You are on page 1of 77

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233272746

Advanced Oxidation Processes for Wastewater Treatment:


Formation of Hydroxyl Radical and Application

Article  in  Critical Reviews in Environmental Science and Technology · February 2012


DOI: 10.1080/10643389.2010.507698

CITATIONS READS

876 26,737

2 authors, including:

Jianlong Wang
Tsinghua University
634 PUBLICATIONS   35,034 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

capacitive deionization View project

Bio-hydrogen production from organic wastes View project

All content following this page was uploaded by Jianlong Wang on 26 March 2015.

The user has requested enhancement of the downloaded file.


This article was downloaded by: [Tsinghua University]
On: 24 September 2012, At: 19:33
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Critical Reviews in Environmental


Science and Technology
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/best20

Advanced Oxidation Processes for


Wastewater Treatment: Formation of
Hydroxyl Radical and Application
a a
JIAN LONG WANG & LE JIN XU
a
Laboratory of Environmental Technology, INET, Tsinghua University,
Beijing, P.R. China

Accepted author version posted online: 04 Jul 2011.Version of


record first published: 01 Dec 2011.

To cite this article: JIAN LONG WANG & LE JIN XU (2012): Advanced Oxidation Processes for
Wastewater Treatment: Formation of Hydroxyl Radical and Application, Critical Reviews in
Environmental Science and Technology, 42:3, 251-325

To link to this article: http://dx.doi.org/10.1080/10643389.2010.507698

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.
The publisher does not give any warranty express or implied or make any representation
that the contents will be complete or accurate or up to date. The accuracy of any
instructions, formulae, and drug doses should be independently verified with primary
sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Critical Reviews in Environmental Science and Technology, 42:251–325, 2012
Copyright © Taylor & Francis Group, LLC
ISSN: 1064-3389 print / 1547-6537 online
DOI: 10.1080/10643389.2010.507698

Advanced Oxidation Processes for Wastewater


Treatment: Formation of Hydroxyl Radical
and Application

JIAN LONG WANG and LE JIN XU


Laboratory of Environmental Technology, INET, Tsinghua University, Beijing, P.R. China
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Advanced oxidation processes (AOPs), defined as those technologies


that utilize the hydroxyl radical (·OH) for oxidation, have received
increasing attention in the research and development of wastewa-
ter treatment technologies in the last decades. These processes have
been applied successfully for the removal or degradation of toxic
pollutants or used as pretreatment to convert recalcitrant pollutants
into biodegradable compounds that can then be treated by con-
ventional biological methods. The efficacy of AOPs depends on the
generation of reactive free radicals, the most important of which is
the hydroxyl radical (·OH). The authors summarize the formation
reactions of ·OH and the mechanisms of pollutants degradation.
They cover six types of advanced oxidation processes, including
radiation, photolysis and photocatalysis, sonolysis, electrochemical
oxidation technologies, Fenton-based reactions, and ozone-based
processes. Controversial issues in pollutants degradation mecha-
nism were discussed. They review the application of these processes
for removal of different kinds of toxic pollutants from wastewater,
including aromatic compounds, dyes, pharmaceutical compounds,
and pesticides, with emphasis on the parameters assessed, removal
effectiveness, and the degradation mechanisms of pollutants. The
authors discuss issues associated with practical wastewater treat-
ment and offer suggestions for the direction for future researches.

KEYWORDS: advanced oxidation processes, application, hy-


droxyl radical, mechanism, toxic pollutants, wastewater treatment

Address correspondence to Jian Long Wang, Laboratory of Environmental Tech-


nology, INET, Tsinghua University, Haidian District, Beijing 100084, P.R. China. E-mail:
wangjl@tsinghua.edu.cn

251
252 J. L. Wang and L. J. Xu

1. INTRODUCTION

Advanced oxidation processes (AOPs) have received increasing attention in


the research and development of wastewater treatment technologies in the
last decades. These processes (e.g., cavitation, photocatalytic oxidation, Fen-
ton’s chemistry, ozonation) have been applied successfully for the removal
or degradation of recalcitrant pollutants, or used as pretreatment to convert
pollutants into shorter-chain compounds that can then be treated by con-
ventional or biological methods (Anjaneyulu et al., 2005; Gogate and Pandit,
2004a, 2004b). The efficacy of AOPs depends on the generation of reac-
tive free radicals, the most important of which is the hydroxyl radical (·OH;
Catalkaya and Kargi, 2007; Gogate and Pandit, 2004a, 2004b; Pera-Titus et al.,
2004).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Free radical species are atoms or molecules that are capable of inde-
pendent existence and possess one or more unpaired electrons (Halliwell,
1991), such as superoxide radical (O2 ·−), hydroperoxyl radical (HO2 ·), hy-
droxyl radical (·OH), and alkoxyl radical (RO·; Gomes et al., 2005; Tai et al.,
2002). Among these various radicals, the hydroxyl radical is thought to play
a central role in AOPs for wastewater treatment (Tai et al., 2002). Due to its
high standard potentials of 2.8 V versus normal hydrogen electrode (NHE) in
acidic media and 1.55 V versus NHE in basic media, the hydroxyl radical is
highly reactive and nonselective that can oxidize and decompose numerous
hazardous compounds to CO2 and inorganic ions (Garcı́a-Montaño et al.,
2008; Pera-Titus et al., 2004; Shin et al., 2008). Many experimental results
have shown that the degradation of organic compounds by AOPs mainly
involves ·OH reaction mechanisms (Boonrattanakij et al., 2009; Liu et al.,
2009; Peller et al., 2001; Rao and Chu, 2009; Song et al., 2007c).
Several studies have attempted to investigate the role of ·OH in the de-
colorization of dyes, the cleavage of naphthalene and benzene rings, the ox-
idation of organic compounds, and the remediation of As(III)-contaminated
water (Dutta et al., 2005; Keenan and Sedlak, 2008b; Lee and Yoon, 2004;
Song et al., 2007b; Song et al., 2007c). Furthermore, Rosenfeldt et al. (2006)
compared the ability of several advanced oxidation processes in terms of
energy required to produce ·OH. The available evidence suggests that the
importance of ·OH reactions in wastewater treatment has been recognized,
and the kinetics and mechanisms for these reactions have been comprehen-
sively studied and reviewed (Cooper et al., 2009; Lau et al., 2007; Zhang
et al., 2007). Previous recent reviews have focused on the experimental
setups, operation parameters, degradation mechanisms, and application of
various processes as well as the comparison of these processes with a note
on their advantages and disadvantages for wastewater treatment (Agustina
et al., 2005; Akpan and Hameed, 2009; Anjaneyulu et al., 2005; Gogate
and Pandit, 2004a, 2004b; Ince et al., 2001; Kasprzyk-Hordern et al., 2003;
Pignatello et al., 2006; Rauf and Ashraf, 2009). However, there has been
Advanced Oxidation Processes for Wastewater Treatment 253

little research summarizing the formation mechanisms of hydroxyl radicals


in various advanced oxidation processes.
It should be noted that the advanced oxidation processes are all chain-
reaction sequences and in general involve radical cycle propagation once the
radical is formed (Getoff, 1996; Gogate and Pandit, 2004b; Kasprzyk-Hordern
et al., 2003). Hydroxyl radicals react with molecules in atom abstraction or
addition reactions driven by bond strength energetics resulting in the de-
composition of pollutants, and a model to predict the pathway of hydroxyl
radical-induced chain reaction mechanisms in aqueous phase advanced ox-
idation processes has been developed recently (Li and Crittenden, 2009).
Moreover, no matter how the radical is formed in AOPs, the presence of
radical scavengers in water can inhibit chain reactions, and some ions (e.g.,
CO3 2−, HCO3 −, Cl−) have influence on the reactions, which also need to be
Downloaded by [Tsinghua University] at 19:33 24 September 2012

summarized.
In this article, we review the formation mechanisms of ·OH in advanced
oxidation processes and their application for wastewater treatment, mainly
based on the recent literatures including the researches carried out by our
research group (Hu et al., 2006; He et al., 2007; Hu and Wang, 2007; Pi
et al., 2007; Song et al., 2007a; Song et al., 2008a; Song et al., 2007b; Song
et al., 2008b; Song et al., 2008c; Song et al., 2010; Wang and Wang, 2007a;
Wang and Wang, 2008a, 2008b; Wang and Wang, 2009a, 2009b; Wang and
Wang, 2007b; Xue and Wang, 2008). Six types of advanced oxidation pro-
cesses (e.g., radiation, photolysis and photocatalysis, sonolysis, electrochem-
ical oxidation technologies, Fenton-based reactions, ozone-based processes)
in which ·OH reactions are involved are described. The focus is on the ·OH
formation reactions in simple systems as well as hybrid methods. The rate
constants for radical reactions are presented, and detailed degradation mech-
anisms in various operating conditions are reviewed. Moreover, we discuss
the application of these processes for wastewater treatment.

2. PROCESSES FOR ·OH FORMATION

A detailed description of the radical formation mechanisms of various AOPs


is provided in this section, focusing on the ·OH formation reactions. In this
section we highlight six different advanced oxidation processes, which are
radiation, photolysis and photocatalysis (including VUV, UV/O3 , UV/H2 O2 ,
UV/H2 O2 /O3 , UV/HOCl, and heterogeneous photocatalysis), sonolysis (in-
cluding US, US/O3 , US/H2 O2 , and US/photocatalysis), electrochemical oxi-
dation technologies (including anodic oxidation and photoelectrocatalysis),
Fenton-based reactions (including homogeneous Fenton process, hetero-
geneous Fenton process, and photo-Fenton, sono-Fenton, electro-Fenton,
and photoelectro-Fenton processes), and ozone-based processes (including
254 J. L. Wang and L. J. Xu

ozonation, O3 /H2 O2 , homogeneous catalytic ozonation, heterogeneous cat-


alytic ozonation, and electrolysis–ozonation).

2.1. Radiation
Radiation technology has been applied to a variety of problems of environ-
mental interest, which has been shown to be useful for the decoloration and
degradation of dyes (Chen et al., 2008; Ma et al., 2007), sewage sludge pro-
cessing (Park et al., 2009), the oxidization of organic pollutants (Al-Sheikhly
et al., 2006; Hu and Wang, 2007), pesticides removal (Basfar et al., 2007),
and the decomposition of pharmaceutical compounds (Sánchez-Polo et al.,
2009). Application of radiation technology to sewage sludge processing and
dyes degradation has been reviewed in depth (Rauf and Ashraf, 2009; Wang
Downloaded by [Tsinghua University] at 19:33 24 September 2012

and Wang, 2007b; Wojnárovits and Takács, 2008), and rate coefficients of
water radical reactions with some aromatic molecules and dyes have been
presented (Wojnárovits and Takács, 2008).
The effect of radiation on the degradation of a specified chemical de-
pends on the property of chemical itself and the amount of energy deposited
in the material exposed to a radiation field. The radiolysis of water has been
very well documented (Basfar et al., 2007; Wasiewicz et al., 2006), and the
radiation chemistry of aqueous solution has been well established (Buxton
et al., 1988; Kubesch et al., 2005; Mozumder, 1999; Pálfi et al., 2007; Rauf and
Ashraf, 2009; Sánchez-Polo et al., 2009; Spinks and Woods, 1990; Wasiewicz
et al., 2006; Wojnárovits and Takács, 2008; Yu et al., 2008). Two radioac-
tive sources, the electron beam irradiation carried out under an electron
accelerator and gamma irradiation mostly performed using a Cobalt-60 or
Caesium-137 source, are presently available (Wang and Wang, 2007b). It
should be noted that the pulse radiolysis technique is generally employed to
study the formation and decay of reactive transients under radiolytic condi-
tions (Mezyk et al., 2004; Pálfi et al., 2007; Rauf and Ashraf, 2009; Zhang et al.,
2007). Under electron beam or gamma irradiation, hydroxyl radical (·OH),
hydrogen atom (·H), hydrated electron (eaq −), H2 , H2 O2 , Haq +, and OHaq −
are formed in dilute aqueous solutions, as given in following reaction. The
number in parentheses in the equation refers to G value, which is the number
of produced or consumed species per 100 eV absorbed energy. The radiation
yield (G value) depends on the linear energy transfer value of the radiation
and the pH value of solution, where in the range of pH 6–8.5, the yield
of these species are shown in Eq. (1) (Wasiewicz et al., 2006).

H2 O → ·OH(2.8), ·H(0.6), eaq − (2.7), H2 (0.45), H2 O2 (0.72),


Haq + (3.2), OHaq − (0.5) (1)

In air-free media, the primary products of water radiolysis are eaq −,


·H, and ·OH (Wojnárovits and Takács, 2008). The hydrated electron (eaq −)
Advanced Oxidation Processes for Wastewater Treatment 255

and the hydrogen atom are the main reductive species, whereas the ·OH
is the main oxidizing species produced in irradiated aqueous solution. In
the presence of air, the reductive radicals, eaq − and ·H, are converted into
peroxyl radicals by reactions (2) and (3) (Buxton et al., 1988; Sánchez-Polo
et al., 2009; Wasiewicz et al., 2006). Perhydroxyl radicals (HO2 ·) and its con-
jugate base O2 ·− exist in a pH-dependent equilibrium, as seen in Eq. (4)
(Wasiewicz et al., 2006). In the absence of metal ions, they have very low
reactivity to many organic compounds, such as phenols and aliphatic acids,
therefore their disproportionation to H2 O2 and O2 at pH < 8 is of major im-
portance (Eqs. [5–7]) (Basfar et al., 2005; Sánchez-Polo et al., 2009). Kubesch
et al. (2005) investigated the influence of oxygen on the radiation-induced
degradation of catechol in distilled water. Under air saturation conditions,
at 12 kGy the total organic carbon (TOC) was reduced by 63%, without air
Downloaded by [Tsinghua University] at 19:33 24 September 2012

saturation by 17.5%.

·H + O2 → HO2 ·, k = 2.0 × 1010 M−1 s−1 (2)


− − −1 −1
eaq + O2 → O2 · , k = 1.9 × 10 M s
10
(3)
+ −
HO2 · ↔ H + O2 · , pK = 4.8 (4)
HO2 · + O2 ·− → H2 O2 + O2 (pH < 7), k = 9.7 × 107 M−1 s−1 (5)
HO2 · + HO2 · → H2 O2 + O2 , k = 8.3 × 105 M−1 s−1 (6)
− − −1 −1
O2 · + O2 · → H2 O2 + O2 , k < 10 M s (7)

Reactions with just the hydroxyl radical are achieved by using a nitrous
oxide (N2 O) saturated solution that quantitatively converts eaq − and ·H to
the ·OH radical in the reactions (8) and (9) (Buxton et al., 1988). Yu et al.
(2008) found that under O2 and N2 O saturations the TOC percentage removal
efficiencies were enhanced in the radiolytic decomposition of cefaclor. The
enhancement of TOC percentage removal efficiencies might be based on the
fast conversion reactions of eaq − and ·H into oxidizing radicals via Eqs. (2),
(3), (8), and (9).

eaq − + N2 O + H2 O → N2 + OH− + ·OH, k = 9.1 × 109 M−1 s−1 (8)


·H + N2 O → ·OH + N2 , k = 2.1 × 106 M−1 s−1 (9)

Radical–radical recombination reactions also happen, which increase at


higher radiation dose rates (Goulet and Jay-Gerin, 1992; Rauf and Ashraf,
2009; Sánchez-Polo et al., 2009). Hydroxyl radicals can react with eaq −, ·H,
and ·OH to produce negative ions OH−, H2 O, and H2 O2 via Eqs. (10–12),
resulting in a decrease in the effective radical concentrations (e.g., ·OH) for
the oxidation of pollutants (Goulet and Jay-Gerin, 1992; Lin et al., 1995; Rauf
and Ashraf, 2009; Sánchez-Polo et al., 2009; Yu et al., 2008). The recombi-
nation of reductive species can also occur via reactions (13–15) (Goulet and
256 J. L. Wang and L. J. Xu

Jay-Gerin, 1992; Rauf and Ashraf, 2009).

·OH + eaq − → OH− , k = 3.0 × 1010 M−1 s−1 (10)


·OH + ·H → H2 O, k = 7.0 × 109 M−1 s−1 (11)
−1 −1
·OH + ·OH → H2 O2 , k = 5.5 × 10 M s 9
(12)
− − −1 −1
eaq + ·H → H2 + OH , k = 2.5 × 10 M s 10
(13)
eaq − + eaq − → OH− + OH− + H2 , k = 5.4 × 109 M−1 s−1 (14)
·H + ·H → H2 , k = 7.8 × 109 M−1 s−1 (15)

The reactions of eaq − and ·H with H2 O2 produce ·OH radicals (Rauf and
Ashraf, 2009). The ·OH can also react with H2 O2 produced in solution and
Downloaded by [Tsinghua University] at 19:33 24 September 2012

generate HO2 · species as shown in Eq. (18). In strong basic solutions, the
·OH can act as a weak acid to undergo the reaction (19).

eaq − + H2 O2 → ·OH + OH− , k = 1.2 × 1010 M−1 s−1 (16)


−1 −1
·H + H2 O2 → ·OH + H2 O, k = 9.0 × 10 M s 7
(17)
·OH + H2 O2 → H2 O + HO2 ·, k = 2.7 × 107 M−1 s−1 (18)
·OH + OH− → O− + H2 O, k = 1.2 × 1010 M−1 s−1 (19)

Rate coefficients of ·OH, eaq − and ·H reactions with some aromatic


compounds and dyes have been summarized in the literature (Wojnárovits
and Takács, 2008). The rate coefficients of ·OH radical reactions with organic
compounds show very little variation with the molecular structure, which
are 0.24–2.0 × 1010 M−1s−1. The ·OH usually abstracts H on reaction with
saturated organic compounds, whereas it undergoes addition reaction with
organic compounds containing double bonds or aromatic systems (Rauf and
Ashraf, 2009; Wasiewicz et al., 2006; Wojnárovits and Takács, 2008).
H2 O2 and O3 can accelerate the degradation of pollutants under radi-
ation. When H2 O2 is added, ·OH radicals are formed by the radiolysis of
H2 O2 from Eq. (20) (the bond dissociation enthalpy of the O–O bond is
210 kJ mol−1) and by the reactions of H2 O2 with eaq − and ·H from Eqs. (16)
and (17). However, with the addition of H2 O2 , it can compete with pollu-
tants for ·OH (Eq. [18]), and therefore may reduce the degradation efficiency
(Rauf and Ashraf, 2009; Sánchez-Polo et al., 2009).

H2 O2 → ·OH + ·OH (20)

In the presence of ozone, besides the direct ozonation of pollutants, the pri-
mary radicals of water radiolysis can also react quickly with ozone, whereby,
among several ozone derived transients, ·OH radicals are formed from
reactions (21–26) (Bielski et al., 1985; Bühler et al., 1984; Hoigné, 1998;
Advanced Oxidation Processes for Wastewater Treatment 257

Kubesch et al., 2005; Sehested et al., 1983; Staehelin and Hoigné, 1982).
Under gamma irradiation, ozone is also radiolyzed and numerous reactive
species (e.g., O·−, O2 ·−, O3 ·−) are formed, which is detailed in literature
(Popov and Getoff, 2004). Based on the synergistic effect of radiation and
even small amounts of ozone, a very efficient degradation of organic pollu-
tants (e.g., chlorinated phenols) can be realized (Getoff, 2002).

eaq − + O3 → O3 ·− , k = 3.7 × 1010 M−1 s−1 (21)


−1 −1
·H + O3 → HO3 ·, k = 3.7 × 10 M s 10
(22)
−1 −1
·OH + O3 → HO2 · + O2 , k = 1.1 × 10 M s
8
(23)
HO3 · ↔ O3 ·− + H+ , pK = 6.1 (24)
O2 ·− + O3 → O3 ·− + O2 , k = 1.5 × 109 M−1 s−1
Downloaded by [Tsinghua University] at 19:33 24 September 2012

(25)
− + −1 −1
O3 · + H → HO3 · → ·OH + O2 , k = 9.0 × 10 M s
10
(26)

Moreover, additives generally present as ions are initially added to the aque-
ous solutions as ionic compounds to improve the industrial process (Rauf
and Ashraf, 2009). For instance, CO3 2− and HCO3 − ions are usually added to
adjust the pH of the solution (Mezyk et al., 2004). The presence of these ions
causes a certain decrease in degradation of pollutants, which is explained on
the basis of their chemical reactions with ·OH radicals already produced in
solution (Mack and Bolton, 1999; Rauf and Ashraf, 2009; Zhao et al., 2009b).
Thus, it is possible to change the types and amount of primary reacting
radicals by selecting appropriate experimental conditions, such as purging
with N2 , O2 or N2 O and adding H2 O2 . Accordingly, specific information
about the mechanism of undergoing reactions and their possible reaction
pathways can be obtained (Wojnárovits and Takács, 2008). Zhang et al.
(2007) established both highly oxidative and reductive conditions for radi-
olytic degradation of nitrobenzene through saturation of irradiated solutions
with N2 O to convert eaq − and ·H to ·OH, or through addition of sodium
formate (HCO2 Na) to scavenge ·OH and ·H and purging the solutions with
argon gas (Ar) to eliminate O2 . The mechanisms and main pathways of
nitrobenzene degradation under oxidative and reductive conditions were
proposed.

