 76% of phenol degradation respectively. However, photo-
as evidenced by XPS results. Under the conditions tested, carbon catalytic activity under visible light (N400 nm) of N,C–TiO2 was four
doped TiO2 was reported to be efficient for the degradation of phenol. times higher than photoactivity of P-25 TiO2. Phenol degradation
The high photocatalytic activity of C-doped TiO2 was attributed to the efficiency for commercial TiO2/UV and doped TiO2/Vis systems was
band gap narrowing effect by vacancy states in the anatase structure reported to be comparable.
stabilized by carbon doping. Doping shifted the absorption edge to a
lower energy level, thus effectively increasing the photocatalytic 4. Conclusion
activity in the visible–light region. Tian et al. [104] studied the
photocatalytic activity of V-doped TiO2 for the degradation of MB and Based on recent representative investigations, the role of various
2,4-dichlorophenol using UV and visible light. In the presence of UV operating parameters on the photocatalytic degradation of various
light, 0.5% V-TiO2 was reported to be efficient for the degradation of phenols and substituted phenols has been explored in this review.
MB relative to pure TiO2. Under visible light irradiation, the TiO2 has been suggested to be efficient for the degradation and
degradation rate of 2,4-dichlorophenol over 1% V-TiO2 is shown to mineralisation of various toxic organic pollutants such as phenols and
be two times higher compared to undoped TiO2. Yuan et al. [73] dyes in wastewater water in the presence of UV, visible or solar light
studied the influence of co-doping of Zn2+ and Fe3+ on the and oxygen. The findings also suggest that various operating
photocatalytic degradation of phenol under solar light irradiation. parameters such as photocatalyst type, light intensity, pollutant
The co-doping of 0.5 mol% Zn2+ and 1 mol% Fe3+ onto TiO2 was found type and initial concentration, catalyst amount, initial pH of the
to be two times more efficient compared to pure TiO2. This behaviour reaction medium, catalyst application mode, oxidizing agents/elec-
was attributed to the coupled influence of the co-dopant and titania tron acceptors, and the presence of ionic components in solution can
energy bands. Lee et al. [116] compared the photocatalytic degrada- influence significantly the photocatalytic degradation rate of phenols.
tion of p-nitrophenol over TiO2 and TiO2/SiO2 nanoparticles prepared Optimising the degradation parameters is crucial from the perspective
by the micro-emulsion method using PFPE-NH4 surfactant. TiO2/SiO2 of efficient design and the application of photocatalytic oxidation
(80:20) nanoparticles were reported to have a higher photocatalytic processes to ensure sustainable operation. The application of this
activity than pure TiO2 and the TiO2/SiO2 (90:10) particles. This effect technique on actual mixed matrices merits further investigation to
has been attributed to the decrease in crystallite size from 17 to 13 nm yield stable pollutant removal through the optimisation of process
with an increase in silica content. Under the conditions tested, Silva parameters. Metal and non-metal doped TiO2 have been reported to
16 S. Ahmed et al. / Desalination 261 (2010) 3–18

result in improved degradation rates. While doped TiO2 has shown [17] O. Legrini, E. Oliveros, A.M. Braun, Photochemical processes for water treatment,
Chemical Reviews 93 (1993) 671–698.
significant improvement in the efficiency of TiO2 to remediate [18] E. Moctezuma, E. Leyva, G. Palestino, H. De Lasa, Photocatalytic degradation of
wastewater in laboratory scale systems, it is evident that the need methyl parathion: reaction pathways and intermediate reaction products,
to develop pilot scale treatment systems and to apply the technique in Journal of Photochemistry and Photobiology A: Chemistry 186 (2007) 71–84.
[19] A. Sobczynski, L. Duczmal, W. Zmudzinski, Phenol destruction by photocatalysis
cost effective wastewater purification processes requires continued on TiO2: an attempt to solve the reaction mechanism, Journal of Molecular
in-depth research. In spite of extensive investigations, the commercial Catalysis A: Chemical 213 (2) (2004) 225–230.
exploitation of photocatalysis has been hindered by the lack of efficient [20] H. Al-Ekabi, N. Serpone, E. Pelizzetti, C. Minero, M.A. Fox, R.B. Draper, Kinetic
studies in heterogeneous photocatalysis 2: TiO2-mediated degradation of 4-
and low cost visible light harvesting catalysts, a relatively poor chlorophenol alone and in a three component mixture of 4-chlorophenol, 2,4-
understanding of reactor design criteria, and inadequate scale up dichlorophenol, and 245-trichlorophneol in air equilibrated aqueous media,
strategies. Future research should focus on the development of a more Langmuir 5 (1989) 250–255.
[21] J.A. Riddick, W.B. Vunger, T.K. Sakano, Organic Solvents, 4th editionJhon Wiley &
reliable photocatalyst that can be activated by visible and solar light or
Sons, New York, 1986.
both. In addition, more work is required on the modelling of the [22] Solubility Data Series, International Union of Pure and Applied Chemistry, Vol.20,
photoreactor to optimise its design for pollutant degradation. In the Pergamon Press, Oxford, 1985.
literature, there is currently little information on this aspect. Present [23] H. Stephen, T. Stephan, Solubilities of Organic and Inorganic Compounds,
MacMillan, NewYork, 1963.
research activities at CQ University and QUT, Australia focus on the [24] Z. Guo, R. Ma, G. Li, Degradation of phenol by nanomaterial TiO2 in wastewater,
computational fluid dynamics modelling of a flat plate reactor to Chemical Engineering Journal 119 (2006) 55–59.
optimise its design and to predict its performance. Although this [25] W. Bahnemann, M. Muneer, M.M. Haque, Titanium dioxide-mediated photo-
catalysed degradation of few selected organic pollutants in aqueous suspensions,
review is not comprehensive in the scope of the photocatalytic Catalysis Today 124 (2007) 133–148.
degradation of organic pollutants, it does however address the [26] M. Qamar, M. Saquib, M. Muneer, Titanium dioxide mediated photocatalytic
fundamental principles and recent applications in this area. degradation of two selected azo dye derivatives, chrysoidine R and acid red 29
(chromotrope 2R), in aqueous suspensions, Desalination 186 (2005) 255–271.
[27] M. Qamar, M. Saquib, M. Muneer, Photocatalytic degradation of two selected dye
derivatives, chromotrope 2B and amido black 10B, in aqueous suspensions of
Acknowledgement titanium dioxide, Dyes and Pigments 6 (2005) 1–9.
[28] M. Saquib, M.A. Tariq, M.M. Haque, M. Muneer, Photocatalytic degradation of
This study is supported under an Australian Research Council disperse blue 1 using UV/TiO2/H2O2 process, Journal of Environmental
Management 88 (2008) 300–306.
(ARC) linkage grant in collaboration with CM Concrete Private Limited
[29] M.A. Tariq, M. Faisal, M. Muneer, Semiconductor-mediated photocatalysed
and Department of Public Works, QLD Government. The authors degradation of two selected azo dye derivatives, amaranth and Bismarck brown
gratefully acknowledge the financial support of the ARC project. One in aqueous suspension, Journal of Hazardous Materials B127 (2005) 172–179.
[30] M.A. Tariq, M. Faisal, M. Saquib, M. Muneer, Heterogeneous photocatalytic
author is also grateful for the financial support of the Queensland
degradation of an anthraquinone and a triphenylmethane dye derivative in
Government through the Smart State fellowship scheme. aqueous suspensions of semiconductor, Dyes and Pigments 76 (2008) 358–365.
[31] M. Faisal, M.A. Tariq, M. Muneer, Photocatalysed degradation of two selected dyes in
UV-irradiated aqueous suspensions of titania, Dyes and Pigments 72 (2007) 233–239.
References [32] M. Haque, M. Muneer, TiO2-mediated photocatalytic degradation of a textile dye
derivative, bromothymol blue, in aqueous suspensions, Dyes and Pigments 75
[1] J. Radcliff, Future directions for water recycling in Australia, Desalination 187 (2007) 443–448.
(2006) 77–87. [33] M.H. Priya, G. Madras, Photocatalytic degradation of nitrobenzenes with
[2] V.G. Mitchell, R.G. Mein, T.A. McMahon, Utilising storm water and wastewater combustion synthesized nano-TiO2, Journal of Photochemistry and Photobiology
resources in urban areas, Australian Journal of Water Resources 6 (2002) 31–43. A: Chemistry 178 (2006) 1–7.