2.2 Photolysis and Photocatalysis


2.2.1 VUV, UV/O3 , UV/H2 O2 , UV/H2 O2 /O3 , AND UV/HOCL PROCESSES
Photolysis of water using the high energy associated with UV radiation of
a wavelength shorter than 190 nm yields powerful oxidizing species (·OH;
the quantum yield OH = 0.42 at 172 nm) and reducing species (·H; H <
0.05 at 172 nm) that can degrade contaminant in water (Azrague et al., 2005;
Heit and Braun, 1997; Kibanova et al., 2009). The vacuum-ultraviolet (V-UV,
258 J. L. Wang and L. J. Xu

172 nm) photolysis of pure water free from dioxygen has been studied,
and the reactions involved with ·OH and H2 O2 are available in the work of
Azrague et al. (2005). Low-pressure mercury vapor UV lamps with a 254 nm
peak emission, which are typically used to produce UV radiation, cannot
be used as an effective procedure for the total mineralization of pollutants,
although it is quite efficient for water disinfection. The direct photolysis of
butylated hydroxyanisole (BHA) by UV at 254 nm has been studied, which
revealed that the treatment process is less effective than the UV/O3 process,
and the dimerization pathway instead of hydroxyl radical reaction is believed
to be the principal reaction pathway (Lau et al., 2007).
Combination of UV with ozone (UV/O3 ) results in a net enhancement
of organic matter degradation due to direct and indirect production of ·OH
radicals (Garoma and Gurol, 2004; Reisz et al., 2003; Song et al., 2008c;
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Tezcanli-Güyer and Ince, 2004). Upon photolysis (λ < 300 nm), O3 is de-
composed into O2 and oxygen atom O(1D) via reaction (27). O(1D) is very
energetic and therefore reacts fast with practically all conceivable substrates,
including water (Eq. [28]), and it is believed that O(1D) reacts predominantly
by insertion into the C–H or the O–H bond. The excess energy of the H2 O2
molecule so formed results in the fragmentation of the O–O bond (Eq. [29])
(Garoma and Gurol, 2004; McKay and Wright, 1998; Reisz et al., 2003; Song
et al., 2008c; Tezcanli-Güyer and Ince, 2004).

O3 + hv → O2 + O(1 D), O ≈ 0.9 (27)


−1 −1
O( D) + H2 O → H2 O2 (hot),
1
k = 1.8 × 10 M s
10
(28)
H2 O2 (hot) → 2·OH (29)

The ozone is destroyed in a short reaction chain consisting of following


reactions, which is initiated by ·OH radicals (Garoma and Gurol, 2004; Reisz
et al., 2003).

O3 + ·OH → HO2 · + O2 , k = 1.1 × 108 M−1 s−1 (23)


+ −
HO2 · ↔ H + O2 · , pK = 4.8 (4)
− − −1 −1
O3 + O2 · → O3 · +O2, k = 1.6 × 10 M s
9
(25)
O3 + HO2 · → ·OH + 2O2 , k = 1.2 × 106 M−1 s−1 (30)

In the presence of OH− as an initiator, the major reactions for the production
of ·OH and the reactions of ·OH with various species in natural water are
given in following reactions (Christensen et al., 1982; Garoma and Gurol,
2004; Reisz et al., 2003; Staehelin and Hoigné, 1982).

O3 + OH− → O2 + HO2 − , k = 48 M−1 s−1 (31)


+ −
HO2 · ↔ H + O2 · , pK = 4.8 (4)
Advanced Oxidation Processes for Wastewater Treatment 259

O3 + O2 ·− → O3 ·− + O2 , k = 1.6 × 109 M−1 s−1 (25)


− + −1 −1
O3 · + H → ·OH + O2 , k = 9.0 × 10 M s
10
(26)
−1 −1
·OH + H2 O2 → H2 O + HO2 ·, k = 2.7 × 10 M s 7
(18)
− − −1 −1
·OH + HO2 → HO2 · + OH , k = 7.5 × 10 M s 9
(32)

Ozone reacts with organic compounds either directly by electrophilic attack


or indirectly by radical chain–type reactions. It is investigated that the reac-
tivity of molecular ozone toward organic compounds is very slow compared
to that of ·OH, which indicates that the reaction of ·OH with organic com-
pounds is the main mechanism in the UV/O3 system (Garoma and Gurol,
2004).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Ultraviolet photolysis combined with hydrogen peroxide (UV/H2 O2 ) is


an alternative for the degradation of toxic organics because this process
may occur in nature itself (Catalkaya and Kargi, 2007). The ·OH produced
through UV/H2 O2 system as shown in Eq. (33) activates organic compounds
for oxidations by subtracting hydrogen atoms or by adding to double bonds
(Baxendale and Wilson, 1957; Catalkaya and Kargi, 2007). Increasing of ini-
tial H2 O2 concentration enhances the oxidation process up to certain point,
at which H2 O2 starts to inhibit the photolytic degradation of organic com-
pounds. At higher H2 O2 concentration acting as a free radical scavenger,
reactions (18) and (32) in the H2 O2 photolysis mechanism become more
important, and less reactive HO2 · radicals are formed (Buxton et al., 1988;
Kusic et al., 2006).

H2 O2 + hv → 2 · OH, OH = 0.5 (33)


−1 −1
·OH + H2 O2 → H2 O + HO2 ·, k = 2.7 × 10 M s 7
(18)
− − −1 −1
·OH + HO2 → HO2 · + OH , k = 7.5 × 10 M s 9
(32)

The addition of H2 O2 to UV/O3 process can accelerate the degradation


of pollutants due to the increased rate of ·OH radical generation (Kusic et al.,
2006). Comparative investigations of these processes, UV, UV/O3 , UV/H2 O2
and UV/H2 O2 /O3 , have been applied for the degradation of organic pollu-
tants, with the UV/H2 O2 /O3 process representing the highest mineralization
extent (Azbar et al., 2004; Beltran-Heredia et al., 2001; Kusic et al., 2006;
Peternel et al., 2006). Peternel et al. (2006) found that the mineralization ex-
tents of organic dye C.I. Reactive Red 45 obtained after a 1-hr treatments were
observed in the following order: UV < UV/H2 O2 < UV/O3 < UV/H2 O2 /O3 .
Many researches have been made to investigate the photolysis of aque-
ous chlorine at sunlight and ultraviolet wavelengths (Nowell and Hoigné,
1992a, 1992b; Vogt and Schindler, 1992; Watts and Linden, 2007). In wa-
ters containing chlorine (e.g., swimming pools) with pH in region of 4–8
and in the absence of free ammonia and amines, aqueous chlorine exists
260 J. L. Wang and L. J. Xu

primarily as an equilibrium mixture of hypochlorite (ClO−) and hypochlor-


ous acid (HOCl) with a pK a of 7.5 (Nowell and Hoigné, 1992a). As seen in
Eq. (34), the direct formation of hydroxyl radicals is the predominant reaction
in the photolysis of aqueous chlorine at sunlight and ultraviolet wavelengths,
and the quantum yield is determined at 254 nm (Vogt and Schindler, 1992;
Watts and Linden, 2007). Indirect formation of hydroxyl radicals (from pro-
tonation of O−) and chlorine radicals is energetically accessible according
to Eqs. (35–37) with the quantum yields determined at 254 nm (Buxton and
Subhani, 1972; Nowell and Hoigné, 1992b; Vogt and Schindler, 1992). In
addition to the primary photolysis of HOCl, other chain reactions have been
proposed in which organic solutes (R) and HOCl may act as chain promoters
and ·OH or Cl· as chain carriers (Eqs. [38–41]), which maintain a steady-state
concentration of ·OH and/or Cl· (Nowell and Hoigné, 1992b; Oliver and
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Carey, 1977).

HOCl + hv → ·OH + Cl·, OH = 1.4 (34)


ClO− → Cl· + O− + O(3 P),  = 0.074 (35)
ClO− → Cl· + O− ,  = 0.278 (36)

ClO → Cl· + O( D), 1
 = 0.133 (37)
·OH + RH → R· + H2 O (38)
R· + HOCl → RCl + ·OH (39)
Cl· + RH → R· + HCl (40)
R· + HOCl → ROH + Cl· (41)

2.2.2 HETEROGENEOUS PHOTOCATALYSIS


To date, the synthesis and application of various photocatalysts (Kwon et al.,
2008; Rupa et al., 2007; Song et al., 2008a), the mechanisms and kinetics of
contaminant degradation (Chakrabarti et al., 2008; Vinu and Madras, 2008;
Yang et al., 2009), the effects of operating parameters (Akpan and Hameed,
2009), photocatalytic carbon–carbon bond formation (Fagnoni et al., 2007),
photochemistry on metal nanoparticles (Watanabe et al., 2006), and solar
photocatalytic detoxification and disinfection of water (Blanco-Galvez et al.,
2007) have been systematically investigated and reviewed.
In the photocatalytic process, anatase titania (TiO2 ) is the most widely
used semiconductor catalyst, and the reaction mechanisms have been dis-
cussed extensively in the literature (Anpo et al., 1985; Bhatkhande et al., 2002;
Chen et al., 2005; Hoffmann et al., 1995; Jaeger and Bard, 1979; Konstantinou
and Albanis, 2004; Martin et al., 1994, 1994b; Rothenberger et al., 1985; Song
et al., 2007b; Yang et al., 2008). Note that to the best of our knowledge, some
Advanced Oxidation Processes for Wastewater Treatment 261

FIGURE 1. Schematic diagram of photocatalytic process acting on the semiconductor photo-


catalyst (Hoffmann et al., 1995; Huang et al., 2009; Linsebigler et al., 1995; Mills and LeHunte,
1997; Rengifo-Herrera et al., 2009) (Color figure available online).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

rate constants and the quantum yield values in this section are not available.
The photoinduced process acting on the semiconductor photocatalyst un-
der ultraviolet or visible light irradiation can be seen in Figure 1 (Hoffmann
et al., 1995; Huang et al., 2009; Linsebigler et al., 1995; Mills and LeHunte,
1997; Rengifo-Herrera et al., 2009). According to Eq. (42), when irradiation
energy (hv) matches or exceeds the band-gap energy of the semiconduc-
tor (E g = 3.2 eV in the case of anatase TiO2 ), electrons (ecb −) are promoted
from the valence band into the conduction band, leaving holes (hvb +) behind
(Bhatkhande et al., 2002; Chen et al., 2005; Fox and Dulay, 1993; Hoffmann
et al., 1995; Mills and LeHunte, 1997; Song et al., 2007b; Yang et al., 2008).
The photogenerated holes that escape direct recombination (Eq. [43]) reach
the surface of TiO2 and react with surface adsorbed hydroxyl groups or wa-
ter to form adsorbed ·OH radicals (·OHads ; Chen et al., 2005; Hoffmann et al.,
1995; Song et al., 2007b). The ·OH radicals that produce at the surface of
semiconductor leave the surface to bulk solution to form free ·OH (·OHfree )
as seen in Eq. (45) (Bhatkhande et al., 2002; Chen et al., 2005; Hoffmann
et al., 1995; Mills and LeHunte, 1997; Song et al., 2007b).

TiO2 + hv → ecb − + hvb + (42)


hvb + + ecb − → heat (43)
+ −
hvb + OHads → ·OHads (44)
+ + +
hvb + H2 O → H2 O → H + ·OHfree (45)

In aerated systems, oxidative species such as O2 ·− and H2 O2 generate from


the reduction site, and radical reactions are given in following reactions
(Al-Ekabi and Serpone, 1988; Chen et al., 2005; Gerischer and Heller, 1991;
262 J. L. Wang and L. J. Xu

Hoffmann et al., 1995; Matthews, 1984; Okamoto et al., 1985a, 1985b; Song
et al., 2007b).

ecb − + O2(ads) → O2 ·− (46)


O2 ·− + ecb − (+2H+ ) → H2 O2 (47)
− −
H2 O2 + O2 · → ·OHads + OH + O2 (48)
− −
H2 O2 + ecb → ·OHads + OH (49)

The enhancement of degradation efficiency by the addition of O3 is attributed


to the reactions with e− and radicals described in Eqs. (21–26) and the
decomposition of O3 by photolysis described in the former section (Agustina
et al., 2005). With the assistance of H2 O2 , the degradation of pollutants is
Downloaded by [Tsinghua University] at 19:33 24 September 2012

reinforced which is due to the electrons on the conduction band initiate


the decomposition of H2 O2 increasing the production of ·OH (Al-Ekabi and
Serpone, 1988; Auguliaro et al., 1990; Okamoto et al., 1985b; Rao and Chu,
2009).

H2 O2 + hv → 2·OH, OH = 0.5 (33)


− −
H2 O2 + ecb → ·OH + OH (49)

An interesting approach should be noted that is the photosensitization


of dyes and pigments under visible light illumination. As shown in Figure 2
(Li et al., 2001; Zhao et al., 2005a; Zhao et al., 1998), the photosensitization
process involves initial excitation of the dye molecules rather than TiO2
particles, which is the difference between the photodegradation of the dye
pollutants under visible light irradiation and that under UV radiation. When
the redox potential of the acceptor dye is lower than the conduction band
of photocatalyst (E cb ∼ –0.5 V vs. NHE in the case of TiO2 ), an electron
is then injected from the excited state into the conduction band, forming

FIGURE 2. The photosensitization pathway of dyes under visible light illumination (Li et al.,
2001; Zhao et al., 2005a; Zhao et al., 1998) (Color figure available online).
Advanced Oxidation Processes for Wastewater Treatment 263

the cationic radical and conduction band electron (Eq. [51]) (Li et al., 2001;
Wu et al., 1999c; Zhao et al., 2005a Zhao et al., 1998). In the presence
of dioxygen, the electron reduces surface adsorbed oxygen, yielding the
oxidizing species such as O2 ·−, HOO·, and ·OH (Li et al., 2001; Wu et al.,
1999c; Zhao et al., 2005a Zhao et al., 1998). When the photodegradation of
an azo dye is conducted in the absence of a catalyst and in the presence of
dissolved oxygen, the possible pathways are homolysis of the excited dye
into radicals, electron transfer of the excited dye to form a radical dye cation,
decomposition by superoxide radical anion, and decomposition by singlet
oxygen (Li et al., 2001; Wu et al., 1999b; Wu et al., 1999c; Zhao et al., 2005a
Zhao et al., 1998).

Dye + hv → Dye∗ (50)


Downloaded by [Tsinghua University] at 19:33 24 September 2012

Dye∗ + TiO2 → Dye+· + TiO2 (e− ) (51)


e− + O 2 → O 2 · − (46)
O2 ·− + H+ → HOO· (52)
HOO· + H+ + e− → H2 O2 (53)
H2 O2 + e− → ·OH + OH− (49)

Recently, more and more studies have focused on the enhancement


of the effectiveness of photocatalysts, such as doping metals, metal ions,
semiconductor oxides, and nonmetal atoms (N, S, I, C), into TiO2 to extend
the absorption spectrum to visible light, slow the recombination rate of the
electron-hole pairs, and enhance the interfacial charge transfer efficiency
(Song et al., 2008a; Zhao et al., 2005a). Hence, the photocatalytic efficiency
under visible light/sunlight illumination can be highly improved, and the
electric power and operating costs can be significantly reduced.
The degradation mechanism of pollutants, especially the role of hvb +,

ecb , ·OHads , and free radicals in photocatalysis, is a very hot topic in photo-
catalysis research. There has been much debate as to whether photoreactions
occurring on the photocatalyst surface involve reactions with positive holes
or surface-bound ·OH or in solution with free radicals (He et al., 2008b; Li
et al., 2008; Song et al., 2007b). Song et al. (2007b) proposed that the decol-
orization of C.I. Reactive Black 5 is primarily initiated by photolysis and/or
O2 ·− in the bulk solution and then the cleavage of the naphthalene and
benzene rings is mainly attributed to the hvb + pathway and ·OHads reactions
on the surface of the catalyst.

2.3 Sonolysis
A large number of studies have been carried out on the use of sonochemi-
cal processes to treat a variety of chemical contaminants mostly in aqueous
systems (Ghodbane and Hamdaoui, 2009; Gogate, 2008; He et al., 2008a;
Sivasankar and Moholkar, 2009b) as well as the basic theory of sonochemical
264 J. L. Wang and L. J. Xu
Downloaded by [Tsinghua University] at 19:33 24 September 2012

FIGURE 3. The formation and collapse of a cavitation bubble, and three reaction zones (i.e.,
cavity interior, gas–liquid interface, and bulk solution) in the cavitation process (Adewuyi,
2001; Chowdhury and Viraraghavan, 2009; Mason and Lorimer, 1988; Suslick, 1989, 1990)
(Color figure available online).

reactions for environmental applications (Adewuyi, 2001; Adewuyi, 2005a,


2005b; Chowdhury and Viraraghavan, 2009; Ensminger, 1973; Fogler and
Timmerhaus, 1966; Hua et al., 1995; Ince et al., 2001; Weavers and Hoffmann,
1998). The formation mechanisms of ·OH under ultrasound (US) irradiation
proceed in the presence of different gases as well as in combination with
other processes (viz. US/O3 , US/H2 O2 , US/O3 /H2 O2 , and US/photocatalysis)
are described subsequently. The rate information in this section is not avail-
able. Sonochemical oxidation techniques involve the use of ultrasonic waves
having frequencies above those within the hearing range of average person
(>16 kHz) to produce an oxidative environment (Adewuyi, 2001).
Sonolysis is principally based on acoustic cavitation including the for-
mation, growth, and implosive collapse of bubbles in a liquid as illustrated
in Figure 3 (Adewuyi, 2001; Chowdhury and Viraraghavan, 2009; Mason and
Lorimer, 1988; Suslick, 1989, 1990). For detailed information of cavitation
formation, readers are requested to refer to the earlier references (Adewuyi,
2001; Chowdhury and Viraraghavan, 2009; Ince et al., 2001; Noltingk and
Neppiras, 1950; Rayleigh, 1917; Serpone and Colarusso, 1994). Positive and
negative pressures are exerted on a liquid by compression and expansion
cycles of ultrasound waves, respectively. When the negative pressure ap-
plied to the liquid is sufficiently large, the average distance between the
molecules would exceed the critical molecular distance necessary to hold the
liquid intact, and the liquid will break apart to form cavities made of vapor-
and gas-filled microbubbles (Adewuyi, 2001; Chowdhury and Viraraghavan,
2009; Ince et al., 2001). Gas and vapors are compressed inside the cavity
Advanced Oxidation Processes for Wastewater Treatment 265

generating heat, which finally produces a short-lived localized hot spot,


creating high local pressures and temperatures (Adewuyi, 2001; Adewuyi,
2005a, 2005b; Chowdhury and Viraraghavan, 2009; Ince et al., 2001).
Among theories of sonochemistry, the hot spot theory is widely accepted
in explaining sonochemical reactions in the environmental field, which sug-
gests that the collapse is so rapid that the compression of the gas and vapor
inside the bubble is adiabatic (Adewuyi, 2001; Adewuyi, 2005a; Chowdhury
and Viraraghavan, 2009; Gogate, 2008; Ince et al., 2001; Ley and Low, 1989;
Noltingk and Neppiras, 1950; Rayleigh, 1917; Serpone and Colarusso, 1994).
The temperatures and pressures within a collapsing microbubble just before
fragmentation can reach values as high as 4200–5000 K and 200–500 atm,
respectively, and consequently a light emission of more than 107 photos per
flash can be achieved (Chowdhury and Viraraghavan, 2009). The localized
Downloaded by [Tsinghua University] at 19:33 24 September 2012

hot spot generated by the rapid collapse of acoustic cavities is very short
lived (<10 µs), implying the existence of extremely high heating and cool-
ing rates in the vicinities of 1010 K/s (Ince et al., 2001). In the structured hot
spot model shown in Figure 3, there are mainly three regions for chemical
reactions: (a) a hot gaseous nucleus in which temperature and pressure are
extremely high, (b) an interfacial region with radical gradient in temperature
and local radical density, and (c) the bulk solution at ambient tempera-
ture (Adewuyi, 2001; Chowdhury and Viraraghavan, 2009; Ghodbane and
Hamdaoui, 2009). The ultrasonic energy influences the chemical reactions
by providing huge heat (pyrolysis) or producing reactive free radicals, and
by increasing the mass transfer rate in an aqueous solution via turbulence
(Adewuyi, 2001, 2005a; Chowdhury and Viraraghavan, 2009). Inside the cav-
itation bubble water molecules are pyrolyzed forming ·OH and ·H radicals in
the gas phase from reaction (54) (Chowdhury and Viraraghavan, 2009; Fis-
cher et al., 1986; Ghodbane and Hamdaoui, 2009; Ince et al., 2001; Serpone
et al., 1994). The substrate either reacts with ·OH or undergoes pyrolysis. In
the interfacial region, a similar reaction occurs but in an aqueous phase, and
the recombination of ·OH radicals to form H2 O2 is the additional reaction
(Eq. [12]) (Chowdhury and Viraraghavan, 2009; Ghodbane and Hamdaoui,
2009; Ince et al., 2001; Serpone et al., 1994). In bulk solution, a small num-
ber of free radicals produced in the cavities or at the interface may move
into the bulk-liquid phase, and the reactions are basically between the sub-
strate and ·OH or H2 O2 (Chowdhury and Viraraghavan, 2009; Ghodbane and
Hamdaoui, 2009).