[3] Introduction to urban stormwater management in Australia. Department of [34] P.R. Shukla, S. Wang, H.M. Ang, M.O. Tade, Photocatalytic oxidation of phenolic
Environment and Heritage, Prepared under the stormwater initiative of the compounds using zinc oxide and sulphate radicals under artificial solar light,
Living Cities Program 2002. Separation and Purification Technology 70 (2010) 338–344.
[4] E. Eriksson, A. Baun, P.S. Mikkelsen, A. Ledin, Risk assessment of xenobiotics in [35] N.H. Salah, M. Bouhelassa, S. Bekkouche, A. Boultif, Study of photocatalytic
stormwater discharged to Harrestup Ao, Denmark, Desalination 215 (2007) degradation of phenol, Desalination 166 (2004) 347–354.
187–197. [36] A.M. Peiro, J.A. Ayllon, J. Peral, X. Domenech, TiO2-photocatalysed degradation of
[5] N.M. Mahmoodi, M. Armani, N.Y. Lymaee, K. Gharanjig, Photocatalytic phenol and ortho-substituted phenolic compounds, Applied Catalysis B:
degradation of agricultural N-heterucyclic organic pollutants using immobilized Environmental 30 (2001) 359–373.
nanoparticles of titania, Journal of Hazardous Materials 145 (1–2) (2007) 65–71. [37] E. Kusvuran, A. Samil, O.M. Atanur, O. Erbatur, Photocatalytic degradation of di-
[6] Department of Environment Conservation. Managing urban stormwater: and tri-substituted phenolic compounds in aqueous solution by TiO2/UV,
harvesting and reuses, NSW, DEC 2006/137. Applied Catalysis B: Environmental 58 (2005) 211–216.
[7] A. Arques, A.M. Amat, A. Garcia-Ripoll, R. Vicente, Detoxification and/or increase [38] H. Lachheb, A. Houas, J.M. Herrmann, Photocatalytic degradation of polynitro-
of the biodegradability of aqueous solutions of dimethoate by means of solar phenols on various commercial suspended or deposited titania catalysts using
photocatalysis, Journal of Hazardous Materials 146 (2007) 447–452. artificial and solar light, International Journal of Photoenergy 2008 (2008),
[8] I. Oller, W. Gernjak, M.I. Maldonado, L.A. Perez-Estrada, J.A. Sanchez-Perez, S. Malato, Article ID 497895, doi:10.1155/2008/497895.
Solar photocatalytic degradation of some hazardous water-soluble pesticides at [39] C. Guillard, J. Disdier, J.M. Herrmann, C. Lehaut, T. Chopin, S. Malato, J. Blanco,
pilot-plant scale, Journal of Hazardous Materials 138 Part B (2006) 507–517. Comparison of various titania samples of industrial origin in the solar
[9] A. Garcýa, A.M. Amat, A. Arques, R. Vicente, M.F. Lopez, I. Oller, M.I. Maldonado, photocatalytic detoxification of water containing 4-chlorophenol, Catalysis
W. Gernjak, Increased biodegradability of Ultracid™ in aqueous solutions with Today 54 (1999) 217–228.
solar TiO2 photocatalysis, Chemosphere 68 (2007) 293–300. [40] K. Selvam, M. Muruganandam, I. Muthuvel, M. Swaminathan, The influence of
[10] R. Goslich, R. Dillert, D. Bahnemann, Solar water treatment principles and inorganic oxidants and metal ions on semiconductor sensitized photodegrada-
reactors, Water Science and Technology 35 (4) (1997) 137–148. tion of 4-flurophenol, Chemical Engineering Journal 128 (2007) 51–57.
[11] G. Busca, S. Berardinelli, C. Resini, L. Arrigi, Technologies for the removal of [41] N. Sobana, M. Swaminathan, The effect of operational parameters on the
phenol from fluid streams. A short review of recent developments, Journal of photocatalytic degradation of acid red 18 by ZnO, Separation and Purification
Hazardous Materials 160 (2008) 265–288. Technology 56 (2007) 101–107.
[12] L.F. Liotta, M. Gruttadauria, G. Di Carlo, G. Perrini, V. Librando, Heterogeneous [42] N. Talebian, M.R. Nilforoushan, Comparative study of the structural, optical and
catalytic degradation of phenolic substrates: catalysts activity, Journal of photocatalytic properties of semiconductor metal oxides toward degradation of
Hazardous Materials 162 (2009) 588–606. methylene blue, Thin Solid Films (2009), doi:10.1016/j.tsf.2009.07.135.
[13] L. Sun, J.R. Bolton, Determination of the quantum yield for the photochemical [43] M. Zhang, T. An, X. Hu, C. Wang, G. Sheng, J. Fu, Preparation and photocatalytic
generation of hydroxyl radicals in TiO2 suspensions, Journal of Physical properties of a nanometer ZnO–SnO2 coupled oxide, Applied Catalysis A: General
Chemistry 100 (1996) 4127–4134. 260 (2004) 215–222.
[14] P.S. Mukherjee, A.K. Ray, Major challenges in the design of a large scale photocatalytic [44] W. Cun, Z. Jincai, W. Xinming, M. Bixian, S. Guoying, P. Ping'an, F. Jiamo, Preparation,
reactor for water treatment, Chemical Engineering Technology 22 (1999) 253. characterization and photocatalytic activity of nano-sized ZnO/SnO2 coupled
[15] A.A. Adesina, Industrial exploitation of photocatalysis: progress, perspectives photocatalysts, Applied Catalysis B: Environmental 39 (2002) 269–279.
and prospects, Catalysis Surveys from Asia 8 (4) (2004) 265–273. [45] M. Ksibi, A. Zemzemi, R. Boukchina, Photocatalytic degradability of substituted
[16] I.K. Konstantinou, T.A. Albanis, Photocatalytic transformation of pesticides in phenols over UV irradiated TiO2, Journal of Photochemistry and Photobiology A:
aqueous titanium dioxide suspensions using artificial and solar light: inter- Chemistry 159 (2003) 61–67.
mediates and degradation pathways, Applied Catalysis B: Environmental 42 [46] A.E. Cassano, O.M. Alfano, Reaction engineering of suspended solid heteroge-
(2003) 319–335. neous photocatalytic reactors, Catalysis Today 58 (2000) 167–197.
S. Ahmed et al. / Desalination 261 (2010) 3–18 17

[47] V. Pareek, S. Chong, M. Tade, A. Adesina, Light intensity distribution in [77] D.Y. Goswami, A review of engineering developments of aqueous phase solar
heterogeneous photocatalytic reactors, Asia–Pacific Journal of Chemical Engi- photocatalytic detoxification and disinfection processes, Journal of Solar Energy
neering 3 (2008) 171–201. Engineering 119 (1997) 101–107.
[48] J.M. Hermann, Heterogeneous photocatalysis: fundamentals and applications to [78] C. Lin, K. Lin, Photocatalytic oxidation of toxic organohalides with TiO2/UV: the effects
the removal of various types of aqueous pollutants, Catalysis Today 53 (1999) of humic substances and organic mixtures, Chemosphere 66 (2007) 1872–1877.
115–129. [79] T. Papadam, N.P. Xekoukoulotakis, I. Poulios, D. Mantzavinos, Photocatalytic
[49] D.F. Ollis, E. Pelizzetti, N. Serpone, Photocatalyzed destruction of water transformation of acid orange 20 and Cr(VI) in aqueous TiO2 suspensions,
contaminants, Environmental Science and Technology 25 (9) (1991) Journal of Photochemistry and Photobiology A: Chemistry 186 (2007) 308–315.
1522–1529. [80] T. Aarthi, G. Madras, Photocatalytic degradation of Rhodamine dyes with nano-
[50] G. Al-Sayyed, J.C. D'Oliveira, P. Pichat, Semiconductor-sensitized photodegrada- TiO2, Industrial and Engineering Chemistry Research 46 (2007) 7–14.
tion of 4-chlorophenol in water, Journal of Photochemistry and Photobiology A: [81] K. Naeem, O. Feng, Parameters effect on heterogeneous photocatalysed
Chemistry 58 (1) (1991) 99–114. degradation of phenol in aqueous dispersion of TiO2, Journal of Environmental
[51] G.L. Puma, P.L. Yue, Effect of the radiation wavelength on the rate of Science 21 (2008) 527–533.
photocatalytic oxidation of organic pollutants, Journal of Industrial and [82] N. Daneshvar, M.H. Rasoulifard, A.R. Khataee, F. Hosseinzadeh, Removal of C.I.