H2 O → ·OH + ·H (54)
·OH + ·OH → H2 O2 , k = 5.5 × 109 M−1 s−1 (12)

Water sonolysis and sonochemical oxidation and reduction can proceed


in the presence of any gas (Adewuyi, 2001; Entezari and Kruus, 1996; Entezari
et al., 1997; Mead et al., 1976; Sivasankar and Moholkar, 2009a, 2009b). When
under O2 atmosphere, sonochemical activities are improved because thermal
266 J. L. Wang and L. J. Xu

dissociation of oxygen molecule may occur, leading to the generation of


additional ·OH radicals via reactions (55–59) (Adewuyi, 2001; Ince et al.,
2001; Mead et al., 1976; Pétrier et al., 1994). Oxygen can also act as a
scavenger of hydrogen atom suppressing the recombination of ·OH and
·H, and another oxidizing agent HO2 · is additionally formed from reactions
(59) and (2) (Adewuyi, 2001; Sivasankar and Moholkar, 2009a). Moreover,
the higher yield of H2 O2 and O3 takes place as seen in Eqs. (6) and (60)
(Adewuyi, 2001; Sivasankar and Moholkar, 2009b).
O2 → 2 · O (55)
·O + H2 O → 2 · OH (56)
·O + H2 → ·OH + ·H (57)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

·O + HO2 · → ·OH + O2 (58)


·O + H2 O2 → ·OH + HO2 · (59)
·H + O2 → HO2 ·, k = 2.0 × 1010 M−1 s−1 (2)
2HO2 · → H2 O2 + O2 , k = 8.3 × 105 M−1 s−1 (6)
·O + O2 → O3 (60)

When sparging the nitrogen through the medium, nitrogen fixation with
HNO2 as the major acid component formed can occur in the cavity (Adewuyi,
2001; Mead et al., 1976). Nitrogen molecules inside the cavitation bubble may
react at high temperature with hydroxyl radicals to give nitrogen oxide via
Eq. (62) (Adewuyi, 2001; Mead et al., 1976). Nitrogen oxide may undergo
further reactions either in the gas phase of acoustic cavities or in the cooler
interfacial zone with free radicals, ultimately transforming to nitrous and
nitric acids (Eqs. [63–67]) (Adewuyi, 2001; Mead et al., 1976; Sivasankar and
Moholkar, 2009b).

N2 → 2 · N (61)
·N + ·OH → NO + ·H (62)
NO + ·OH → HNO2 (63)
NO + ·OH → NO2 + ·H (64)
NO + ·H → ·N + ·OH (65)
2NO2(aq) + H2 O → HNO2 + HNO3 (66)
NO2 + ·OH → HNO3 (67)

For the air bubble, nitrous oxide formed is unstable under the high-
temperature conditions of the cavitation bubble, which is decomposed in the
gas phase and may be further transformed to NO via Eqs. (68–70) (Adewuyi,
2001; Mead et al., 1976). Nitrogen oxide is additionally formed by reaction
Advanced Oxidation Processes for Wastewater Treatment 267

(71), and finally results in the formation of nitrous and nitric acids as re-
actions (63–67) (Adewuyi, 2001). Under a hydrogen atmosphere, nitrogen
fixation is inhibited due to ·OH is scavenged by H2 from reaction (72), and
oxidation reactions are almost completely suppressed owing to the strong
reducing ability of ·H (Eq. [73]) (Adewuyi, 2001). Likewise, the inhibition
of CO on nitrogen fixation is attributed to the depletion of oxygen in the
oxidation of CO (Eq. [74]) (Adewuyi, 2001). In the presence of CO (or CO2 in
small amount), HCHO can be formed in the cavitation bubble from reactions
(75) and (76) (Adewuyi, 2001).

N2 + ·OH → N2 O + ·H (68)
N2 + ·O → N2 O (69)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

N2 O + ·O → 2NO (70)
N2 + O2 → 2NO (71)
H2 + ·OH → H2 O + ·H (72)
H2 → 2·H (73)
CO + ·O → CO2 (74)
CO2 + ·H → HCOO· (75)
HCOO· + ·H → HCHO + ·O (76)

Furthermore, argon gas penetrating into a cavity can contribute to the transfer
of electron excitation to water (Adewuyi, 2001; Mead et al., 1976; Sivasankar
and Moholkar, 2009a, 2009b). As illustrated in reactions (77–79), the ex-
cited argon atom (Ar∗ ) can participate in energy transfer reactions, ultimately
generating more ·OH (Adewuyi, 2001; Mead et al., 1976).

Ar → Ar∗ (77)
∗ ∗
Ar + H2 O → H2 O +Ar (78)
H2 O∗ → ·H + ·OH (79)

Sivasankar and Moholkar (2009a, 2009b) summarized the simulation results


during bubbling of gases through the aqueous solutions of phenol, nitroben-
zene, p-nitrophenol, and 2,4-dichlorophenol. The results show that the num-
bers of ·OH radicals produced per cavitation bubble (NOH ) are 1.809 × 10−3,
2.838 × 10−4, 2.126 × 10−3, and 1.145 × 10−1 for oxygen, nitrogen, air, and
argon bubbles, respectively. Accordingly, the trend in production of ·OH
radical that is responsible for hydroxylation reaction varies as Ar > Air >
O2 > N2 .
268 J. L. Wang and L. J. Xu

The principle behind the beneficial effects observed using ultrasound


in combination with ozone or hydrogen peroxide as compared to the indi-
vidual application lies in the fact that the rate of generation of free radicals
is significantly enhanced in the case of combination technique (Adewuyi,
2005a, 2005b; Gogate, 2008; Kidak and Ince, 2007), which is very similar
to the UV/O3 or UV/H2 O2 processes as discussed previously, with the only
difference being the energy required for the generation of free radicals from
dissociation of O3 or H2 O2 is given by the cavitating bubbles as against the
UV light. The synergistic effect of combining ozonation with ultrasonic irradi-
ation is observed only when free radical attack is the controlling mechanism
for degradation process (Gogate, 2008). An additional pathway of hydroxyl
radical generation arises upon the decomposition of ozone in the gaseous
bubbles during implosive collapse as given in Eqs. (80) and (81) (Adewuyi,
Downloaded by [Tsinghua University] at 19:33 24 September 2012

2005a, 2005b; Gogate, 2008; Kidak and Ince, 2007).

O3 → O2 + O(3 P) (80)
O( P) + H2 O → 2 · OH
3
(81)

Similarly, in the case of reactions in which the controlling mechanism is the


free radical attack, the combination of cavitation with hydrogen peroxide
should enhance the degradation rates due to the reversible reaction of dis-
sociation of hydrogen peroxide liberating additional free radicals (Gogate,
2008). However, not many studies have been reported in the literature, which
may be attributed to the fact that the concentration of hydrogen peroxide
plays a crucial role in deciding the extent of enhancement obtained for the
combined process, and hydrogen peroxide also acts as the scavenger of the
generated free radicals, as discussed previously in Eq. (18) (Gogate, 2008).
It has also been reported that using a catalyst on the dissociation reaction
of hydrogen peroxide could further improve the combined effect of ultra-
sound and hydrogen peroxide (Gogate, 2008). In addition, the presence of
hydrogen peroxide simultaneously may also further enhance the contribu-
tion of ozone in the overall generation of free radicals by following reactions
(Adewuyi, 2005a; Lesko et al., 2006):

H2 O2 → HO2 − +H+ , pK a = 11.8 (82)


O3 + HO2 − → ·OH + O2 + O2 ·− (83)

The application of ultrasound to photocatalytic oxidation (either using


UV light or solar energy) constitutes a thriving field, including the basic
reaction mechanism (Adewuyi, 2005a, 2005b). Ultrasound plays a profound
role in combination of irradiations (UV and ultrasound) due to substantial
increase in the number of active sites as well as the available surface area
of photocatalyst, defragmentation of the catalyst agglomerates under the
action of turbulence generated by acoustic streaming, and an increase in
Advanced Oxidation Processes for Wastewater Treatment 269

the diffusional rates of the contaminants (Adewuyi, 2005a, 2005b). Thus, the
turbulence induced by the cavitation phenomena can enhance photocatalytic
oxidation, and more free radicals will be available for the reaction, thereby
greatly increasing the rates of reaction.
Many researches have been made on the comparison of single and
combined processes for their effectiveness at degrading organic pollutants
(Kidak and Ince, 2007; Wu, 2009). Kidak and Ince (2007) reported that at
pH 2.0 the pseudo-first-order phenol decay rates was US/UV/O3 > O3 /UV >
US/O3 > US/UV > O3 > US > UV, whereas at pH 10.0 the order was
US/UV/O3 > O3 /UV > O3 > US/O3 > US/UV > UV > US. In the work of
Wu (2009), at pH 7 the decolorization rates of C.I. Reactive Red 2 followed
the order UV/US/TiO2 > UV/TiO2 > US/TiO2 .
Downloaded by [Tsinghua University] at 19:33 24 September 2012

2.4 Electrochemical Oxidation Technologies


2.4.1 ANODIC OXIDATION
Electrochemical oxidation, with the main advantages of environmental com-
patibility, versatility, energy efficiency, and cost-effectiveness, appears as a
promising procedure for removing pollutants from wastewater, which has
become an intense area of research recently (Liu et al., 2009; Martı́nez-Huitle
and Brillas, 2009). The development, design, and application of electrochem-
ical technologies in water and wastewater treatment have been reviewed
(Chen, 2004; Martı́nez-Huitle and Brillas, 2009; Panizza and Cerisola, 2005),
and the oxidation mechanisms of pollutants at anodes have been investigated
(Comninellis, 1994; Liu et al., 2009; Martı́nez-Huitle et al., 2008; Song et al.,
2010; Zhu et al., 2008). In this section, the rate constants are unavailable to
the best of our knowledge.
At the anode, water molecules are oxidized leading to the formation of
radicals, whereas at the cathode hydrogen gas is produced and is not in-
volved in the oxidation of pollutants. Therefore, the electrochemical method
used in the wastewater treatment mainly focuses on anodic oxidation. It has
been shown that the electrode materials play a key role in electrochemical
oxidation by influencing the effectiveness of oxidation, degradation path-
ways, and reaction mechanisms (Zhu et al., 2008). Based on the mechanisms
involved in the pollutant oxidation, the electrode materials have been classi-
fied into two main groups that are active anodes including Pt, IrO2 , and RuO2 ,
and nonactive anodes including PbO2 , SnO2 , and boron-doped diamond
(BDD) (Comninellis, 1994; Martı́nez-Huitle and Brillas, 2009; Martı́nez-Huitle
et al., 2008; Zhu et al., 2008). The initial reaction in both kinds of anodes
(generally denoted as M) corresponds to the oxidation of water molecules
generating physisorbed active oxygen, adsorbed hydroxyl radicals M(·OH),
(Martı́nez-Huitle and Brillas, 2009; Martı́nez-Huitle et al., 2008):

M + H2 O → M(·OH) + H+ + e− (84)
270 J. L. Wang and L. J. Xu

The surface of active anodes interacts strongly with the ·OH radicals,
resulting in the transformation into a higher oxide or superoxide (MO) as
seen in reaction (85) (Comninellis, 1994; Martı́nez-Huitle and Brillas, 2009;
Martı́nez-Huitle et al., 2008; Martı́nez-Huitle and Ferro, 2006). Thus, when
higher oxidation states are available for a metal oxide anode, above the
standard potential for oxygen evolution (E 0 = 1.23 V vs. standard hydrogen
electrode [SHE]), the adsorbed hydroxyl radicals may form chemisorbed ac-
tive oxygen (Martı́nez-Huitle and Brillas, 2009; Martı́nez-Huitle et al., 2008).

M(·OH) → MO + H+ + e− (85)

The MO, which has weaker oxidizing capacity than ·OH, is able to participate
in the oxidation of organic pollutants (R), and the redox couple MO/M acts
Downloaded by [Tsinghua University] at 19:33 24 September 2012

as a mediator in reaction (86). Meanwhile, there exists competition reaction


that is the side reaction of oxygen evolution via chemical decomposition
of MO from reaction (87) (Martı́nez-Huitle and Brillas, 2009; Martı́nez-Huitle
et al., 2008).

MO + R → M + RO (86)
MO → M + (1/2)O2 (87)

In contrast, in a nonactive anode, the surface does not provide any cat-
alytic active site for the adsorption of organics form the aqueous medium,
and it acts only as an electron sink for the removal of electrons (Comninellis,
1994; Martı́nez-Huitle and Brillas, 2009; Martı́nez-Huitle et al., 2008). Hence,
the surface interacts so weakly with ·OH radicals that it allows the direct re-
action of organics with M(·OH) to give fully oxidized reaction products such
as CO2 . In reaction (88), R is an organic compound with m carbon atoms and
2n hydrogen atoms without any heteroatom (Martı́nez-Huitle et al., 2008).
This mineralization reaction with the physisorbed heterogeneous hydroxyl
radical competes with the side reactions of M(·OH) similar to direct oxida-
tion to O2 from reaction (89) or indirect consumption through dimerization
to hydrogen peroxide by reaction (90) (Martı́nez-Huitle and Brillas, 2009;
Martı́nez-Huitle et al., 2008).

M(·OH) + R → M + mCO2 + nH2 O + H+ + e− (88)


M(·OH) → M + (1/2)O2 + H+ + e− (89)
2M(·OH) → 2M + H2 O2 (90)

It should be mentioned that a weaker oxidant as ozone can be generated


from water discharge at the anode (E 0 = –1.51 V vs. SHE) by the overall
reaction (91), and a small amount of hydrogen peroxide can also be pro-
duced as reaction (92) (Michaud et al., 2000; Michaud et al., 2003; Panizza
and Cerisola, 2005). That is to say, besides physisorbed ·OH as the strongest
Advanced Oxidation Processes for Wastewater Treatment 271

oxidant, other reactive oxygen species such as H2 O2 by reactions (90) and


(92), and O3 by reaction (91), are generated in electrochemical oxidation.

3H2 O → O3 + 6H+ + 6e− (91)


2H2 O → H2 O2 + 2H+ + 2e− (92)

Besides the electrode materials, the efficiency of the electrochemical


oxidation is also in direct relation to supporting medium. When sulfate (or
bisulfate), carbonate (or bicarbonate) and phosphate are employed as elec-
trolytes, other weaker oxidizing species such as peroxodisulfate, peroxodi-
carbonate, and peroxodiphosphate can also be formed in anodic oxidation
(Martı́nez-Huitle and Brillas, 2009; Michaud et al., 2000; Michaud et al., 2003;
Serrano et al., 2002).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

2HSO4 − → S2 O8 2− + 2H+ + 2e− (93)



2SO4 2−
→ S2 O8 2−
+ 2e (94)
2HCO3 − → C2 O6 2− + 2H+ + 2e− (95)
2PO4 3− → P2 O8 4− + 2e− (96)

During all the supporting electrolytes, it is important to highlight the role


of chloride because the degradation efficiency of organics may also be
improved using chloride-mediated wastewaters. This is attributed to the gen-
eration of soluble chlorine, trichloride ion, or hypochlorite from chloride ions
(Eqs. [97–100]) or by the reaction of physisorbed ·OH with chloride ions (Eq.
[101]) (Carvalho et al., 2007; del Rı́o et al., 2009; Martı́nez-Huitle and Brillas,
2009; Panizza et al., 2005; Szpyrkowicz et al., 2000). Martı́nez-Huitle and Bril-
las (2009) illustrated the formation of active chlorine species (Cl2(aq) , Cl3 −,
HClO, and ClO−) at various pH values, which shows that Cl3 − is formed
in very low concentration up to pH 4, whereas the predominant species is
Cl2(aq) until pH near 3, HClO in the pH range 3–8, and ClO− for pH > 8.
Because of the higher standard potential of Cl2(aq) (E 0 = 1.36 V vs. SHE) and
HClO (E 0 = 1.49 V vs. SHE) than ClO− (E 0 = 0.89 V vs. SHE), the mediated
oxidation of organics with these species is then expected to be faster in
acidic than in alkaline media.

Anode: 2Cl− → Cl2(aq) + 2e− (97)


Bulk solution: Cl2(aq) + Cl− ↔ Cl3 − (98)
+ −
Cl2(aq) + H2 O → HClO + H + Cl (99)
+ −
HClO ↔ H + ClO , pK a = 7.55 (100)
Cl− + ·OH + H+ → HClO + H2 O (101)
272 J. L. Wang and L. J. Xu
Downloaded by [Tsinghua University] at 19:33 24 September 2012

FIGURE 4. Oxidation mechanisms of organic pollutants at active anodes (dashed line frame),
and at nonactive anodes (Zhu et al., 2008; Liu et al., 2009) (Color figure available online).

In summary, oxidation mechanisms of organic pollutants at nonactive


anodes are elucidated in Figure 4 (Liu et al., 2009; Zhu et al., 2008), and
mechanisms at active anodes can be seen in dashed line frame. The main
approaches for the pollution abatement in wastewaters at active anodes are
(a) direct anodic oxidation (or direct electron transfer to the anode), which
yields very poor decontamination; (b) chemical reaction with chemisorbed
active oxygen (oxygen in the lattice of a metal oxide [MO] anode), leading
to partial decontamination which is due to the surface redox couple MO/M
is much more selective than the hydroxyl radical; and (c) indirect electro-
chemical oxidation mediated by electrogenerated oxidants, such as perox-
odisulfates (in the presence of SO4 2−) and active chlorine (in the presence
of Cl−). Meanwhile, the degradation of pollutants at nonactive anodes may
be attributed to (a) direct electrochemical oxidation on the anode surface,
(b) indirect electrochemical oxidation mediated by electrogenerated oxidants
(e.g., S2 O8 2−, H2 O2 , Cl2 , HClO, ClO−), and (c) indirect electrochemical ox-
idation mediated by hydroxyl radicals leading to total decontamination of
organics.

2.4.2 PHOTOELECTROCATALYSIS
Photoelectrocatalysis, an emerging technology in which a TiO2 -based thin
film anode is irradiated with an UV light, has recently received great attention
for wastewater remediation (Gao et al., 2008; Kim et al., 2009; Martı́nez-
Huitle and Brillas, 2009). We discussed the mechanisms of photocatalysis
and anodic oxidation in Sections 2.2.2 and 2.4.1, and the formation of ·OH
was presented in Eqs. (44–49) and Eq. (84).
Advanced Oxidation Processes for Wastewater Treatment 273

With the assistance of photocatalysis, the electrochemical technology


can provide much higher efficiency for wastewater remediation. The pho-
toinduced electrons are continuously extracted from the anode by an ex-
ternal circuit, which causes the suppression of electron–hole recombination
(Eq. [43]) and the inhibition of electron-mediated reactions (Eqs. [46–49])
(Martı́nez-Huitle and Brillas, 2009). Simultaneously, the generation of higher
amount of holes from reaction (42) and ·OH from reactions (44) and (45) is
favored, or photogenerated holes acquire longer lifetime promoting redox
reactions (Gao et al., 2008). Thus, photoelectrocatalysis largely enhances the
oxidation of pollutants in comparison with photocatalysis.
When anodic potential is low, the applied bias is not intended to create
electrochemical oxidation of pollutants. However, the low bias is efficiently
high enough to inhibit the recombination of holes and electrons efficiently,
Downloaded by [Tsinghua University] at 19:33 24 September 2012

so the effect of anodic potential on photoelectrons is dominating (Martı́nez-


Huitle and Brillas, 2009). As anodic potential increases, a large amount of
current carrier (photoelectrons) passes through the anode, and photocurrent
is saturated when the transportation and creation of photoelectrons reaches
equilibrium (Kim et al., 2009). Therefore, synergistic effects are obtained
and lower current density is supplied in photoelectrocatalysis, reducing the
energy required to operate the purely electrochemical system (Malpass et al.,
2007).
Recent researches on removing pollutants by photoelectrocatalysis have
demonstrated its better performance than the corresponding photocatalysis
and electrochemical oxidation (Asmussen et al., 2009; Gao et al., 2008).
Asmussen et al. (2009) used TiO2 /Ti/Ta2 O5 –IrO2 bifunctional electrodes
for photoelectrochemical oxidation of 4-nitrophenol and 2-nitrophenol. The
results show that in the degradation of 4-nitrophenol a rate constant of
1.06 × 10−2 min−1 was created by photoelectrocatalysis, which was much
higher than the degradation rate produced by the individual photochemical
oxidation (1.11 × 10−4 min−1) and that by the electrochemical oxidation
(5.74 × 10−3 min−1).