Engineering Chemistry Research 41 (2002) 5594–5600. Acid Orange 7 from aqueous solution by UV irradiation in the presence of ZnO
[52] S. Kaneco, M.A. Rahman, T. Suzuki, H. Katsumata, K. Ohta, Optimization of solar nanopowder, Journal of Hazardous Materials 143 (2007) 95–101.
photocatalytic degradation conditions of bisphenol A in water using titanium [83] S. Anandan, P. Satish Kumar, N. Pugazhenthiran, J. Madhavan, P. Maruthamuthu,
dioxide, Journal of Photochemistry and Photobiology A: Chemistry 163 (2004) Effect of loaded silver nanoparticles on TiO2 for photocatalytic degradation of
419–424. Acid Red 88, Solar Energy Materials & Solar Cells 92 (2008) 929–937.
[53] N. Venkatachalam, M. Palanichamy, V. Murugesan, Sol–gel preparation and [84] K.A. Halhouli, Effect of pH and temperature on degradation of dilute
characterization of alkaline earth metal doped nano TiO2: efficient photocata- dihydroxybenzene, in aqueous titanium dioxide suspension irradiated by UV
lytic degradation of 4-chlorophenol, Journal of Molecular Catalysis A: Chemical light, Journal of Photochemistry and Photobiology 200 (2009) 421–425.
273 (2007) 177–185. [85] T. An, J. Liu, G. Li, S. Zhang, H. Zhao, X. Zeng, G. Sheng, J. Fu, Structural and
[54] C.H. Chiou, C.Y. Wu, R.S. Juang, Influence of operating parameters on photocatalytic degradation characteristics of hydrothermally treated mesopor-
photocatalytic degradation of phenol in UV/TiO2 process, Chemical Engineering ous TiO2, Applied Catalysis A: General 350 (2008) 237–243.
Journal 139 (2008) 322–329. [86] G. Pecchi, P. Reyes, P. Sanhueza, J. Villasenor, Photocatalytic degradation of
[55] W. Han, W. Zhu, P. Zhang, Y. Zhang, L. Li, Photocatalytic degradation phenols in pentachlorophenol on TiO2 sol–gel catalysts, Chemosphere 43 (2001) 141–146.
aqueous solution under irradiation of 254 and 185 nm UV light, Catalysis Today [87] P. Gorska, A. Zaleska, E. Kowalska, T. Klimczuk, J.W. Sobczak, E. Skwarek, W.
90 (2004) 319–324. Janusz, J. Hupka, TiO2 photoactivity in vis and UV light: The influence of
[56] S.S. Hong, C.S. Ju, C.G. Lim, B.H. Ahn, K.T. Lim, G.D. Lee, A photocatalytic calcinations temperature and surface properties, Applied Catalysis B: Environ-
degradation of phenol over TiO2 prepared by sol–gel method, Journal of mental 84 (2009) 440–447.
Industrial and Engineering Chemistry 7 (2) (2001) 99–104. [88] R. Silveyra, L.D. Torre Saenz, W.A. Flores, V.C. Martýnez, A.A. Elguezabal, Doping
[57] K.M. Parida, S. Parija, Photocatalytic degradation of phenol under solar radiation of TiO2 with nitrogen to modify the interval of photocatalytic activation towards
using microwave irradiation zinc oxide, Solar Energy 80 (2006) 1048–1054. visible radiation, Catalysis Today 107–108 (2005) 602–605.
[58] K.M. Parida, S.S. Dash, D.P. Das, Physico-chemical characterization and photo- [89] J. Liqiang, S. Xiaojun, X. Baifu, W. Baiqi, C. Weimin, F. Honggang, The preparation
catalytic activity of zinc oxide presented by various methods, Journal of Colloid and characterization of La doped TiO2 nanoparticles and their photocatalytic
and Interface Science 298 (2006) 787–793. activity, Journal of Solid State Chemistry 177 (2004) 3375–3382.
[59] S.K. Pardeshi, A.B. Patil, A simple route for photocatalytic degradation of phenol in [90] B. Wawrzyniak, A.W. Morawski, B. Tryba, Preparation of TiO2-nitrogen-doped
aqueous zinc oxide suspension using solar energy, Solar Energy 82 (2008) 700–705. photocatalyst active under visible light, International Journal of Photoenergy
[60] S. Lathasree, A.N. Rao, B. Siva Sankar, V. Sadasivam, K. Rengaraj, Heterogeneous 2006 (2006), Article ID 68248, doi:10.1155/IJP/2006/68248.
photocatalytic mineralization of phenols in aqueous solutions, Journal of [91] R.S. Sonawane, M.K. Dongare, Sol–gel synthesis of Au/TiO2 thin films for
Molecular Catalysis A: Chemical 223 (2004) 101–105. photocatalytic degradation of phenol in sunlight, Journal of Molecular Catalysis
[61] M.A. Barakat, H. Schaeffer, G. Hayes, S. Ismat-Shah, Photocatalytic degradation of A: Chemical 243 (2006) 68–79.
2-chlorophenol by co-doped TiO2 nanoparticles, Applied Catalysis B: Environ- [92] J. Lukac, M. Klementova, P. Bezdicka, S. Bakardjieva, J. Subrt, L. Szatmary, Z. Bastl,
mental 57 (2005) 23–30. J. Jirkovsky, Influence of Zr and TiO2 doping ion on photocatalytic degradation of
[62] C.H. Chiou, R.S. Juang, Photocatalytic degradation of phenol in aqueous solutions 4-chlorophenol, Applied Catalysis B: Environmental 74 (2007) 83–91.
by Pr-doped TiO2 nanoparticles, Journal of Hazardous Materials 149 (2007) 1–7. [93] X. Hong, Z. Wang, W. Cai, F. Lu, J. Zhang, Y. Yang, N. Ma, Y. Liu, Visible-light
[63] C.H. Chiou, C.Y. Wu, R.S. Juang, Photocatalytic degradation of phenol and m- activated nanoparticle photocatalyst of Iodine-doped titanium dioxide, Journal
nitrophenol using irradiated TiO2 in aqueous solutions, Separation and of Chemistry of Materials 17 (2005) 1548–1552.
Purification Technology 62 (2008) 559–564. [94] Z. Wang, W. Cai, X. Hong, X. Zhao, F. Xu, C. Cai, Photocatalytic degradation of
[64] R.W. Mathews, Purification of water with near-UV illuminated suspensions of phenol in aqueous nitrogen-doped TiO2 suspensions with various light sources,
titanium dioxide, Water Research 24 (5) (1990) 653–660. Applied Catalysis B: Environmental 57 (2005) 223–231.
[65] R.W. Mathews, Kinetics of photocatalytic oxidation of organic solutes over [95] Y. Cheng, H. Sun, W. Jin, N. Xu, Photocatalytic degradation of 4-chlorophenol
titanium dioxide catalysis, Journal of Catalysis 111 (1988) 264–272. with combustion synthesized TiO2 under visible light irradiation, Chemical
[66] A. Mills, S. Morris, Photo-mineralization of 4-chlorophenol sensitized by titanium Engineering Journal 128 (2007) 127–133.
dioxide: a study of the initial kinetics of carbon dioxide photo-generation, Journal of [96] S. Shakthivel, M. Janczarek, H. Kisch, Visible light activity and photoelectro-
Photochemistry and Photobiology A: Chem. 71 (1993) 75–83. chemical properties of nitrogen-doped TiO2, The Journal of Physical Chemistry B
[67] C.S. Turchi, D.F. Ollis, Photocatalytic degradation of organic water contaminant 108 (50) (2004) 19384–19387.
mechanisms involving hydroxyl radical attack, Journal of Catalysis 122 (1990) [97] J.F. Porter, Y.G. Li, C.K. Chan, The effect of calcination on the microstructural
178–192. characteristics and photoreactivity of Degussa P-25 TiO2, Journal of Materials
[68] W.T. Tsai, M.K. Lee, T.Y. Su, Y.M. Chang, Photo degradation of biphenol-A in a batch Science 34 (1999) 1523–1531.