2.5 Fenton-Based Reactions


Various reviews have maintained attention on the complex mechanism of
Fenton and Fenton-like reactions (Masarwa et al., 2005; Pignatello et al.,
2006), the important factors influencing the efficiency of these processes
in applications to water and soil treatment (Neyens and Baeyens, 2003;
Pignatello et al., 2006), the progress of all electrochemical technologies based
on Fenton’s reaction chemistry (Brillas et al., 2009), and the application of
Fenton oxidation to the treatment of industrial wastewaters (Bautista et al.,
2008). Fenton and related reactions, viewed as potentially convenient and
economical techniques, encompass reactions of peroxides (usually H2 O2 ) or
dissolved oxygen with iron ions to generate oxidizing species for treating
274 J. L. Wang and L. J. Xu

organic or inorganic pollutants (Pignatello et al., 2006). Hydroxyl radical


production is thought to be the key step in the classical or free radical
Fenton chain reaction, and the sequence of reactions is discussed in detail
subsequently. It should be noticed that some rate constants are not supplied
because, as far as we know, they are not available.

2.5.1 FENTON PROCESS


The generally accepted mechanism of homogeneous Fenton process, initi-
ated by the generation of hydroxyl radicals in acidic medium, is a number of
cyclic reactions, which utilize the ferrous or ferric ions as a catalyst to decom-
pose the H2 O2 (Eqs. [102] and [103]) (Barb et al., 1949, 1951a, 1951b; Bautista
et al., 2008; Brillas et al., 2009; Haber and Weiss, 1934; Pignatello et al., 2006).
The rate constants are determined at 298 K and I = 0.1 M (HClO4 /NaClO4
Downloaded by [Tsinghua University] at 19:33 24 September 2012

solution) for a second-order reaction rate (Brillas et al., 2009; De Laat and
Gallard, 1999; Fan et al., 2009). The Fenton reaction (102) is propagated by
Fe2+ regeneration, which can take place by the reduction of Fe3+ with H2 O2
from reaction (103), HO2 · from reaction (106), an organic radical R· from
reaction (107), or O2 ·− from reaction (108) (Barb et al., 1949, 1951a; Brillas
et al., 2009; Haber and Weiss, 1934; Rothschild and Allen, 1958). Reaction
(103) is associated with a two-step transformation in which Fe(III)-peroxo
complexes formed in the equilibrium reaction (104) further decompose into
Fe2+ and HO2 · following the reaction (105) (Barb et al., 1949, 1951b; Brillas
et al., 2009).

Fe2+ + H2 O2 → Fe3+ + ·OH + OH− , k = 63 M−1 s−1 (102)


Fe3+ + H2 O2 → Fe2+ + HO2 · + H+ , k = 0.002–0.01 M−1 s−1 (103)
Fe3+ + H2 O2 ↔ Fe(OOH)2+ + H+ , k = 3.1 × 10−3 M−1 s−1 (104)
−3 −1 −1
Fe(OOH) 2+
→ Fe 2+
+ HO2 ·, k = 2.7 × 10 M s (105)
+ −1 −1
Fe 3+
+ HO2 · → Fe 2+
+ O2 + H , k = 2.0 × 10 M s 3
(106)
Fe3+ + R· → Fe2+ + R+ (107)
Fe3+ + O2 ·− → Fe2+ + O2 , k = 5.0 × 107 M−1 s−1 (108)

The decomposition of organic pollutants (RH) by ·OH principally via ab-


stracting H from C–H, N–H or O–H bonds (Eq. [109]), and adding to C C
bonds or aromatic rings, depending on the ionization potential of organic
pollutants (Brillas et al., 2009; Legrini et al., 1993; Pignatello et al., 2006;
Walling, 1975). The produced intermediate organic radicals (R·) can react
with Fe3+ and H2 O2 , forming R+ and ROH (Eqs. [107], [110] and [111]) that
can be further oxidized (Bautista et al., 2008; Brillas et al., 2009; Neyens
and Baeyens, 2003; Walling, 1975). In the presence of oxygen, R· radicals
may react with O2 to give HO2 · (Eq. [112]), peroxyl radicals (ROO·), or oxyl
Advanced Oxidation Processes for Wastewater Treatment 275

radicals (RO·), as seen in Eq. (113), which can be ultimately degraded into
CO2 , H2 O, and organic acids (Du et al., 2007b; Pignatello et al., 2006). Nev-
ertheless, a number of competitive reactions occurred, as seen in reactions
(114–117), can negatively affect the oxidation process (Bautista et al., 2008;
Brillas et al., 2009; Du et al., 2007a). Among them, reactions (114) and (115),
promoting the consumption of ·OH by the Fenton’s reagent, are the major
waste reactions (Brillas et al., 2009).

·OH + RH → H2 O + R·, k = 107 −109 M−1 s−1 (109)


R· + Fe3+ → R+ + Fe2+ (107)
+ −
R + OH → ROH (110)
R· + H2 O2 → ROH + ·OH (111)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

R· + O2 → → R(−H+ ) + HO2 · (112)


R· + O2 → ROO· →→ RO· (113)
− −1 −1
·OH + Fe 2+
→ OH + Fe , 3+
k = 3.2 × 10 M s 8
(114)
·OH + H2 O2 → H2 O + HO2 ·, k = 2.7 × 107 M−1 s−1 (115)
·OH + HO2 · → H2 O + O2 , k = 1.0 × 1010 M−1 s−1 (116)
−1 −1
·OH + ·OH → H2 O2 , k = 4.2 × 10 M s
9
(117)

The significance of following reactions and reactions (116) and (117) is


relatively low because of the poor concentrations of radical species (e.g.,
HO2 ·, O2 ·−) in the bulk solution, despite the quite high rate constants of
these reactions (Brillas et al., 2009).

Fe2+ + HO2 · + H+ → Fe3+ + H2 O2 , k = 1.2 × 106 M−1 s−1 (118)


Fe2+ + O2 ·− + 2H+ → Fe3+ + H2 O2 , k = 1.0 × 107 M−1 s−1 (119)
− + −1 −1
O2 · + HO2 · + H → H2 O2 + O2 , k = 9.7 × 10 M s
7
(120)
HO2 · + HO2 · → H2 O2 + O2 , k = 8.3 × 105 M−1 s−1 (6)
− − −1 −1
O2 · + ·OH → OH + O2 , k = 1.01 × 10 M s 10
(121)

It is worth noting that hypervalent iron complexes (e.g., ferryl or Fe(IV)


ions) have been found under certain operating conditions in the Fenton pro-
cess, which can only oxidize organic molecules by electron transfer (Boss-
mann et al., 1998). For example, previous studies have provided evidence
for the production of an Fe(IV) species by the Fenton reaction, the photo-
Fenton reaction, the reaction of nanoparticulate zero-valent iron and oxygen
at neutral pH values, and the Fenton reaction in the presence of ligands
(e.g., ethylenediaminetetraacetic acid, polyoxometalate; Keenan and Sedlak,
2008b, 2008c; Lee et al., 2008a). The formation of a mononuclear Fe(IV) oxo
276 J. L. Wang and L. J. Xu

complex has been proposed in reaction (122) (Brillas et al., 2009; Pignatello
et al., 1999).

Fe2+ + H2 O2 → [Fe(OH)2 ]2+ → Fe3+ + ·OH + OH− (122)

Many researches utilize Fe3+ instead of Fe2+ to decompose H2 O2 as


Fenton-like reactions (Barb et al., 1949, 1951b; De Laat and Le, 2006; Fan
et al., 2009; Siedlecka et al., 2008). From the previous reactions, it can be seen
that both ferrous and ferric species are present simultaneously, regardless of
which is used to initiate the reaction. Therefore, in this review we consider
reactions initiated with Fe2+ and Fe3+ as homogeneous Fenton processes.
For the application of homogeneous Fenton processes, there are some disad-
vantages, including the requirements of further removal for the iron sludge,
Downloaded by [Tsinghua University] at 19:33 24 September 2012

the relatively high cost and risks related to the storage and transportation of
H2 O2 , and the acidification of effluents at pH 2–4 before decontamination or
the neutralization of treated solutions before disposal (Brillas et al., 2009).
To overcome these disadvantages of homogeneous Fenton process,
solid iron-containing catalysts have been used in heterogeneous Fenton pro-
cesses, which have recently received much attention (Deng et al., 2008;
Keenan and Sedlak, 2008c; Lee et al., 2008a). The investigations have mainly
been focused on zero-valent iron; iron oxides, such as Fe2 O3 , Fe3 O4 and
Fe0/Fe3 O4 ; iron-immobilized materials; and natural iron-containing materi-
als, such as goethite and magnetite (Deng et al., 2008; Lee et al., 2008a).
Oxygen instead of H2 O2 can be used to react with solid iron-containing
catalysts, leading to ·OH induced oxidation. The mechanism of zero-valent
iron oxidation by dissolved oxygen via electron transfer from the particle
surface to oxygen has been proposed (Keenan and Sedlak, 2008b, 2008c;
Lee et al., 2008a; Lee and Sedlak, 2008). The H2 O2 produced by reaction
(123) is either reduced to H2 O (Eq. [124]) or converted to ·OH (Eq. [102])
or Fe(IV) (Eq. [125]) (Lee et al., 2008a; Lee and Sedlak, 2008). Under neu-
tral pH conditions, the oxidation of Fe2+ by oxygen also produces H2 O2
via reactions (126) and (119), which subsequently yields oxidants through
the homogeneous Fenton reaction (Lee et al., 2008a; Lee and Sedlak, 2008).
In general, the solid iron-containing catalysts are first transferred into Fe2+
and Fe3+ ions in the presence of H2 O2 or O2 , and then generate oxidizing
species via the homogeneous Fenton reaction. Thus, oxidation of pollutants
may potentially occur via iron ions released into solution or via reactions
that take place between solutes and surface-bound species.

Fe0 + O2 + 2H+ → Fe2+ + H2 O2 (123)


Fe0 + H2 O2 + 2H+ → Fe2+ + 2H2 O (124)
− −1 −1
Fe 2+
+ H2 O2 → Fe 3+
+ ·OH + OH , k = 63 M s (102)
Fe 2+
+ H2 O2 → Fe(IV) (e.g., FeO ) + H2 O
2+
(125)
Advanced Oxidation Processes for Wastewater Treatment 277

Fe2+ + O2 → Fe3+ + O2 ·− , k = 5.0 × 107 M−1 s−1 (126)

Fe2+ + O2 ·− + 2H+ → Fe3+ + H2 O2 , k = 1.0 × 107 M−1 s−1 (119)

It is noteworthy that the form of iron species and their catalytic activities
are mainly affected by the solution pH (Brillas et al., 2009; Katsoyiannis et al.,
2008). At pH 0–2.5, the sole species is [Fe(H2 O)6 ]3+ or simply Fe3+. At pH 2.8
the concentration of Fe2+ available in the reaction medium is at its maximum,
which gives rise to the highest rate of Fenton’s reaction (102), and Fe3+ is
always present in the bulk (Brillas et al., 2009). At pH 5, [Fe(OH)2 ]+ is the
main species and Fe(III) species precipitate as Fe(OH)3 more quickly at
higher pH values, thereby decreasing the quantity of catalyst in the solution
and breaking the H2 O2 into O2 and H2 O (Brillas et al., 2009; Katsoyiannis
Downloaded by [Tsinghua University] at 19:33 24 September 2012

et al., 2008). Therefore, the optimum pH range in the homogeneous Fenton


reaction is 2–4, which needs a strict pH control, whereas the heterogeneous
Fenton process possesses a wider applicable pH range.
The nature of the reactive oxidant produced by the Fenton reaction,
·OH versus Fe(IV), has been a controversial issue for decades (Brillas et al.,
2009; Jiang et al., 2008a, 2008b; Katsoyiannis et al., 2008, 2009; Keenan and
Sedlak, 2008a, 2008b; Lee et al., 2008a, 2008b; Noubactep, 2009). Keenan
and Sedlak (2008b) studied the oxidation of methanol, ethanol, benzoic acid,
and 2-propanol by the reaction of nanoparticulate zero-valent iron with O2 in
the absence of ligands. They tested the hypothesis that the nature of the ox-
idants changed with pH from ·OH at low pH to another less reactive species
(e.g., the ferryl ion) at neutral pH values. Nevertheless, Jiang et al. (2008a)
suggested that the results obtained in the experiment might be as such be-
cause the reactivity of Fe(IV) with the probe compounds decreased with
the increase of pH. Again, Lee et al. (2008a) concluded that at neutral pH
the active oxidant changed from Fe(IV) in the absence of polyoxometalate
(POM) to ·OH in the presence of POM for the oxidation of organic com-
pounds by the Fenton process. However, Jiang et al. (2008b) argued that the
reactivity of Fe(IV) toward organics might be enhanced in the presence of
POM. Anyway, despite being a controversial question, there is a consensus
that the hydroxyl radicals mechanism and Fe(IV)-based mechanism coexist
and predominate one or the other depending on the operation conditions
such as pH and the presence of ligands.
Another possibility to overcome the disadvantages of Fenton process
and to enhance the degradation efficiency is the use of hybrid processes
such as photo-Fenton, sono-Fenton, and electro-Fenton processes, which is
detailed in subsequent sections.
2.5.2 PHOTO-FENTON PROCESS
The Fenton process assisted by ultraviolet or visible light irradiation, namely
the photo-Fenton process, invariably leads to the increase of degradation
278 J. L. Wang and L. J. Xu

efficiency (Brillas et al., 2009; Ntampegliotis et al., 2006; Pignatello et al.,


2006). In addition to the reactions described in Section 2.5.1, the following
processes are occurring synchronously. Photolysis of H2 O2 from reaction
(33) occurs using UV light (λ < 285 nm) because H2 O2 has a maximum ab-
sorbance at 210–230 nm (Chevaldonnet et al., 1986; Elmolla and Chaudhuri,
2009).

H2 O2 + hv → 2·OH, OH = 0.5 (33)

Photoreduction of aqueous ferric ions (Eq. [127]) also takes place under
wavelength 365 nm, which reproduces Fe2+, enhancing the production of
·OH from reaction (102) (Elmolla and Chaudhuri, 2009; Ntampegliotis et al.,
2006; Safarzadeh-Amiri et al., 1996).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Fe3+ + hv ↔ Fe2+ (127)

The photolysis of ferric iron/aquo complexes that have adsorption bands in


the UV–Vis range (λ < ca. 450 nm) can both efficiently regenerate Fe2+ and
further produce ·OH via reaction (128) (Faust and Hoigné, 1990; Lapertot
et al., 2006). The monohydroxyl complex, Fe(OH)2+, predominant in acidic
conditions, is the most photoactive species. The quantum yield for Fe2+
formation by the photolysis of Fe(OH)2+ is wavelength dependent, which is
0.14–0.19 at 313 nm and 0.017 at 360 nm (Pignatello et al., 2006).

[Fe3+ (OH− )x (H2 O) y ] + hv → Fe2+ + (x − 1)OH− + yH2 O + ·OH (128)

Under visible light irradiation, much faster degradation and mineralization of


various dyes has been observed in the photo-Fenton reaction due to effective
electron transfer from the visible light-excited dyes into Fe3+, leading to the
catalytic cycling of Fe3+/Fe2+ (Eqs. [50] and [129]) (Ma et al., 2005; Wu et al.,
1999a).

Dye + hv → Dye∗ (50)


Dye∗ + Fe3+ → Fe2+ + Dye+ · (129)

In addition, UV or visible light can induce the photodegradation of certain


target compounds, their by-products, or their complexes with Fe(III) promot-
ing Fe2+ regeneration. This is the case for the photoreduction of ferrioxalate
complexes according to reaction (130) (Hatchard and Parker, 1956; Wang
et al., 2008; Wu et al., 1999a).

[Fe(C2 O4 )n]3−2n + hv → [FeII (C2 O4 )n−1 ]4−2n + CO2 + CO2 ·− (130)

Solid iron-containing photocatalysts, such as iron (hydr)oxides (Du et al.,


2008) and Fe(III)-Al2 O3 (Muthuvel and Swaminathan, 2008), have been used
in the heterogeneous photo-Fenton process, which has been discussed with
Advanced Oxidation Processes for Wastewater Treatment 279

respect to the mechanisms of photocatalysis and the Fenton process. As


an example, the synergistic effects of iron oxide and TiO2 on the polymer
film are observed during the heterogeneous photo-Fenton degradation of
hydroquinone due to the photocatalysis of iron oxide and TiO2 (described
in Section 2.2.2) and the photo-Fenton reactions (described in Section 2.5.1
and the present section; Mazille et al., 2009).
2.5.3 SONO-FENTON PROCESS
In the case of ultrasound coupled with Fenton process, the degradation
of pollutants is accelerated due to the following reactions (Bremner et al.,
2009; Chakinala et al., 2009; Pradhan and Gogate, 2010). The decomposition
of H2 O2 under ultrasonic irradiation generates a higher concentration of
·OH according to the reaction (131) (Chakinala et al., 2009; Pradhan and
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Gogate, 2010). Besides the Fenton reactions described in Section 2.5.1, the
complex Fe(OOH)2+ can be effectively dissociated into Fe2+ and HO2 · under
ultrasonic irradiation (Eq. [132]), which makes more Fe2+ available in the
solution giving rise to the higher rate of Fenton’s reaction (102) (Chakinala
et al., 2009; Pradhan and Gogate, 2010). Moreover, in heterogeneous sono-
Fenton process, acoustic cavitations can enhance mass transfer, dispersion,
and disaggregation of catalyst particles, and increase the surface defects or
active sites on the catalyst surface (Bremner et al., 2009).

H2 O2 → 2·OH (131)
Fe + H2 O2 → Fe3+ + ·OH + OH− , k = 63 M−1 s−1
2+
(102)
Fe3+ + H2 O2 ↔ Fe(OOH)2+ + H+ , k = 3.1 × 10−3 M−1 s−1 (104)
Fe(OOH)2+ → Fe2+ + HO2 ·, k = 2.7 × 10−3 M−1 s−1 (132)

2.5.4 ELECTRO-FENTON PROCESS


The electro-Fenton technology is based on the electrochemical reactions that
are used to generate in situ one or both of the Fenton reagents (H2 O2 and
Fe2+) throughout the process (Brillas et al., 2009). The reagents generated
depend on cell configuration, the nature of the electrodes, applied potential
or current, and solution conditions (Brillas et al., 2009; El-Desoky et al., 2010;
Pignatello et al., 2006).
Most studies focus on the continuous electrogeneration of H2 O2 at a
suitable cathode from the two-electron reduction of oxygen gas directly
injected as pure gas or bubbled air, as expressed in reaction (133) with E 0 =
0.695 V/SHE (Brillas et al., 2000; Chen and Liang, 2008; Khataee et al., 2009;
Liu et al., 2007; Sudoh et al., 1986). The H2 O2 generated in acidic solution
reacts with the added Fe2+ or iron catalysts to produce ·OH via Fenton’s
reaction (102).

O2 + 2H+ + 2e− → H2 O2 (133)


280 J. L. Wang and L. J. Xu

Ferrous ions can be electrochemically regenerated by cathodically re-


duction of ferric ions in reaction (134), with E 0 = 0.77 V/SHE, which accel-
erates the production of ·OH from Fenton’s reaction (102) and minimizes
the quantity of iron sludge (Brillas et al., 2009; Chou et al., 1999; Hsiao and
Nobe, 1993; Zhang et al., 2009).

Fe3+ + e− → Fe2+ (134)

Soluble Fe2+ can also be supplied by oxidative dissolution of sacrificial an-


odes such as iron metal or titanium and iron via reaction (135), with E 0 =
–0.44 V/SHE, following a faradaic behavior (Brillas et al., 2009; Huang et al.,
1999; Pignatello et al., 2006; Pratap and Lemley, 1994).

Fe0 → Fe2+ + 2e−


Downloaded by [Tsinghua University] at 19:33 24 September 2012

(135)

Other mineral iron oxides such as wüstite (FeO), magnetite (Fe3 O4 ), and
hematite (α-Fe2 O3 ) are also used to release Fe2+ to the bulk to react with
electrogenerated H2 O2 via Fenton’s reaction (102) (Brillas et al., 2009). Fur-
thermore, in an undivided cell, pollutants are destroyed not only by ·OH
produced homogeneously from Fenton’s reaction (102) described previously
but also by the action of heterogeneously formed hydroxyl radical, M(·OH),
described in Section 2.4.1 (Brillas et al., 2009).
Recently, there is increasing interest in the use of the photoelectro-
Fenton method for wastewater treatment (Brillas et al., 2009; Casado et al.,
2005; Li and Qu, 2009; Skoumal et al., 2009). The solution is treated under
electro-Fenton conditions and irradiated with UV or visible light during or
after electrolysis to accelerate the mineralization rate of pollutants (Casado
et al., 2005). The synergistic effect can be associated with (a) the regeneration
of more Fe2+ and production of additional ·OH from the photoreduction of
Fe(OH)2+ that is the predominant Fe(III) species in acid medium (Eq. [136]),
and (b) the photolysis of complexes of Fe(III) with regenerated Fe2+ by
reaction (130) (Brillas et al., 2009; Faust and Hoigné, 1990; Li and Qu, 2009;
Sun and Pignatello, 1993).