TiO2 suspension reactor, Journal of Hazardous Materials 168 (2009) 269–275. [98] K. Lv, C.S. Lu, Different effects of fluoride surface modification on the
[69] H.K. Singh, M. Saquib, M. Haque, M. Muneera, D. Bahnemann, Titanium dioxide photocatalytic oxidation of phenol in anatase and rutile TiO2 suspensions,
mediated photocatalysed degradation of phenoxyacetic acid and 2,4,5-trichlor- Chemical Engineering Technology 31 (9) (2008) 1272–1276.
ophenoxyacetic acid, in aqueous suspensions, Journal of Molecular Catalysis A: [99] X.M. Song, J.M. Wu, M. Yan, Photocatalytic degradation of selected dyes by titania
Chemical 264 (2007) 66–72. thin films with various nanostructures, Thin Solid Films 517 (2009) 4341–4347.
[70] M.M. Haque, M. Muneer, D.W. Bahnemann, Semiconductor-mediated photo- [100] S. Kato, Y. Hirano, M. Iwata, T. Sano, K. Takeuchi, S. Matsuzawa, Photocatalytic
catalyzed degradation of a herbicide derivative, chlorotoluron, in aqueous degradation of gaseous sulphur compounds by silver-deposited titanium
suspensions, Environmental Science and Technology 40 (2006) 4765–4770. dioxide, Applied Catalysis B: Environmental 57 (2005) 109–115.
[71] J. Saien, S. Khezrianjoo, Degradation of the fungicide carbendazim in aqueous [101] M. Kirilov, B. Koumanova, L. Spasov, L. Petrov, Effects of Ag and Pd modifications
solutions with UV/TiO2 process: optimization, kinetics and toxicity studies, of TiO2 on the photocatalytic degradation of p-chlorophenol in aqueous solution,
Journal of Hazardous Materials 157 (2008) 269–276. Journal of the University of Chemical Technology and Metallurgy 41 (3) (2006)
[72] Z. Yuan, J.H. Jia, L.D. Zhang, Influence of co-doping of Zn(II) and Fe(III) on the 343–348.
photocatalytic activity of TiO2 for phenol degradation, Materials Chemistry and [102] C. Adan, A. Bahamonde, M. Fernandez-Garica, A. Martinez-Arias, Structure and
Physics 73 (2002) 323–326. activity of nanosized iron-doped anatase TiO2 catalysts for phenol photocatalytic
[73] A.A. Khodja, T. Sehili, J.-F. Pilichowski, P. Boule, Photocatalytic degradation of 2- degradation, Applied Catalysis B: Environmental 72 (2007) 11–17.
phenylphenol on TiO2 and ZnO in aqueous suspensions, Journal of Photochem- [103] A.M.T. Silva, C.G. Silva, G. Draic, J.L. Faria, Ce-doped TiO2 for photocatalytic
istry and Photobiology A: Chemistry 141 (2001) 231–239. degradation of chlorophenol, Catalysis Today 144 (2009) 13–18.
[74] F. Akbal, N. Onar, Photocatalytic degradation of phenol, Environmental [104] B. Tian, C. Li, F. Gu, H. Jiang, Y. Hu, J. Zhang, Flame sprayed V-doped TiO2
Monitoring and Assessment 83 (2003) 295–302. nanoparticles with enhanced photocatalytic activity under visible light irradi-
[75] H. Yu, X. Zheng, Z. Yin, F. Tao, B. Fang, K. Hou, Preparation of nitrogen-doped TiO2 ation, Chemical Engineering Journal 151 (2009) 220–227.
nanoparticle catalyst and its catalytic activity under visible light, Chinese Journal [105] K. Naeem, F. Ouyang, Preparation of Fe3+-doped TiO2 nanoparticles and its
of Chemical Engineering 15 (6) (2007) 802–807. photocatalytic activity, Physica B: Condensed Matter 405 (1) (2009) 221–226.