Fe(OH)2+ + hv → Fe2+ + ·OH (136)

The electro-Fenton process can also be improved by ultrasounds, and


is called the sonoelectro-Fenton process (Brillas et al., 2009). Cavitation can
accelerate the mineralization process by the additional generation of ·OH
by sonolysis from reaction (54), the enhancement of mass transfer, and the
pyrolysis of organics.
The degradation of pollutants by various processes has been compar-
atively studied (Esquivel et al., 2009; Skoumal et al., 2009). Esquivel et al.
(2009) investigated the TOC removal of azo dye Orange II in the differ-
ent electrochemical advanced oxidation processes, and the mineralization
Advanced Oxidation Processes for Wastewater Treatment 281

extents after 60 min treatments followed the order electro-Fenton photo-


electrocatalytic oxidation > electro-Fenton > electro-assisted photocatalytic
oxidation > anodic electrooxidation.

2.6 Ozone-Based Processes


Processes based on ozone are likely promising for efficient treatment of
wastewater-containing recalcitrant organic compounds, such as chlorophe-
nols, organic pesticides, and pharmaceuticals, as seen in many reviews
(Agustina et al., 2005; Esplugas et al., 2007; Ikehata and El-Din, 2005a,
2005b; Ikehata et al., 2006). The oxidation mechanisms of pollutants using
ozonation, catalytic ozonation (including homogeneous and heterogeneous
catalytic ozonation), and hybrid processes (e.g., radiolysis/O3 , UV/O3 and
Downloaded by [Tsinghua University] at 19:33 24 September 2012

US/O3 ) are discussed in following sections.


2.6.1 OZONATION
Ozone is a very selective oxidant with standard potentials of 2.07 V (vs. NHE)
in acidic solution and 1.25 V (vs. NHE) in basic solution, and has recently
received much attention in water treatment and purification (Hoigné and
Bader, 1979, 1983a, 1983b; Hoigné et al., 1985; Staehelin and Hoigné, 1985;
Hoigné, 1998; Kishimoto et al., 2005; Pi et al., 2007). There are two oxidation
mechanisms, namely direct electrophilic attack by molecular ozone and in-
direct attack through the formation of hydroxyl radicals (Hoigné and Bader,
1975, 1979; Pera-Titus et al., 2004; Staehelin and Hoigné, 1982, 1985). The
pH of the solution is a major factor determining the efficiency of ozonation
because it can alter the kinetics and pathways of the reaction. At low pH,
the direct ozonation dominates and is selective, whereas at basic conditions,
the indirect pathway prevails (Agustina et al., 2005; Pera-Titus et al., 2004;
Staehelin and Hoigné, 1985). The formation of ·OH and relevant radical re-
actions are very complex and influenced by many substances. In ozonation
of water, ·OH is generated as seen in the SBH model (Bühler et al., 1984;
Staehelin and Hoigné, 1982, 1985) for the neutral pH region or the TFG
model (Nemes et al., 2000a, 2000b; Tomiyasu et al., 1985) for a higher pH
range, and the rate constants were determined by Chelkowska et al. (1992).
Reaction (137) corresponds to the formation of one O2 ·− anion and one
HO2 · radical, whereas reaction (141) corresponds to a two-electron transfer
or an oxygen atom transfer reaction. No matter the ozone decomposition
process is initiated by reactions (137) or (141), ·OH is produced by means
of reaction between O3 and OH− ions.

SBH model: O3 + OH− → O2 ·− + HO2 ·, k = 140 M−1 s−1 (137)


− − −1 −1
O2 · + O 3 → O3 · + O 2 , k = 1.6 × 10 M 9
s (138)
− +
O3 · + H ↔ HO3 ·, pK a = 6.2 (139)
282 J. L. Wang and L. J. Xu

HO3 · → ·OH + O2 , k = 5.0 × 104 M−1 s−1 (140)


TFG model: O3 + OH− → HO2 − + O2 , k = 120 M−1 s−1 (141)
HO2 − + O3 → O3 ·− + HO2 ·, k = 1.5 × 106 M−1 s−1 (142)
O2 ·− + O3 → O3 ·− + O2 , k = 1.6 × 109 M−1 s−1 (143)
− − −1 −1
O3 · + H2 O → ·OH + O2 + OH , k = 15 M s (144)

Moreover, at extreme conditions of very high pH values, a direct one-electron


transfer process is observed (Staehelin and Hoigné, 1982; Tomiyasu et al.,
1985).

O3 + OH− → O3 ·− + ·OH, k = 70 ± 7 M−1 s−1 (145)


Downloaded by [Tsinghua University] at 19:33 24 September 2012

O3 ·− + H2 O → ·OH + O2 + OH− , k = 15 M−1 s−1 (144)

The decomposition of ozone in basic aqueous solutions has been elaborated


in light of the detailed reactions shown in literature (Chelkowska et al., 1992).
It is indicated that the combination of ozone with hydrogen perox-
ide (O3 /H2 O2 ) can improve the degradation of organic compounds (Adams
et al., 1994; Ormad et al., 1997; Pera-Titus et al., 2004). In addition to the
mechanism of ozonation exposed former, additional ·OH is generated from
the reaction of ozone with anion HO2 − via reactions (82) and (83), and the
direct reaction of ozone with the undissociated H2 O2 is negligible due to a
very low kinetic constant, as seen in reaction (146) (Pera-Titus et al., 2004;
Staehelin and Hoigné, 1982).

O3 + H2 O2 → H2 O + O2 , k< 0.01 M−1 s−1 (146)

However, ozonation has been shown to achieve a very limited min-


eralization of organic compounds due to the relatively low solubility and
stability of ozone in water and the slow reaction with some organic com-
pounds (Kasprzyk-Hordern et al., 2003). Recently, a research focus has been
laid on the catalytic ozonation processes and hybrid processes to enhance
the ·OH production and the degradation efficiency.
2.6.2 CATALYTIC OZONATION
More recently, catalytic ozonation has become an attractive and increasingly
important research field in the use of ozone (Kasprzyk-Hordern et al., 2003;
Pirgalioğlu and Özbelge, 2009; Wu et al., 2008; Zhao et al., 2009d). It can
be considered as homogeneous catalytic ozonation that is based on ozone
activation by metal ions present in aqueous solution, and as heterogeneous
catalytic ozonation in the presence of metal oxides or metals/metal oxide on
supports (Kasprzyk-Hordern et al., 2003; Pirgalioğlu and Özbelge, 2009). It
should be noticed that in this section some rate constants are not available.
Advanced Oxidation Processes for Wastewater Treatment 283

In homogeneous catalytic ozonation studies, the catalysts proposed are


transition metals such as Fe(II), Mn(II), Ni(II), Co(II), Cd(II), Cu(II), Ag(I),
Cr(III), and Zn(II) (Abd El-Raady et al., 2005; Kasprzyk-Hordern et al., 2003;
Logager et al., 1992; Ma and Graham, 1997; Pirgalioğlu and Özbelge, 2009).
The mechanism of homogeneous catalytic ozonation is based on an ozone
decomposition reaction initiated by metal ions present in the bulk solution
following by the generation of hydroxyl radicals (Kasprzyk-Hordern et al.,
2003). Possible pathways of ·OH production for O3 /Mn(II) have been pro-
posed as seen in reaction (147) (Ma and Graham, 1997; Wu et al., 2008). In
O3 /Fe(II) and O3 /Fe(III) systems, an ozone decomposition reaction followed
by the generation of radicals is observed as reactions (148–152) (Li et al.,
2006; Logager et al., 1992; Piera et al., 2000; Wu et al., 2008).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Mn2+ + O3 + H+ → Mn3+ + ·OH + O2 (147)



Fe 2+
+ O3 → Fe 3+
+ O3 · (148)
O3 ·− + H+ → O2 + ·OH (149)
Fe2+ + O3 → FeO2+ + O2 , k = 8.2 × 105 M−1 s−1 (150)
FeO2+ + H2 O → Fe3+ + ·OH + OH− , k = 1.3 × 10−2 M−1 s−1 (151)
Fe3+ + O3 + H2 O → FeO2+ + H+ + ·OH + O2 (152)

Thus, the mechanism of homogeneous catalytic ozonation with the usage of


transition metal ions such as Mn(II) and Fe(II) can be briefly expressed as
Eq. (153) (Wu et al., 2008). If the catalyst dosage exceeds the optimum, then
hydroxyl radicals are scavenged from reaction (154) and the degradation rate
declined (Wu et al., 2008).

M(n−1)+ + O3 + H+ → Mn+ + ·OH + O2 (153)



M (n−1)+
+ ·OH → M n+
+ OH (154)

As proposed by previous researchers (Pines and Reckhow, 2002; Rakitskaya


et al., 1999; Wu et al., 2008), the possible reaction mechanisms of O3 /Co(II)
and O3 /Ni(II) systems are shown in reactions (155–157).

Co2+ /Ni2+ + O3 + H2 O → Co(OH)2+ /Ni(OH)2+ + ·OH + O2 (155)


·OH + O3 → HO2 · + O2 , k = 5.0 × 108 M−1 s−1 (156)
HO2 · + Co(OH) /Ni(OH) 2+ 2+
→ Co /Ni
2+ 2+
+ H2 O + O2 (157)

Moreover, based on the suggestions of Pines and Reckhow (2002),


the degradation of oxalic acid with the O3 /Co(II) system at near-neutral
pH follows a two-step reaction. First, a Co(II)–oxalate complex is formed,
which is subsequently oxidized by ozone to form Co(III)–oxalate, and then
284 J. L. Wang and L. J. Xu

Co(III)–oxalate is decomposed to form an oxalate radical and Co(II). This


system shows that homogeneous catalytic ozonation does not relay solely
on the formation of hydroxyl radicals, and hydroxyl radicals are byproducts
of Co(II)-catalyzed ozonation of oxalate.
It has to be emphasized that the efficiency and the mechanism of ho-
mogeneous catalytic ozonation are influenced by several factors, such as the
pH of the solution, the concentration of ozone, and catalyst dose (Kasprzyk-
Hordern et al., 2003). Therefore, the mechanism is very complicated and
the attempt to propose catalytic ozonation pathways is difficult due to the
variety of the catalysts used and influencing factors. Generally, the homo-
geneous catalytic ozonation of organic molecules proceeds via reactions of
hydroxyl radicals, or via formation of complexes between metal ions and
organic molecules and further attack of ozone molecules on the complex
Downloaded by [Tsinghua University] at 19:33 24 September 2012

(Kasprzyk-Hordern et al., 2003; Pines and Reckhow, 2002; Rakitskaya et al.,


1999; Wu et al., 2008).
In the heterogeneous catalytic ozonation process using metal oxides
(e.g., MnO2 , TiO2 , Al2 O3 ) and supported metal oxides (e.g., Cu-Al2 O3 , Cu-
TiO2 , Ru-CeO2 , Fe2 O3 /Al2 O3 ) as catalysts, adsorption of ozone and its further
decomposition lead to surface-bound O radicals and hydroxyl radicals on the
surface of catalysts (S) (Beltrán et al., 2002; Kasprzyk-Hordern et al., 2003).
Two mechanisms can be highlighted, depending on the pH value. At acidic
pH to near-neutral pH (pH 2–6), reactions (158–160) that lead to surface-
bound O radicals rather than free radicals are proposed to account for the
lower ozone catalytic decomposition (Beltrán et al., 2002; Kasprzyk-Hordern
et al., 2003).

O3 + S ↔ O 3 − S (158)
O3 − S ↔ O − S + O 2 (159)
O3 + O − S ↔ 2O2 + S (160)

For pH > 6, reactions (161–164) are proposed and lead to the produc-
tion of ·OH and a quicker mineralization of organics (Beltrán et al., 2002;
Kasprzyk-Hordern et al., 2003).

OH− + S ↔ OH − S (161)
O3 + OH − S ↔ ·O3 − S + ·OH (162)
·O3 − S ↔ ·O − S + O2 (163)

O3 + ·O − S ↔ O2 · + S + O2 (164)

More recently, Ma’s group (Zhang et al., 2008; Zhang and Ma, 2008; Zhao
et al., 2009a; Zhao et al., 2009c; Zhao et al., 2009d) has proposed that the
initiation efficiency of ·OH is mainly influenced by the surface characteristics
Advanced Oxidation Processes for Wastewater Treatment 285

variation of the catalyst in the process of heterogeneous catalytic ozonation.


The density of the surface hydroxyl groups on the heterogeneous catalytic
surface that are believed to be crucial for the initiation of ·OH from the
decomposition of ozone is dependent to a great extent on the pH at the
point of zero charge (pHPZC ) of the catalyst (Zhang et al., 2008; Zhao et al.,
2009c; Zhao et al., 2009d). The modification of catalysts with metals results in
the conversion of the density of surface hydroxyl groups and the pHPZC . The
pH of aqueous solution is also an important factor that determines the charge
properties of surface hydroxyl groups at oxide/water interface, as inferred
from Eqs. (165) and (166) (Zhang et al., 2008; Zhao et al., 2009c; Zhao et al.,
2009d). Most of the surface hydroxyl groups are in a neutral state when
the pH of aqueous solution is near the pHPZC of the oxide, namely nearly
no surface charge exists. Otherwise, the oxide surface becomes protonated
Downloaded by [Tsinghua University] at 19:33 24 September 2012

or deprotonated when the pH of aqueous solution is below or above the


pHPZC . Molecule ozone in aqueous solution should interact with the surface
–OH2 +, generating ·OH.

SOH + H+ ↔ SOH2 + (pH < pHPZC ) (165)


− −
SOH + OH ↔ SO + H2 O (pH > pHPZC ) (166)

The formation of HO2 ·, HO3 ·, and HO4 · also can cause the initiation of ·OH
through the reactions with ozone and the intermediates derived from the
decomposition of ozone in homogeneous aqueous solution, according to
the following reactions (Ernst et al., 2004; Pera-Titus et al., 2004; Sehested
et al., 1998; Staehelin et al., 1984; Zhao et al., 2009c):

O3 + HO2 · ↔ ·OH + 2O2 , k < 104 M−1 s−1 (167)


HO2 · + HO2 − → ·OH + HO− + O2 (168)
HO3 · → ·OH + O2 , k = 1.1 × 108 M−1 s−1 (169)
−1 −1
HO2 · + HO2 · → H2 O2 + O2 , k = 8.0 × 10 M s
5
(170)
HO3 · + HO3 · → H2 O2 + 2O2 , k = 5.0 × 109 M−1 s−1 (171)
HO4 · + HO4 · → H2 O2 + 2O3 , k = 5.0 × 109 M−1 s−1 (172)
−1 −1
HO4 · + HO3 · → H2 O2 + O2 + O3 , k = 5.0 × 10 M s
9
(173)
−1 −1
H2 O2 + O3 → ·OH + HO2 · + O2 , k = 0.025 M s (174)
H2 O2 + HO2 · → ·OH + H2 O + O2 (175)
H2 O2 + O2 ·− → ·OH + HO− + O2 (176)

Consequently, the radical chain reactions are initiated on the surface of the
catalyst and in the bulk of the aqueous phase through the synergetic effect
between homogeneous and heterogeneous reaction systems.
286 J. L. Wang and L. J. Xu

The heterogeneous catalytic ozonation mechanism is still controversial


because the catalysts used are various and the pathway is so complex. The
diversity is mainly due to the differences in the catalyst performance, the
properties of target organic compounds, and the operating conditions used
by the different research groups (Beltrán et al., 2002). Thus, generalization
concerning the mechanism of the catalytic ozonation process is also diffi-
cult. Various mechanisms have been detailed by Kasprzyk-Hordern et al.
(2003), which are generally described in three ways: (a) adsorption of target
material on the catalyst surface, then further oxidation with aqueous ozone
and radicals; (b) adsorption of ozone on the catalyst surface, then reac-
tion of surface-bonded ozone molecules and radicals with aqueous target
compounds; and (c) adsorption of ozone and target materials, then further
reaction on the catalyst surface.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

2.6.3 HYBRID PROCESSES


The enhancement of ozonation by radiolysis, UV radiation, and sonolysis
has been previously described in Sections 2.1, 2.2.1, and 2.3, respectively.
In addition, the electrolysis assisted with ozonation has also been used for
the removal of refractory organic matter from wastewater (Kishimoto et al.,
2005).
Besides the indirect process of ozonation described in reactions
(141–144), several pathways can be assumed on the advanced oxidation
mechanism of electrolysis–ozonation (Kishimoto et al., 2005). The indirect
process of ozonation is promoted near the cathodes because of the high
local pH formed near the cathodes. In the other three paths that are re-
actions (177–179), active oxygen species are produced by cathode reac-
tions, initiating radical chain reactions (4) and (142–144) (Bühler et al., 1984;
Sehested et al., 1998). In addition, the cathodic reactions (180) and (181) can
be assumed in electrolysis–ozonation.

O2 + H2 O + 2e− → HO2 − (aq) + OH− (−0.06491 V vs .NHE) (177)


− −
O2 + e → O2 · (aq) (−0.331 V vs .NHE) (178)
− −
O3 + e → O3 · (aq) (179)
− − −1 −1
HO2 + O3 → O3 · + HO2 ·, k = 1.5 × 10 M s 6
(142)
− + −5 −1 −1
HO2 · ↔ O2 · + H , k = 1.59 × 10 M s (4)
− − −1 −1
O2 · + O 3 → O 3 · + O 2 , k = 1.6 × 10 M s
9
(143)
O3 ·− + H2 O → ·OH + O2 + OH− , k = 15 M−1 s−1 (144)
− −
O3 + H2 O + 2e → O2 + 2OH (1.25 V vs .NHE) (180)
− −
O2 + 2H2 O + 4e → 4OH (0.401 V vs .NHE) (181)

Kishimoto et al. (2005) proposed that reactions (177) and (178) are unlikely
to proceed in electrolysis–ozonation because a cathodic reaction with high
Advanced Oxidation Processes for Wastewater Treatment 287

standard potential generally occurs prior to reactions with lower potential.


Accordingly, the indirect process of ozonation as reactions (141–144) and
reactions (179) and (180) may possibly constitute the advanced oxidation
mechanism.

3. THE APPLICATION IN TOXIC POLLUTANTS


REMOVAL FROM WASTEWATER

Recent studies have revealed the application of advanced oxidation pro-


cesses for the treatment of wastewater that contain different kinds of toxic
compounds (Akpan and Hameed, 2009; Dutta et al., 2005; Lapertot et al.,
2006; Liu et al., 2009; Sánchez-Polo et al., 2009; Shankar et al., 2008; Song
Downloaded by [Tsinghua University] at 19:33 24 September 2012

et al., 2008a). In this section, we summarize some of the representative ap-


plications of AOPs to the wastewater treatment and highlight the parameters
assessed, removal effectiveness, and degradation mechanisms.

3.1 Removal of Aromatic Compounds From Wastewater


As seen in Table 1, the removal of armomatic compounds such as phe-
nol, chlorophenols, and nitrophenols by various AOPs has been extensively
investigated because these compounds are anthropogenic, toxic, and biore-
fractory pollutants that give foul odor to the water (Gogate, 2008; Liu et al.,
2009; Sivasankar and Moholkar, 2009b). Various operating parameters that
affect the advanced oxidation processes have been assessed in laboratory-
scale studies. For instance, Chen and Liang (2008) carried out the exper-
iments to elucidate the influence of temperature, dosage of oxygen gas,
sulfuric acid concentration, and dosage of ferrous ions on the performance
of oxidative degradation of dinitrotoluene isomers and 2,4,6-trinitrotoluene
in spent acid by electro-Fenton reagents. Under optimum operating con-
ditions, high removal efficiency can be obtained, as shown in Table 1.
It is interesting to find that molecular structures can also influence the
degradation of nitrophenols, studied by Liu et al. (2009), and the results
show that the value of pseudo-first-order rate constants of nitrophenols
lies in the order of 2,6-dinitrophenol >2,5-dinitrophenol >2,4-dinitrophenol
>2,4,6-trinitrophenol. Moreover, the degradation mechanisms have been de-
duced and demonstrated in experiments, as described comprehensively in
the preceding sections.