[76] M. Abdullah, G. Low, R.W. Mathews, Effects of common inorganic ions on rates of [106] N. Venkatachalm, M. Palanichamy, B. Arabindoo, V. Murugesan, Enhanced
photocatalytic oxidation of organic carbon over illuminated titanium dioxide, photocatalytic degradation of 4-chlorophenol by Zr4+ doped nano TiO2, Journal
Journal of Physical Chemistry 94 (1990) 6820. of Molecular Catalysis A: Chemical 266 (2007) 158–165.
18 S. Ahmed et al. / Desalination 261 (2010) 3–18

[107] S. Rengaraj, X.Z. Li, Enhanced photocatalytic activity of TiO2 by doping with Ag [114] A. Zaleska, J.W. Sobczak, E. Grabowska, J. Hupka, Preparation and photocatalytic
for degradation of 2,4,6-trichlorophenol in aqueous suspension, Journal of activity of boron-modified TiO2 under UV and visible light, Applied Catalysis B:
Molecular Catalysis A: Chemical 243 (2006) 60–67. Environmental 78 (2008) 92–100.
[108] I.M. Arabatzis, T. Stergiopoulos, D. Andreeva, S. Kitova, S.G. Neophytides, P. [115] S. Lee, C.Y. Yun, M.S. Hahn, J. Lee, J. Yi, Synthesis and characterization of carbon-
Falaras, Characterization and photocatalytic activity of Au/TiO2 thin films for doped titania as a visible-light-sensitive photocatalyst, Korean Journal of
azo-dye degradation, Journal of Catalysis 220 (2003) 127–135. Chemical Engineering 25 (4) (2008) 892–896.
[109] A. Burns, W. Li, C. Baker, S.I. Shah, Sol–gel synthesis and characterization of [116] M.S. Lee, G.D. Lee, S.S. Park, C.S. Ju, K.T. Lim, S.S. Hong, Synthesis of TiO2/SiO2
neodymium-ion doped nanostructured titania thin film, Proceedings of the nanoparticles in a water-in-carbon-dioxide microemulsion and their photo-
Materials Research Society Symposium 703 (2002) 193–198. catalytic activity, Journal of Research on Chemical Intermediates 31 (4) (2005)
[110] M. Bellardita, M. Addamo, A. Di Paola, L. Palmisano, Photocatalytic behaviour of 379–389.
metal-loaded TiO2 aqueous dispersions and films, Chemical Physics 339 (2007) [117] Z. Wang, W. Cai, X. Hong, X. Zhao, F. Xu, C. Cai, Photocatalytic degradation of
94–103. phenol in aqueous nitrogen-doped TiO2 suspensions with various light sources,
[111] N. Sobana, M. Muruganandam, M. Swaminathan, Characterization of AC-ZnO Applied Catalysis B: Environmental 57 (2005) 223–231.
catalyst and its photocatalytic activity on 4-acetylphenol degradation, Catalysis [118] D.H. Kim, D.K. Choi, S.J. Kim, K.S. Lee, The effect of phase type on photocatalytic
Communications 9 (2008) 262–268. activity in transition metal doped TiO2 nanoparticles, Catalysis Communications
[112] R. Wang, J.H. Xin, Y. Yang, H. Liu, L. Xu, J. Hu, The characteristics and 9 (2008) 654–657.
photocatalytic activities of silver doped ZnO nanocrystallites, Applied Surface [119] M.S. Nahar, K. Hasegawa, S. Kagaya, Photocatalytic degradation of phenol by
Science 227 (2004) 312–317. visible light-responsive iron-doped TiO2 and spontaneous sedimentation of the
[113] J.A. Rengifo-Herrera, C. Pulgarin, Photocatalytic activity of N, S co-doped and N- TiO2 particles, Chemosphere 65 (2006) 1976–1982.
doped commercial anatase TiO2 powders towards phenol oxidation and E. coli [120] P. Gorska, A. Zaleska, J. Hupka, Photodegradation of phenol by UV/TiO2 and Vis/N,
inactivation under simulated solar light irradiation, Journal of Solar Energy 84 C–TiO2 processes: comparative mechanistic and kinetic studies, Separation and
(1) (2010) 37–43. Purification Technology 68 (2009) 90–96.

You might also like