3.2 Removal of Dyes From Wastewater


The release of wastewaters containing textile dyes and industrial dyestuffs
in the environment is a considerable source of nonaesthetic pollution and
eutrophication (Akpan and Hameed, 2009). These dyes can present toxic
Downloaded by [Tsinghua University] at 19:33 24 September 2012

288
TABLE 1. The removal of aromatic compounds from wastewater
Results (methods with
Pollutant Applied processes Parameters assessed highest removal) Degradation mechanism Reference

Phenol US/O3 Effect of acoustic frequency, Complete and rapid The degradation was due (Lesko et al.,
comparison of various elimination of TOC to direct O3 attack and 2006)
processes, kinetics was observed using ·OH reaction
US/O3 process
Phenol Sono-Fenton (US + Effects of ultrasonic frequency, Final TOC degradation The oxidizing species (Bremner et al.,
Fe2 O3 /SBA-15 + H2 O2 concentration and of about 30% was were mainly hydroxyl 2009)
H2 O2 ) catalyst loadings, latent observed at 584 kHz and hydroperoxyl
remediation in the presence of radicals
0.6 g L−1 catalyst
p-Chlorophenol Ozonation H2 O2 detection, effects of About 95% p-chlorophenol There were direct ozone (Pi et al., 2007)
scavenger and radical probe and 26% DOC were reactions and indirect
compound, intermediates removed after 9 min ·OH reactions, and ·OH
identification reaction reactions played an
important role
Chlorophenols γ -radiation Effect of absorbed dose, Cl− When the initial The reactive species (Hu and Wang,
release, the addition of concentration was 100 mg were ·OH, ·H, and eaq − 2007)
H2 O2 , kinetics L−1 and the dosage was 6
kGy, the removal
efficiencies were 44.54%
for 2-CP, 91.46% for 3-CP,
82.72% for 4-CP and
93.25% for 2,4-DCP
4-Chloro-3-methyl Anodic oxidation Electrode characterization and Complete removal of The ·OH and active (Song et al.,
phenol (Ti/SnO2 –Sb/PbO2 stability, effects of initial 4-chloro-3-methyl phenol chlorine on the 2010)
anode) concentration and current after 90 min reaction and electrode surface had a
density, intermediates and about 49% TOC removal dominant role in the
anions detection, identification after 8-hr reaction were electrooxidation
of the reactive species achieved process
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Nitrobenzene γ -radiation Contribution of the initial Under oxidative Under the oxidative (Zhang et al.,
reactive species, kinetics, conditions, complete conditions, ·OH was the 2007)
intermediates identification mineralization was predominant species;
achieved, whereas no under the reductive
total organic carbon conditions, eaq − played a
reduction was key role
observed under
reductive conditions
Nitrobenzene Heterogeneous catalytic Comparison experiments, About 77.9% The degradation was (Zhao et al.,
ozonation production of oxidative nitrobenzene was mainly attributed to ·OH 2009d)
(Cu-cordierite + O3 ) species, initiation of ·OH converted and oxidation
from catalyst surface approximately 62.3%
of TOC was removed
after 20 min
p-Nitrophenol Anodic oxidation Comparison of degradation When the applied The mineralization capacity (Zhu et al.,
(Ti/boron-doped efficiencies using different electrical charge was of BDD anodes would 2008)
diamond (BDD), anodes, concentration 3.2 Ah L−1, the removal be mainly attributed to
Ti/SnO2 –Sb/PbO2 , evolution of ·OH, oxidation rate of TOC was 98% in the existence of free
Ti/SnO2 –Sb anodes) of p-substituted phenols, the Ti/BDD anode cell hydroxyl radicals in the
kinetics BDD anode cell
p-Nitrophenol Sono-Fenton Investigation of various Maximum extent of ·OH reaction mechanism (Pradhan and
operating conditions degradation (66.4%) Gogate, 2010)
was observed for 0.5%
p-nitrophenol
concentration (w/v)
4-Nitrophenol, Photoelectrocatalysis Characterization of electrode, A rate constant of 1.06 × Photocatalysis and (Asmussen
2-nitrophenol (TiO2 /Ti/Ta2 O5 –IrO2 photocurrent and 10−2 min−1 was created electrochemical et al., 2009)
catalyst) electrochemical current for 4-nitrophenol, and mechanisms
responses, degradation of a rate constant of
organics 1.93 × 10−2 min−1 was
created for
2-nitrophenol
Nitrophenols Anodic oxidation Characteristics of anode, The degradation These nitrophenols were (Liu et al., 2009)
(Ti/Bi–PbO2 anode) voltammetric study, efficiencies followed mainly degraded by
degradation of nitrophenols, the order means of indirect
kinetics, intermediates 2,6-dinitrophenol oxidation
analysis >2,5-dinitrophenol
>2,4-dinitrophenol
>2,4,6-trinitrophenol

289
(Continued on next page)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

290
TABLE 1. The removal of aromatic compounds from wastewater (Continued)
Results (methods with
Pollutant Applied processes Parameters assessed highest removal) Degradation mechanism Reference

2,6-dimethyl- Homogeneous Fenton Kinetics, intermediates Nearly complete ·OH reaction mechanism (Boonrattanakij
aniline process identification degradation was et al., 2009)
obtained after 10 min
2,4-Dinitrotoluene Homogeneous catalytic Effects of Mn2+, oxalic acid and 65% 2,4-dinitrotoluene was Hydroxyl radical-type (Xiao et al.,
ozonation (Mn2+ + scavenger, kinetics, degraded in 15 min by mechanism 2008)
O3 + oxalic acid) intermediates analysis homogeneous catalytic
ozonation
2,4,6- Electro-Fenton Effects of reaction temperature, Nearly complete ·OH reaction mechanism (Chen and
trinitrotoluene dosage of oxygen, sulfuric mineralization of TOC Liang, 2008)
acid concentration and could be achieved
dosage of ferrous ions,
intermediates detection
4-Chlorobenzoic Electrolysis-ozonation Combination effect of Complete degradation was Indirect process of (Kishimoto
acid ozonation and electrolysis, reached after 30 min ozonation and the et al., 2005)
mathematical model analysis, electron transfer forming
effect of electric current O3 ·− and ·OH may
constitute the oxidation
mechanism
Aniline, Photoelectro-Fenton Effects of electrolysis time, pH, Almost complete Photodecomposition of (Casado et al.,
nitrobenzene, initial concentration, and mineralization was some intermediates and 2005)
4-chlorophenol solar or UVA irradiation, reached after about indirect electrooxidation
running costs 50 min mechanism
Advanced Oxidation Processes for Wastewater Treatment 291

effects to microorganisms, aquatic life, and human beings, and reduce light
penetration in contaminated waters (Akpan and Hameed, 2009; AlHamedi
et al., 2009; Chen et al., 2008). The decolorization and degradation of dyes
by AOPs has therefore received increasing attention, as shown in Table 2.
Effects of operating parameters, decolorization, and degradation efficiencies
and degradation mechanisms have been studied (as shown in Table 2). For
example, Ma et al. (2007) investigated the decomposition and mineralization
of Congo Red by γ -rays with the changes of absorption spectra, degrada-
tion efficiency, TOC removal, and pH changes of the solutions saturated
with different gases, and proposed the degradation mechanism. It should be
noted that a faster decolorization rate has been observed, and the complete
decolorization of dyes does not mean that dyes have been oxidized com-
pletely into CO2 and H2 O because in most cases some stable intermediates
Downloaded by [Tsinghua University] at 19:33 24 September 2012

are formed in solution (El-Desoky et al., 2010; Song et al., 2007b).

3.3 Removal of Pharmaceutical Compounds From Wastewater


A large amount of pharmaceutical compounds, including antipyretics, anal-
gesics, blood lipid regulators, antibiotics, antidepressants, chemotherapy
agents, and contraceptive drugs, are consumed annually throughout the
world (Ikehata et al., 2006; Sánchez-Polo et al., 2009; Yang et al., 2009;
Yu et al., 2008). The occurrence of these pharmaceuticals in the aquatic
environment as well as in finished drinking water can cause some adverse
effects, such as aquatic toxicity, resistant development in pathogenic bacte-
ria, genotoxicity, and endocrine disruption (Ikehata et al., 2006). Recently,
concerns have been raised over the removal of pharmaceutical compounds
from wastewater by various AOPs, as depicted in Table 3. Effects of op-
erating parameters, presence of radical promoters, and scavengers, as well
as the influence of chemical composition of waters, reaction kinetics, and
biodegradability improvement have been studied, as have the degradation
mechanisms, which indicate that the ·OH is the most dominant species in
the oxidation processes.

3.4 Removal of Pesticides From Wastewater


Pesticides, divided into several types, depending on their usage, including
herbicides, insecticides, fungicides, rodenticides, nematicides, and microbio-
cides, have been extensively used in agricultural and nonagricultural sectors,
and residues are often found in surface and groundwater resources (Ikehata
and El-Din, 2005a, 2005b; Nélieu et al., 2009; Shankar et al., 2008). Due to
their persistence in aquatic environment and potential adverse health effects,
the wide range of these chemical substances can make efficient disposal
difficult (Ikehata and El-Din, 2005a, 2005b; Malpass et al., 2007). Many re-
searches to remove pesticides from wastewater using AOPs have assessed the
Downloaded by [Tsinghua University] at 19:33 24 September 2012

TABLE 2. The removal of dyes from wastewater

292
Results (methods with
Pollutant Applied processes Parameters assessed highest removal) Degradation mechanism Reference

Congo Red γ -radiation Effects of absorption In O2 saturated solutions, Reductive species such as (Ma et al., 2007)
spectra, saturated with TOC removal reached eaq − were most efficient
different gases and pH, 78% at 8.4 kGy; in N2 O to discolor the dye;
products analysis saturated solutions, TOC oxidative species such as
removal reached 86% at ·OH led high TOC
11.9 kGy; in N2 saturated removal rate of the
solutions, the highest TOC solution
removal was 20% at
6.7 kGy
Methyl orange γ -irradiation Effects of γ -ray dose rate, At a dose rate of 92 Gy Under reductive (Chen et al.,
initial dye concentration min−1, more than 90% conditions, eaq − and 2008)
and saturated with gases, dye degradation was COO·− were the main
kinetics, intermediates achieved after 120 min reactive species; under
analysis oxidative conditions,
·OH was the main
reactive species
Diacryl Red Photolysis Effects of temperature, pH, About 70% dye was Four pathways: homolysis (Zhao et al.,
X-GRL radiation flow rate of UV degraded after 300 min of excited dye to 2005b)
lamp, carrier gas flow radicals, electron transfer
rate, initial dye of excited dye to form
concentration, radical dye cation,
concentration of t-BuOH decomposition by O2 ·−,
and the concentration of and decomposition by
dissolved oxygen, kinetics singlet oxygen
CI Reactive UV/O3 Effects of solution pH, initial The TOC removal reached The UV-enhanced (Song et al.,
Yellow 145 dye concentration, ozone about 80% after 150 min ozonation was controlled 2008c)
feed and temperature, under 175 W UV by both mass transfer
intermediates detection irradiation and chemical reactions
Rhodamine B UV/H2 O2 Effects of dye concentration, 73% decoloration of the dye ·OH reaction mechanism (AlHamedi
pH, H2 O2 dose and was achieved et al., 2009)
irradiation time, kinetics,
intermediates detection,
addition of chemicals
Downloaded by [Tsinghua University] at 19:33 24 September 2012

CI Reactive Photocatalysis (UV + Effect of pH, purging gas, At pH 12.0, complete dye The decolorization (Song et al.,
Black 5 SrTiO3 /CeO2 effect of scavengers decolorization was almost primarily proceeded by 2007b)
catalyst) achieved after 120 min; photolysis and/or O2 ·−
57% of the TOC was in the bulk solution;
removed after a reaction cleavage of the
time of 300 min naphthalene and
benzene rings was
mainly attributed to the
hvb + pathway and
·OHads reactions at the
catalyst surface
Acid Orange 7 Photocatalysis Influence of scavengers, About 90% dye Dye molecules first (Chen et al.,
(UV/TiO2 ) effect of TiO2 surface concentration was adsorbed on the catalyst 2005)
characteristic reduced after 4-hr reaction surface and following
degradation was mostly
initiated by hvb +
CI Direct Red 23 US/O3 Effects of dye concentration, Approximately 98% of the ·OH reaction mechanism (Song et al.,
pH, ozone dose and dye was removed after 2007c)
ultrasonic density, 1 min treatment
kinetics, intermediates
detection
CI Reactive US/O3 Effects of initial dye About 65% TOC removal O3 , ·OH and cavitation (He et al.,
Blue 19 concentration, pH, ozone was achieved after bubbles played a major 2008a)
dose and ultrasonic 120 min role in dye degradation
energy density,
intermediates detection
Methylene blue Photoelectrocatalysis Characterization of catalyst, Irradiation of the composite The applied potential (Gao et al.,
(CNTs/TiO2 comparison experiments at a sufficiently positive could facilitate the 2008)
catalyst) potential resulted in separation of the
enhanced degradation photo-generated
efficiency electron/hole pairs at the
catalyst surface
(Continued on next page)

293
Downloaded by [Tsinghua University] at 19:33 24 September 2012

294
TABLE 2. The removal of dyes from wastewater (Continued)
Results (methods with
Pollutant Applied processes Parameters assessed highest removal) Degradation mechanism Reference

Crystal violet Homogeneous Effects of reagent Complete degradation was The species ·OH was the (Fan et al.,
Fenton process concentration and ratio observed after 10 min main oxidant 2009)
(Fe2+/Fe3+ + and pH, intermediates reaction at pH 3
H2 O2 ) identification
Orange II Heterogeneous Characterization of FeVO4 , Degradation efficiency The activation of H2 O2 by (Deng et al.,
Fenton process catalytic activity of FeVO4 , reached 94.2% after 60 both Fe(III) and V(V) in 2008)
(FeVO4 + H2 O2 ) effect of pH, stability of min FeVO4 producing ·OH
FeVO4
Reactive dyes Solar photo-Fenton Effect of Fe(II) 82% DOC removal for The ·OH as the dominant (Garcı́a-
(Procion Red concentration, dye Procion Red H-E7B 86% reactive species Montaño
H-E7B, disappearance, dissolved DOC removal for et al., 2008)
Cibacron Red organic carbon (DOC) Cibacron Red FN-R
FN-R) removal, H2 O2 and Fe
evolution, carboxylic
acids generation,
inorganic heteroatoms
release
Levafix Blue Electro-Fenton Effects of applied potential, Complete decolorization The ·OH was the mainly (El-Desoky
CA, Levafix pH, nature of supporting and approximately oxidative species et al., 2010)
Red CA electrolyte, concentration 90–95% mineralization
of catalytic ferric ions and were achieved
dye concentration,
kinetics, treatment of real
industrial wastewater
Downloaded by [Tsinghua University] at 19:33 24 September 2012

TABLE 3. The removal of pharmaceutical compounds from wastewater

Applied Results (methods with


Pollutant processes Parameters assessed highest removal) Degradation mechanism Reference

Nitroimidazoles γ -irradiation Effects of absorbed dose, Around 70% of the TOC Oxidation by ·OH (Sánchez-
(Metronidazole, contaminant concentration was radicals, and reduction Polo et al.,
Dimetridazole, concentration and pH, reduced after 700 Gy by eaq − and ·H 2009)
Tinidazole) decomposition kinetics, of treatment for
effect of radical Metronidazole
promoters and
scavengers, influence
of chemical composition
of waters
Cefaclor γ -radiation Effects of radiation dose Cefaclor, 30 mg/L, was ·OH radicals were more (Yu et al.,
and initial aqueous completely degraded closely associated with 2008)
concentration, kinetics, with 1000 Gy of radiation the radiolytic
effect of scavengers, decomposition of
purging gases cefaclor than other
radicals
Antibiotics Photo-Fenton Effects of UV irradiation, Under optimum operating ·OH reaction mechanism (Elmolla and
(amoxicillin, H2 O2 /COD molar ratio, conditions, complete Chaudhuri,
ampicillin, H2 O2 /Fe2+ molar ratio, degradation of antibiotics 2009)
cloxacillin) pH, initial antibiotics occurred in 2 min
concentration and
irradiation time,
biodegradability
improvement
Paracetamol Photocatalysis Effect of oxygen, purging More than 80% paracetamol The ·OH was the most (Yang et al.,
(UV/TiO2 ) gases and scavengers, was degraded after dominant species 2009)
intermediates detection 300 min
Triclosan Sonolysis Sonication of various Complete conversion was Degradation occurred via (Sanchez-
samples (deionised water, achieved at reasonable pyrolysis and/or Prado et al.,
seawater, water treatment times in nearly hydroxyl 2008)
containing 3.5% NaCl, all cases radical-induced
urban runoff water, reactions
centrifuged urban
runoff water, influent
wastewater), detection

295
of intermediates
(Continued on next page)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

296
TABLE 3. The removal of pharmaceutical compounds from wastewater (Continued)

Applied Results (methods with


Pollutant processes Parameters assessed highest removal) Degradation mechanism Reference

Levodopa and Sonolysis Effects of initial solute About 91% levodopa and Pollutants degradation (Isariebel
paracetamol concentration, ultrasonic 95% paracetamol were proceeded principally et al., 2009)
frequency, ultrasonic removed after 240 min through hydroxyl
power, and radical of sonolysis with actual radical-mediated
scavenger and promoter, power of 32 W for reactions
reaction kinetics 574 kHz, pollutant initial
concentration of 25 mg
L−1 and temperature
of 20◦ C
Drug ibuprofen Electron- Comparison experiment, The most potent method ·OH reaction mechanism (Skoumal
Fenton, UVA effects of Fe2+ content, was solar et al., 2009)
photoelectro- pH and current density, photoelectron-Fenton
Fenton, solar kinetics, intermediates with BDD giving 92%
photoelectro- detection mineralization
Fenton
Propranolol Ozonation The identification of major The reaction rate of Significant ·OH radical (Benner and
oxidation products, propranolol was formed during Ternes,
effects of pH and kapp, OH, propranolol was ozonation 2009)
scavenger, fragmentation 1 × 1010 M−1s−1
pathways
Antibacterial Ozonation Analysis and interpretation Nearly all of the model The oxidation resulted (Dodd et al.,
compounds of microdilution assay antibacterial compounds from reactions with 2009)
data, deactivation of were quantitatively both O3 and ·OH
model antibacterial deactivated
compounds
Advanced Oxidation Processes for Wastewater Treatment 297

effects of operating parameters to obtain the optimum operating conditions


(Table 4). The toxicity of wastewaters containing pesticides can be reduced
and the biodegradability can be enhanced after the treatment. Degradation
mechanisms are deduced by determining the influence of scavengers and
identifying intermediates.

3.5 Removal of Other Pollutants From Wastewater


The advanced oxidation processes have also been applied to deal with other
pollutants, such as carboxylic acids (Kishimoto et al., 2005; Song et al.,
2008a), heavy metals (Dutta et al., 2005; Neppolian et al., 2009), pathogens
(Li et al., 2008), and real industrial wastewaters (Chakinala et al., 2009; Kallel
Downloaded by [Tsinghua University] at 19:33 24 September 2012

et al., 2009), as seen in Table 5. The operating conditions using various AOPs
have been optimized, and reaction kinetics and degradation mechanisms
have been investigated. Although most researches have been conducted un-
der laboratory conditions, Chakinala et al. (2009) presented an advancement
over the previous work in terms of experimental data on a larger scale of
operation with an aim of enhancing the confidence among prospective users
for commercial applications.
It can be concluded that the effects of various operating parameters in
AOPs have been investigated to maximize the degradation efficiency and to
reduce the overall cost of treatment. For instance, in radiolytic degradation of
contaminant in wastewater, the operating parameters that affect the process
are irradiation dose, solution pH, and initial concentration of contaminant,
saturated with different gases and the addition of radical promoters and scav-
engers. In photocatalysis, the influence of pollutant concentration, pH, cata-
lyst loading, light intensity, dissolved oxygen concentration, and scavengers
on the degradation efficiency has been studied. Under optimum operating
conditions, a contaminant can be completely removed or converted into less
harmful compounds that are amenable to biodegradation. Furthermore, these
studies have deduced and demonstrated the degradation mechanisms in ex-
periments, which were summarized in the preceding sections. It has been
shown that hydroxyl radicals play a significant role in wastewater treatment,
which can undergo a number of different reactions with contaminant, in-
cluding addition to C C and C N double bonds, H-atom abstraction, and
electron transfer.
Generally, the combination of various advanced oxidation processes
appears to be much more beneficial for removal of pollutants as compared
with individual techniques. The extent of synergism may depend on the
enhancement in the generation of hydroxyl radicals, eventually resulting in
higher oxidation rates, or on the alteration of the reactor conditions or con-
figuration, leading to a better contact of the generated free radicals with the
pollutant molecules, and also on the better utilization of the oxidants and
Downloaded by [Tsinghua University] at 19:33 24 September 2012

298
TABLE 4. The removal of pesticides from wastewater

Applied Results (methods with


Pollutant processes Parameters assessed highest removal) Degradation mechanism Reference

Linuron Photocatalysis Comparison of various More than 70% linuron The ·OH played a (Rao and
(TiO2 /H2 O2 / processes, effect of could be decomposed predominant role in Chu, 2009)
visible light) scavengers, monitoring this process
the generation of
photocurrent,
intermediates
identification
2,4-dichloro- Sonolysis Effect of pH, purging Nearly 90% degradation ·OH radicals as the (Peller et al.,
phenoxyacetic various gases, was achieved in less primary reactive 2001)
acid intermediates generation, than 100 min species
effect of ·OH scavenger
2,4-dichloro- Heterogeneous Characterization of catalyst, About 82% of TOC was The ·OH was the main (Xing et al.,
phenoxyacetic catalytic ozonation of 2,4-D, removed at pH 3.7 with active species in 2008)
acid (2,4-D) ozonation effects of pH and MnOx /MZIW catalyst catalytic ozonation
radical scavenger dosage of 1.5 g L−1,
and gaseous ozone
concentration of
30 mg L−1
Diazinon γ -irradiation Effects of initial Complete degradation The mechanism of (Basfar et al.,
concentration of could be achieved at degradation was 2007)
contaminant and the lowest targeted probably based on
irradiation doses, concentration ·OH attack
intermediates detection
Methamidophos Anodic oxidation Effects of different electrode Complete mineralization The degradation was (Martı́nez-
(Pb/PbO2 , materials, current density was achieved using attributed to an indirect Huitle
Ti/SnO2 , and pH, products analysis Si/BDD electrode mineralization et al., 2008)
Si/BDD mechanism mainly
anodes) by ·OH
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Atrazine Photoelectro- Comparison experiments, Complete atrazine removal The degradation was (Malpass
catalysis effects of current density, was achieved at low ascribed to the direct et al., 2007)
electrolyte flow rate and current densities photolysis and
different supporting hydroxyl radicals
electrolytes
Diuron, linuron, Photolysis Effect of nitrate About 88% diuron, 91% The ·OH and nitrogen (Shankar
chlorotoluron concentration, linuron and 71% reactive species et al., 2008)
comparison of chlorotoluron were played a role in
phenylureas, by-product degraded after 16-hr the degradation
distribution, matrix effect, irradiation
degradation kinetics, tank
experiments
Pesticides (alachlor, Solar photo- Contaminant disappearance All of the pesticides tested ·OH reaction mechanism (Lapertot
atrazine, Fenton and mineralization, were mineralized, et al., 2006)
chlorfenvinphos, toxicity reduction, biodegradability was
diuron, enhancement of enhanced (70%
isoproturon) biodegradability considered
biodegradable) by the
photo-Fenton treatment
after 12–25 min

299
Downloaded by [Tsinghua University] at 19:33 24 September 2012

300
TABLE 5. The removal of other pollutants from wastewater

Applied Results (methods with


Pollutant processes Parameters assessed highest removal) Degradation mechanism Reference

Oxalic acid Photocatalysis Catalyst characterization, The concentration of oxalic Oxalic acid was first (Song et al.,
(visible light + photocatalyst activity, acid decreased was about adsorbed on the 2008a)
Ce-I-TiO2 effect of scavenger and 92%, and TOC removal surface of the catalysts,
catalyst) purging gas was about 89% and then oxidized by
·OHads
Perfluorooctane Sonolysis Sonochemical degradation About 28% PFOS and 63% Most of the PFOS and (Moriwaki
sulfonate (PFOS) of pollutants under air PFOA were removed after PFOA molecules were et al., 2005)
and perfluorooc- and argon atmosphere, 60 min irradiation under pyrolyzed at the
tanoic acid identification of products air atmosphere interfacial region
(PFOA) between the cavitation
bubbles and the bulk
solution
Polypropylene Sonolysis Effects of initial More than 60% The degradation was due (Desai et al.,
concentrations of polypropylene was to the generation of 2008)
polymer, reaction degraded using p-xylene local hot spots, highly
volumes and type of as solvent with initial reactive free radicals
solvents polymer concentration of and liquid turbulence
0.005 gm/ml and 75 ml
volume
Arsenite (As(III)) Photocatalysis Effects of As(III) The photocatalytic The ·OH was the (Dutta et al.,
(UV/TiO2 ) concentration, pH, oxidation of As(III) to dominant oxidant for 2005)
catalyst loading, light As(V) followed by As(III) oxidation
intensity, dissolved adsorption of As(V) could
oxygen concentration, completely remove
type of TiO2 surfaces and arsenic from water even
ferric ions, kinetics below the WHO drinking
water limit of 10 µg/L
(0.13 µM)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Arsenite (As(III)) Sonolysis Effects of benzoic acid, Complete oxidation of ·OH and O2 ·− radicals (Neppolian
radical scavengers, arsenite was observed were involved in the et al., 2009)
dissolved nitrogen and after 30 min sonication oxidation process
oxygen, arsenic time with initial arsenic
concentration, pH, concentration of
acoustic amplitude, 0.0013 mM, tip diameter
solution volume of the of 19 mm and power of
reactor, and humic acid 36 W at pH 7
Coliphage MS2 Photocatalysis Catalyst characterization, By combining adsorption Production of ·OH (Li et al.,
(visible light + photochemical removal and photocatalysis, final radicals was proved 2008)
TiON/PdO of coliphage MS2, ·OH virus removal rates
catalyst) formation reached 99.75–99.94%
Olive mill Heterogeneous Effects of H2 O2 dose, Fe0 Coloration of olive mill ·OH reaction mechanism (Kallel et al.,
wastewater Fenton process dose, pH and initial COD wastewater disappeared 2009)
(zero valent concentration; phenolic and phenolic compound
iron + H2 O2 ) compounds analyses decreased to 50% of
initial concentration after
3-hr reaction time
Real industrial Sono-Fenton Effects of inlet pressure, About 70% of the initial ·OH reaction mechanism (Chakinala
wastewater (US + zero temperature and the TOC was removed after et al., 2009)
valent iron + presence of copper 150 min continuous
H2 O2 ) windings, latent treatment
remediation

301
302 J. L. Wang and L. J. Xu

catalytic activity. However, the main problem of AOPs, which restricts their
industrial applications, lies in the high cost of energy sources, such as γ -ray,
ultraviolet light, or reagents such as ozone and hydrogen peroxide. For the
photocatalytic methods, many studies have focused on the development of
photocatalysts using solar radiation as an energy source that can reduce costs.
In electrochemical technologies, the energy demand could be considerably
decreased by in situ electrochemical reagent generation, and electrode ma-
terials play an important role. Solid iron-containing catalysts can react with
dissolved oxygen instead of H2 O2 in Fenton processes, which may make the
treatment far more affordable. Furthermore, many catalysts have been used
to improve the degradation efficiency of ozonation. Thus, the development
of novel catalytic materials may lead to improved and expanded application
of various types of AOPs.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

4. ISSUES ASSOCIATED WITH PRACTICAL


WASTEWATER TREATMENT

The presence of additives in solution matrix has effects on the degradation


of pollutants by AOPs, due to the fact that these additives can react with
·OH radicals (AlHamedi et al., 2009; Rauf and Ashraf, 2009; Rauf et al., 2007;
Spinks and Woods, 1990). Many additives normally present in wastewater as
ions are Fe2+, Zn2+, Ag+, Na+, CO3 2−, HCO3 −, PO4 3−, SO4 2−, Cl−, and NO2 −
(Rauf and Ashraf, 2009; Rauf et al., 2007). The addition of these ions causes
a certain decrease in the degradation of pollutants (Rauf et al., 2007). For
example, Fe2+ ions most likely undergo the chemical reaction (182) in solu-
tion with ·OH radicals, decreasing the concentration of ·OH radicals and the
degradation efficiency (Rauf et al., 2007). Likewise, the presence of CO3 2−,
HCO3 −, H2 PO4 −, and HPO4 2− ions, which are usually added to adjust the
pH of the solution, also decreases the degradation efficiency because of the
reactions with ·OH radicals, as seen in Eqs. (183–186) (Zhao et al., 2009b).
The sulfate anions in solution can react with ·OH radicals according to reac-
tion (187) (k not available), resulting in a decrease in percentage degradation
(AlHamedi et al., 2009). In the presence of chloride ions, ·OH radicals are
involved into an equilibrium reaction yielding HOCl·− that evolves into Cl·
at pH < 7, as shown in Eqs. (188) and (189) (AlHamedi et al., 2009; Mack
and Bolton, 1999; Rauf and Ashraf, 2009; Zhao et al., 2009b). Furthermore,
it has been reported that NO2 − is a very efficient ·OH scavenger, as seen in
reaction (190) (Mack and Bolton, 1999). However, it has been demonstrated
that nitrate and nitrite ions can induce the photodegradation of pollutants
under simulated sunlight because light excitation of NO3 − and NO2 − ions in
water result in the formation of ·OH, as shown in reactions (191) and (192)
Advanced Oxidation Processes for Wastewater Treatment 303

(Fischer and Warneck, 1996; Mack and Bolton, 1999; Nélieu et al., 2009;
Shankar et al., 2008; Warneck and Wurzinger, 1988).

Fe2+ + ·OH → Fe3+ + OH− , k = 3.5 × 108 M−1 s−1 (182)


− − −1 −1
CO3 2−
+ ·OH → CO3 · + OH , k = 3.9 × 10 M s
8
(183)
HCO3 − + ·OH → CO3 ·− + H2 O, k = 8.5 × 106 M−1 s−1 (184)
H2 PO4 − + ·OH → H2 PO4 · + OH− , k < 105 M−1 s−1 (185)
HPO4 2− + ·OH → HPO4 ·− + OH− , k < 107 M−1 s−1 (186)
− −
SO4 2−
+ ·OH → SO4 · + OH (187)
Cl− + ·OH ↔ HOCl·− , k = 4.3 × 109 M−1 s−1 (188)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

HOCl·− + H+ → Cl· + H2 O, k = 2.1 × 1010 M−1 s−1 (189)


− − −1 −1
NO2 + ·OH → NO2 · + OH , k = 1.0 × 10 M s
10
(190)
− −
NO3 + H2 O + hv → ·NO2 + ·OH + OH , OH = 0.009–0.017 (191)
NO2 − + H2 O + hv → ·NO + ·OH + OH− , OH = 0.015–0.076 (192)

The influence of various scavengers on the degradation efficiency has


been determined usually to investigate the possible degradation mechanism
and pathway of pollutants (San et al., 2001; Song et al., 2007b). Alcohols,
such as methanol, 2-propanol, n-butanol, 2-butanol, and t-butanol, have
been used to scavenger hydroxyl radicals produced in solution (Buxton
et al., 1988; Jonsson et al., 2008; Song et al., 2007b; Xue et al., 2009; Yu
et al., 2008). For instance, methanol and t-butanol react more rapidly with
·OH than with ·H and e−, as shown in Eqs. (193–198) (Buxton et al., 1988;
Sánchez-Polo et al., 2009; Yu et al., 2008). Excess 2-propanol in solution
as an ·OH scavenger is rapidly oxidized by hydroxyl radicals (k = 3.0 ×
109 M−1s−1) and is less reactive with superoxide O2 ·− (k = 1.0 × 106 M−1s−1)
(Xue et al., 2009).

CH3 OH + ·OH → ·CH2 OH + H2 O,


k = (0.78–1.0) × 109 M−1 s−1 (193)
CH3 OH + ·H → ·CH2 OH + H2 , k = 2.6 × 106 M−1 s−1 (194)
CH3 OH + eaq − → CH3 O− + ·H, k < 1.0 × 104 M−1 s−1 (195)
CH3 C(CH3 )2 OH + ·OH → CH2 C(CH3 )2 OH + H2 O,
k = 6 × 108 M−1 s−1 (196)
304 J. L. Wang and L. J. Xu

CH3 C(CH3 )2 OH + ·H → CH2 C(CH3 )2 OH + H2 ,


k = 1.7 × 105 M−1 s−1 (197)
− − +
CH3 C(CH3 )2 OH + eaq → CH3 C(CH3 )2 O + H ,
k = 4.0 × 105 M−1 s−1 (198)

Persulfate ions (S2 O8 2−) can trap electrons in solution, producing sulfate
radical anions (SO4 −·) that can further react with water molecules to form
·OH radicals, as shown in Eqs. (199) and (200) (Poulios and Tsachpinis,
1999; Rauf and Ashraf, 2009; San et al., 2001; Wang and Hong, 1999).

S2 O8 2− + e− → SO4 −· +SO4 2− , E 0 = 2.6 V (199)


−· +
+ H2 O → ·OH + SO4 2−
+ H , (k not available)
Downloaded by [Tsinghua University] at 19:33 24 September 2012

SO4 (200)

Bromate ion (BrO3 −) is also an efficient electron scavenger, as seen in reac-


tion (201) (Martin et al., 1995; Rauf et al., 2007; San et al., 2001). The bromide
ions (Br−) produced in solution can then react with ·OH radicals (Eq. [202])
(Rauf et al., 2007; San et al., 2001).

BrO3 − + 6e− + 6H+ → Br− + 3H2 O, E 0 = 1.44 V (201)


− − −1 −1
Br + ·OH → ·Br + OH , k = 1.1 × 10 M s 9
(202)

In photocatalysis, the fluoride ion (F−) shows strong adsorption on the cata-
lyst surface, and it is very stable against oxidation even by the valence band
holes (hvb +), as the redox potential of the couple F·/F− is about 3.6 V (Chen
et al., 2005; Yang et al., 2006). For example, in the presence of fluoride
ions, the concentration of surface hydroxyl group on the TiO2 surface can
be controlled by adopting fluoride exchange, as shown in Eq. (203), and
then free ·OH can be generated as Eq. (204) (Chen et al., 2005).

≡ Ti–OH + F− ↔ ≡ Ti–F + OH− , pKF = 6.2 (203)


≡ Ti–F + H2 O/OH− + hvb + → ≡ Ti–F + ·OHfree
+ H+ , (k not available) (204)

The iodide ion (I−) is a scavenger that reacts with hvb + and surface-bounded
·OHads , reducing the number of oxidizing species available on the cata-
lyst surface (He et al., 2008b; Martin et al., 1995). A likely reaction path-
way initiated by valence-band hole oxidation is shown in Eqs. (205–207),
and the oxidation by a surficial hydroxyl radical is also possible, as seen
Eqs. (208–210) (Martin et al., 1995).

I− (+hvb + ) ↔ I· + e− , E 0 = −1.3 V (205)


− − −1 −1
I· + I → I2 · , k = 1.0 × 10 M s 10
(206)
Advanced Oxidation Processes for Wastewater Treatment 305

I2 ·− ↔ I2 + e− , E 0 = − 0.2 V (207)
I− + ·OHads → I · +OH− , k = 1.2 × 1010 M−1 s−1 (208)
·OHads + I2 → HOI + I·, k = 1.1 × 1010 M−1 s−1 (209)
− − −1 −1
I · + I → I2 · , k = 1.0 × 10 M s 10
(210)

Tetranitromethane (TNM) is an effective radical scavenger that can react with


·H, HO2 ·, and O2 ·− as described in reactions (211–213) (Al-Sheikhly et al.,
2006; Henglein et al., 1959; Yoon and Lee, 2005). In photocatalysis, TNM
eliminates the species responsible for the formation of free ·OH radicals in
solution, yet it does not remove adsorbed ·OHads radicals because TNM is
not strongly adsorbed to the surface of catalyst (El-Morsi et al., 2000).
Downloaded by [Tsinghua University] at 19:33 24 September 2012

·H + C(NO2 )4 → H+ + C(NO2 )3 − + NO2 (211)


+ −
HO2 · + C(NO2 )4 → H + C(NO2 )3 + NO2 + O2 (212)
O2 ·− + C(NO2 )4 → C(NO2 )3 − + NO2 + O2 , k = 1.9 × 109 M−1 s−1 (213)

The application of these AOPs to the real wastewater treatment has


attracted many researchers, which has been detailed in previous reviews
(Bautista et al., 2008; Belgiorno et al., 2007; Martı́nez-Huitle and Brillas,
2009; Satyawali and Balakrishnan, 2008). Various AOPs have been demon-
strated in full-scale applications as feasible technologies for the treatment of
a wide diversity of industrial wastewaters. Bandala et al. (2008) investigated
the application of the photo-Fenton process to colored real wastewater, and
the result showed that a decrease of 62.6% in the concentration of COD
could be observed. An industrial pharmaceutical wastewater was treated by
a Fe2 O3 /SBA-15 nanocomposite catalyst and hydrogen peroxide, and the re-
sults showed that a high oxidation degree of organic compounds in the outlet
effluent was obtained, enhancing its biodegradability (Melero et al., 2009).
Moreover, Sirtori et al. (2009) used the solar photo-Fenton as finishing step
for biological treatment of a pharmaceutical wastewater, and total degrada-
tion of nalidixic acid and reduction in toxicity were observed, demonstrating
that the photo-Fenton process applied to finish off biological treatment is
a useful approach. Although the efficient treatment of wastewater is exten-
sively observed, majority of the work is still on a laboratory scale, which can
be seen from Tables 1–5. Thus, more efforts are still required in the studies
with real effluents, and a large quantum of work needs to be done in terms
of the design strategies for the scale-up.

5. CONCLUDING REMARKS

The ideas presented in this article provide a holistic view on the progress
of advanced oxidation processes for wastewater treatment based on the
306 J. L. Wang and L. J. Xu

formation mechanisms of the hydroxyl radical. Generally, free radicals are


generated via the decomposition of water molecules by radiation, photol-
ysis, and sonolysis, which subsequently attack the contaminant molecules.
Ozonation and use of hydrogen peroxide work on the direct attack of the ox-
idants or via formation of free radicals combined with radiation, photolysis,
or sonolysis. Catalytic technologies such as photocatalysis, anodic oxidation,
photoelectrocatalysis, Fenton processes, and catalytic ozonation use a com-
bination of irradiation, oxidation (O3 , H2 O2 ), electrons, and catalysts as a
means of generating free radicals. However, there are still many controver-
sial issues in degradation mechanism of contaminant, such as the role of
various radicals in photocatalysis and the nature of the reactive oxidant, ·OH
vs Fe(IV), produced by the Fenton reaction. Moreover, the mechanisms of
catalytic ozonation are rather complicated and not fully understood. There-
Downloaded by [Tsinghua University] at 19:33 24 September 2012

fore, further research is needed to explore the degradation mechanism of


specific process.
These processes have been applied to remove contaminants from
wastewater or convert pollutants into more biodegradable compounds as
a pretreatment stage. AOPs may become applicable on an industrial scale
when every effectiveness factor is optimized and different processes are com-
bined to eliminate some of the drawbacks of the individual techniques for
the final goal to achieve maximum efficiency with minimum energy input.
In particular, more efforts are required to handle some key problems, such
as the development of novel catalytic materials. Furthermore, the amount
of ·OH produced is an important parameter for efficiency assessment and
proper comparison of different advanced oxidation technologies. The qual-
itative and quantitative analysis of ·OH is also another attractive option to
explore the detailed degradation mechanism, but more work is required in
this direction.

ACKNOWLEDGMENTS

The authors are grateful for the financial support provided the National
Natural Science Foundation of China (Grant Nos. 10876016; 50978145).

REFERENCES

Abd El-Raady, A., Nakajima, T., and Kimchhayarasy, P. (2005). Catalytic ozonation
of citric acid by metallic ions in aqueous solution. Ozone Sci. Eng., 27, 495–498.
Adams, C.D., Scanlan, P.A., and Secrist, N.D. (1994). Oxidation and biodegradability
enhancement of 1,4-Dioxane using hydrogen peroxide and ozone. Environ. Sci.
Technol., 28, 1812–1818.
Adewuyi, Y.G. (2001). Sonochemistry: Environmental science and engineering ap-
plications. Ind. Eng. Chem. Res., 40, 4681–4715.
Advanced Oxidation Processes for Wastewater Treatment 307

Adewuyi, Y.G. (2005a). Sonochemistry in environmental remediation. 1. Combina-


tive and hybrid sonophotochemical oxidation processes for the treatment of
pollutants in water. Environ. Sci. Technol., 39, 3409–3420.
Adewuyi, Y.G. (2005b). Sonochemistry in environmental remediation. 2. Heteroge-
neous sonophotocatalytic oxidation processes for the treatment of pollutants in
water. Environ. Sci. Technol., 39, 8557–8570.
Agustina, T.E., Ang, H.M., and Vareek, V.K. (2005). A review of synergistic effect of
photocatalysis and ozonation on wastewater treatment. J. Photochem. Photobiol.
C, 6, 264–273.
Akpan, U.G., and Hameed, B.H. (2009). Parameters affecting the photocatalytic
degradation of dyes using TiO2 -based photocatalysts: A review. J. Hazard.
Mater., 170, 520–529.
Al-Ekabi, H., and Serpone, N. (1988). Kinetic studies in heterogeneous photocatal-
ysis. 1. Photocatalytic degradation of chlorinated phenols in aerated aqueous
Downloaded by [Tsinghua University] at 19:33 24 September 2012

solutions over TiO2 supported on a glass matrix. J. Phys. Chem., 92, 5726–
5731.
Al-Sheikhly, M., Poster, D.L., An, J.C., Neta, P., Silverman, J., and Huie, R.E. (2006).
Ionizing radiation-induced destruction of benzene and dienes in aqueous media.
Environ. Sci. Technol., 40, 3082–3088.
AlHamedi, F.H., Rauf, M.A., and Ashraf, S.S. (2009). Degradation studies of Rho-
damine B in the presence of UV/H2 O2 . Desalination, 239, 159–166.
Anjaneyulu, Y., Chary, N.S., and Raj, D.S.S. (2005). Decolourization of industrial
effluents—available methods and emerging technologies—a review. Rev. Envi-
ron. Sci. Biotechnol., 4, 245–273.
Anpo, M., Shima, T., and Kubokawa, Y. (1985). Electron-spin-resonance and photo-
luminescence evidence for the photocatalytic formation of hydroxyl radiclas on
small TiO2 particles. Chem. Lett., 1799–1802.
Asmussen, R.M., Tian, M., and Chen, A.C. (2009). A new approach to wastewa-
ter remediation based on bifunctional electrodes. Environ. Sci. Technol., 43,
5100–5105.
Auguliaro, V., Davı̀, E., Palmisano, L., Schiavello, M., and Sclafani, A. (1990). In-
fluence of hydrogen peroxide on the kinetics of phenol photodegradation in
aqueous titanium dioxide dispersion. Appl. Catal., 65, 101–116.
Azbar, N., Yonar, T., and Kestioglu, K. (2004). Comparison of various advanced ox-
idation processes and chemical treatment methods for COD and color removal
from a polyester and acetate fiber dyeing effluent. Chemosphere, 55, 35–43.
Azrague, K., Bonnefille, E., Pradines, V., Pimienta, V., Oliveros, E., Maurette, M.T.,
and Benoit-Marquié, F. (2005). Hydrogen peroxide evolution during V-UV pho-
tolysis of water. Photochem. Photobiol. Sci., 4, 406–408.
Bandala, E.R., Peláez, M.A., Garcı́a-López, A.J., Salgado, M.D., and Moeller, G. (2008).
Photocatalytic decolourisation of synthetic and real textile wastewater contain-
ing benzidine-based azo dyes. Chem. Eng. Process., 47, 169–176.
Barb, W.G., Baxendale, J.H., George, P., and Hargrave, K.R. (1949). Reactions of
ferrous and ferric ions with hydrogen peroxide. Nature, 163, 692–694.
Barb, W.G., Baxendale, J.H., George, P., and Hargrave, K.R. (1951a). Reactions of
ferrous and ferric ions with hydrogen peroxide. Part I. The ferrous ion reaction.
Trans. Faraday Soc., 47, 462–500.
308 J. L. Wang and L. J. Xu

Barb, W.G., Baxendale, J.H., George, P., and Hargrave, K.R. (1951b). Reactions of
ferrous and ferric ions with hydrogen peroxide. Part II. The ferric ion reaction.
Trans. Faraday Soc., 47, 591–616.
Basfar, A.A., Khan, H.M., Al-Shahrani, A.A., and Cooper, W.J. (2005). Radiation
induced decomposition of methyl tert-butyl ether in water in presence of chlo-
roform: Kinetic modelling. Water Res., 39, 2085–2095.
Basfar, A.A., Mohamed, K.A., Al-Abduly, A.J., Al-Kuraiji, T.S., and Al-Shahrani, A.A.
(2007). Degradation of diazinon contaminated waters by ionizing radiation.
Radiat. Phys. Chem., 76, 1474–1479.
Bautista, P., Mohedano, A.F., Casas, J.A., Zazo, J.A., and Rodriguez, J.J. (2008).
An overview of the application of Fenton oxidation to industrial wastewaters
treatment. J. Chem. Technol. Biotechnol., 83, 1323–1338.
Baxendale, J.H., and Wilson, J.A. (1957). The photolysis of hydrogen peroxide at
high light intensities. Trans. Faraday Soc., 53, 344–356.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Belgiorno, V., Rizzo, L., Fatta, D., Della Rocca, C., Lofrano, G., Nikolaou, A., Naddeo,
V., and Meric, S. (2007). Review on endocrine disrupting-emerging compounds
in urban wastewater: Occurrence and removal by photocatalysis and ultrasonic
irradiation for wastewater reuse. Desalination, 215, 166–176.
Beltrán, F.J., Rivas, J., Álvarez, P., and Montero-de-Espinosa, R. (2002). Kinetics of
heterogeneous catalytic ozone decomposition in water on an activated carbon.
Ozone Sci. Eng., 24, 227–237.
Beltran-Heredia, J., Torregrosa, J., Dominguez, J.R., and Peres, J.A. (2001). Compari-
son of the degradation of p-hydroxybenzoic acid in aqueous solution by several
oxidation processes. Chemosphere, 42, 351–359.
Benner, J., and Ternes, T.A. (2009). Ozonation of propranolol: Formation of oxida-
tion products. Environ. Sci. Technol., 43, 5086–5093.
Bhatkhande, D.S., Pangarkar, V.G., and Beenackers, A.A.C.M. (2002). Photocat-
alytic degradation for environmental applications—a review. J. Chem. Technol.
Biotechnol., 77, 102–116.
Bielski, B.H.J., Cabelli, D.E., Arudi, R.L., and Ross, A.B. (1985). Reactivity of HO2 /O2
radicals in aqueous solution. J. Phys. Chem. Ref. Data, 14, 1041–1100.
Blanco-Galvez, J., Fernández-Ibáñez, P., and Malato-Rodrı́guez, S. (2007). Solar pho-
tocatalytic detoxification and disinfection of water: Recent overview. J. Sol.
Energ.-T. Asme, 129, 4–15.
Boonrattanakij, N., Lu, M.C., and Anotai, J. (2009). Kinetics and mechanism of
2,6-dimethyl-aniline degradation by hydroxyl radicals. J. Hazard. Mater., 172,
952–957.
Bossmann, S.H., Oliveros, E., Göb, S., Siegwart, S., Dahlen, E.P., Payawan, L., Straub,
M., Wörner, M., and Braun, A.M. (1998). New evidence against hydroxyl radicals
as reactive intermediates in the thermal and photochemically enhanced Fenton
reactions. J. Phys. Chem. A, 102, 5542–5550.
Bremner, D.H., Molina, R., Martı́nez, F., Melero, J.A., and Segura, Y. (2009). Degra-
dation of phenolic aqueous solutions by high frequency sono-Fenton systems
(US-Fe2 O3 /SBA-15-H2 O2 ). Appl. Catal. B: Environ., 90, 380–388.
Brillas, E., Calpe, J.C., and Casado, J. (2000). Mineralization of 2,4-D by advanced
electrochemical oxidation processes. Water Res., 34, 2253–2262.
Advanced Oxidation Processes for Wastewater Treatment 309

Brillas, E., Sirés, I., and Oturan, M.A. (2009). Electro-Fenton process and related
electrochemical technologies based on Fenton’s reaction chemistry. Chem. Rev.,
109, 6570–6631.
Bühler, R.E., Staehelin, J., and Hoigné, J. (1984). Ozone decomposition in water
studied by pulse radiolysis. 1. HO2 /O2 − and HO3 /O3 − as intermediates. J. Phys.
Chem., 88, 2560–2564.
Buxton, G.V., Greenstock, C.L., Helman, W.P., and Ross, A.B. (1988). Critical review
of rate constants for reactions of hydrated electrons, hydrogen atoms and hy-
droxyl radicals (·OH/·O−) in aqueous solution. J. Phys. Chem. Ref. Data, 17,
518–886.
Buxton, G.V., and Subhani, M.S. (1972). Radiation chemistry and photochemistry of
oxychlorine ions. 2. Photodecomposition of aqueous solutions of hypochlorite
ions. J. Chem. Soc., Faraday Trans. I, 68, 958.
Carvalho, C., Fernandes, A., Lopes, A., Pinheiro, H., and Gonçalves, I. (2007). Elec-
Downloaded by [Tsinghua University] at 19:33 24 September 2012

trochemical degradation applied to the metabolites of Acid Orange 7 anaerobic


biotreatment. Chemosphere, 67, 1316–1324.
Casado, J., Fornaguera, J., and Galán, M.I. (2005). Mineralization of aromatics in
water by sunlight-assisted electro-Fenton technology in a pilot reactor. Environ.
Sci. Technol., 39, 1843–1847.
Catalkaya, E.C., and Kargi, F. (2007). Color, TOC and AOX removals from pulp
mill effluent by advanced oxidation processes: A comparative study. J. Hazard.
Mater., 139, 244–253.
Chakinala, A.G., Gogate, P.R., Burgess, A.E., and Bremner, D.H. (2009). Industrial
wastewater treatment using hydrodynamic cavitation and heterogeneous ad-
vanced Fenton processing. Chem. Eng. J., 498–502.
Chakrabarti, S., Chaudhuri, B., Bhattacharjee, S., Das, P., and Dutta, B.K. (2008).
Degradation mechanism and kinetic model for photocatalytic oxidation of PVC-
ZnO composite film in presence of a sensitizing dye and UV radiation. J. Hazard.
Mater., 154, 230–236.
Chelkowska, K., Grasso, D., Fábián, I., and Gordon, G. (1992). Numerical simulations
of aqueous ozone decomposition. Ozone Sci. Eng., 14, 33–49.
Chen, G.H. (2004). Electrochemical technologies in wastewater treatment. Sep. Purif.
Technol., 38, 11–41.
Chen, W.S., and Liang, J.S. (2008). Decomposition of nitrotoluenes from trinitro-
toluene manufacturing process by electro-Fenton oxidation. Chemosphere, 72,
601–607.
Chen, Y.P., Liu, S.Y., Yu, H.Q., Yin, H., and Li, Q.R. (2008). Radiation-induced
degradation of methyl orange in aqueous solutions. Chemosphere, 72, 532–536.
Chen, Y.X., Yang, S.Y., Wang, K., and Lou, L.P. (2005). Role of primary active
species and TiO2 surface characteristic in UV-illuminated photodegradation of
Acid Orange 7. J. Photochem. Photobiol. A: Chem., 172, 47–54.
Chevaldonnet, C., Cardy, H., and Dargelos, A. (1986). Ab initio CI calculations on
the PE and VUV spectra of hydrogen peroxide. Chem. Phys., 102, 55–61.
Chou, S.S., Huang, Y.H., Lee, S.N., Huang, G.H., and Huang, C.P. (1999). Treatment
of high strength hexamine-containing wastewater by electro-Fenton method.
Water Res., 33, 751–759.
310 J. L. Wang and L. J. Xu

Chowdhury, P., and Viraraghavan, T. (2009). Sonochemical degradation of chlori-


nated organic compounds, phenolic compounds and organic dyes: A review.
Sci. Total Environ., 407, 2474–2492.
Christensen, H., Sehested, K., and Corfitzen, H. (1982). Reactions of hydroxyl radicals
with hydrogen-peroxide at ambient and elevated-temperatures. J. Phys. Chem.,
86, 1588–1590.
Comninellis, C. (1994). Electrocatalysis in the electrochemical conversion/
combustion of organic pollutants for waste water treatment. Electrochim. Acta,
39, 1857–1862.
Cooper, W.J., Cramer, C.J., Martin, N.H., Mezyk, S.P., O’Shea, K.E., and von Son-
ntag, C. (2009). Free radical mechanisms for the treatment of methyl tert-butyl
ether (MTBE) via advanced oxidation/reductive processes in aqueous solutions.
Chem. Rev., 109, 1302–1345.
De Laat, J., and Gallard, H. (1999). Catalytic decomposition of hydrogen peroxide
Downloaded by [Tsinghua University] at 19:33 24 September 2012

by Fe(III) in homogeneous aqueous solution: Mechanism and kinetic modeling.


Environ. Sci. Technol., 33, 2726–2732.
De Laat, J., and Le, T.G. (2006). Effects of chloride ions on the iron(III)-catalyzed
decomposition of hydrogen peroxide and on the efficiency of the Fenton-like
oxidation process. Appl. Catal. B: Environ., 66, 137–146.
Del Rı́o, A.I., Molina, J., Bonastre, J., and Cases, F. (2009). Influence of electrochem-
ical reduction and oxidation processes on the decolourisation and degradation
of C.I. Reactive Orange 4 solutions. Chemosphere, 75, 1329–1337.
Deng, J.H., Jiang, J.Y., Zhang, Y.Y., Lin, X.P., Du, C.M., and Xiong, Y. (2008). FeVO4
as a highly active heterogeneous Fenton-like catalyst towards the degradation
of Orange II. Appl. Catal. B: Environ., 84, 468–473.
Desai, V., Shenoy, M.A., and Gogate, P.R. (2008). Degradation of polypropylene
using ultrasound-induced acoustic cavitation. Chem. Eng. J., 140, 483–487.
Dodd, M.C., Kohler, H.P.E., and Von Gunten, U. (2009). Oxidation of antibacterial
compounds by ozone and hydroxyl radical: Elimination of biological activity
during aqueous ozonation processes. Environ. Sci. Technol., 43, 2498–2504.
Du, W.P., Xu, Y.M., and Wang, Y.S. (2008). Photoinduced degradation of Orange
II on different iron (hydr)oxides in aqueous suspension: Rate enhancement on
addition of hydrogen peroxide, silver nitrate, and sodium fluoride. Langmuir,
24, 175–181.
Du, Y.X., Zhou, M.H., and Lei, L.C. (2007a). Kinetic model of 4-CP degradation by
Fenton/O2 system. Water Res., 41, 1121–1133.
Du, Y.X., Zhou, M.H., and Lei, L.C. (2007b). The role of oxygen in the degradation
of p-chlorophenol by Fenton system. J. Hazard. Mater., 139, 108–115.
Dutta, P.K., Pehkonen, S.O., Sharma, V.K., and Ray, A.K. (2005). Photocatalytic
oxidation of arsenic(III): Evidence of hydroxyl radicals. Environ. Sci. Technol.,
39, 1827–1834.
El-Desoky, H.S., Ghoneim, M.M., El-Sheikh, R., and Zidan, N.M. (2010). Oxidation
of Levafix CA reactive azo-dyes in industrial wastewater of textile dyeing by
electro-generated Fenton’s reagent. J. Hazard. Mater., 175, 858–865.
El-Morsi, T.M., Budakowski, W.R., Abd-El-Aziz, A.S., and Friesen, K.J. (2000). Photo-
catalytic degradation of 1,10-dichlorodecane in aqueous suspensions of TiO2 : A
Advanced Oxidation Processes for Wastewater Treatment 311

reaction of adsorbed chlorinated alkane with surface hydroxyl radicals. Environ.


Sci. Technol., 34, 1018–1022.
Elmolla, E.S., and Chaudhuri, M. (2009). Degradation of the antibiotics amoxicillin,
ampicillin and cloxacillin in aqueous solution by the photo-Fenton process.
J. Hazard. Mater., 172, 1476–1481.
Ensminger, D. (1973). Ultrasonic: The low and high-intensity-applications, Marcel
Dekker: New York.
Entezari, M.H., and Kruus, P. (1996). Effect of frequency on sonochemical reactions.
2. Temperature and intensity effects. Ultrason. Sonochem., 3, 19–24.
Entezari, M.H., Kruus, P., and Otson, R. (1997). The effect of frequency on sono-
chemical reactions. 3. Dissociation of carbon disulfide. Ultrason. Sonochem., 4,
49–54.
Ernst, M., Lurot, F., and Schrotter, J.C. (2004). Catalytic ozonation of refractory or-
ganic model compounds in aqueous solution by aluminum oxide. Appl. Catal.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

B: Environ., 47, 15–25.


Esplugas, S., Bila, D.M., Krause, L.G.T., and Dezotti, M. (2007). Ozonation and
advanced oxidation technologies to remove endocrine disrupting chemicals
(EDCs) and pharmaceuticals and personal care products (PPCPs) in water ef-
fluents. J. Hazard. Mater., 149, 631–642.
Esquivel, K., Arriaga, L.G., Rodrı́guez, F.J., Martı́nez, L., and Godı́nez, L.A. (2009).
Development of a TiO2 modified optical fiber electrode and its incorporation
into a photoelectrochemical reactor for wastewater treatment. Water Res., 43,
3593–3603.
Fagnoni, M., Dondi, D., Ravelli, D., and Albini, A. (2007). Photocatalysis for the
formation of the C–C bond. Chem. Rev., 107, 2725–2756.
Fan, H.J., Huang, S.T., Chung, W.H., Jan, J.L., Lin, W.Y., and Chen, C.C. (2009).
Degradation pathways of crystal violet by Fenton and Fenton-like systems: Con-
dition optimization and intermediate separation and identification. J. Hazard.
Mater., 171, 1032–1044.
Faust, B.C., and Hoigné, J. (1990). Photolysis of Fe (III)-hydroxy complexes as
sources of OH radicals in clouds, fog and rain. Atmos. Environ. A, 24, 79–89.
Fischer, C.H., Hart, E.J., and Henglein, A. (1986). Ultrasonic irradiation of water in
the presence of oxygen 18,18O2 : Isotope exchange and isotopic distribution of
H2 O2 . J. Phys. Chem., 90, 1954–1956.
Fischer, M., and Warneck, P. (1996). Photodecomposition of nitrite and undissociated
nitrous acid in aqueous solution. J. Phys. Chem., 100, 18749–18756.
Fogler, H.S., and Timmerhaus, K.D. (1966). Effect of ultrasonic waves on mass
transfer rates of selected fluids. AICHE J., 12, 90.
Fox, M.A., and Dulay, M.T. (1993). Heterogeneous photocatalysis. Chem. Rev., 93,
341–357.
Gao, B., Peng, C.A., Chen, G.Z., and Puma, G.L. (2008). Photo-electro-catalysis en-
hancement on carbon nanotubes/titanium dioxide (CNTs/TiO2 ) composite pre-
pared by a novel surfactant wrapping sol-gel method. Appl. Catal. B: Environ.,
85, 17–23.
Garcı́a-Montaño, J., Pérez-Estrada, L., Oller, I., Maldonado, M.I., Torrades, F., and
Peral, J. (2008). Pilot plant scale reactive dyes degradation by solar photo-Fenton
and biological processes. J. Photochem. Photobiol. A: Chem., 195, 205–214.
312 J. L. Wang and L. J. Xu

Garoma, T., and Gurol, M.D. (2004). Degradation of tert-butyl alcohol in dilute
aqueous solution by an O3 /UV process. Environ. Sci. Technol., 38, 5246–5252.
Gerischer, H., and Heller, A. (1991). The role of oxygen in photooxidation of organic
molecules on semiconductor particles. J. Phys. Chem., 95, 5261–5267.
Getoff, N. (1996). Radiation-induced degradation of water pollutants: State of the art.
Radiat. Phys. Chem., 47, 581–593.
Getoff, N. (2002). Factors influencing the efficiency of radiation-induced degradation
of water pollutants. Radiat. Phys. Chem., 65, 437–446.
Ghodbane, H., and Hamdaoui, O. (2009). Intensification of sonochemical decoloriza-
tion of anthraquinonic dye Acid Blue 25 using carbon tetrachloride. Ultrason.
Sonochem., 16, 455–461.
Gogate, P.R. (2008). Treatment of wastewater streams containing phenolic com-
pounds using hybrid techniques based on cavitation: A review of the current
status and the way forward. Ultrason. Sonochem., 15, 1–15.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Gogate, P.R., and Pandit, A.B. (2004a). A review of imperative technologies for
wastewater treatment I: Oxidation technologies at ambient conditions. Adv.
Environ. Res., 8, 501–551.
Gogate, P.R., and Pandit, A.B. (2004b). A review of imperative technologies for
wastewater treatment II: Hybrid methods. Adv. Environ. Res., 8, 553–597.
Gomes, A., Fernandes, E., and Lima, J. (2005). Fluorescence probes used for detec-
tion of reactive oxygen species. J. Biochem. Biophys. Methods, 65, 45–80.
Goulet, T., and Jay-Gerin, J.P. (1992). On the reactions of hydrated electrons with
OH· and H3 O+. Analysis of photoionization experiments. J. Chem. Phys., 96,
5076–5087.
Haber, F., and Weiss, J. (1934). The catalytic decomposition of hydrogen peroxide
by iron salts. Proc. Roy. Soc. A., 134, 332–351.
Halliwell, B. (1991). Drug antioxidant effects: A basis for drug selection. Drugs, 42,
569–605.
Hatchard, C.G., and Parker, C.A. (1956). A new sensitive chemical actinometer. II.
Potassium ferrioxalate as a standard chemical actinometer. Proc. R. Soc. London,
Ser. A, 235, 518–536.
He, Z.Q., Lin, L.L., Song, S., Xia, M., Xu, L.J., Ying, H.P., and Chen, J.M. (2008a).
Mineralization of C.I. Reactive Blue 19 by ozonation combined with sonolysis:
Performance optimization and degradation mechanism. Sep. Purif. Technol., 62,
376–381.
He, Z.Q., Song, S., Ying, H.P., Xu, L.J., and Chen, J.M. (2007). p-aminophenol degra-
dation by ozonation combined with sonolysis: Operating conditions influence
and mechanism. Ultrason. Sonochem., 14, 568–574.
He, Z.Q., Xu, X., Song, S., Xie, L., Tu, J.J., Chen, J.M., and Yan, B. (2008b). A
visible light-driven titanium dioxide photocatalyst codoped with lanthanum and
iodine: An application in the degradation of oxalic acid. J. Phys. Chem. C, 112,
16431–16437.
Heit, G., and Braun, A.M. (1997). VUV-photolysis of aqueous systems: Spatial dif-
ferentiation between volumes of primary and secondary reactions. Water Sci.
Technol., 35, 25–30.
Henglein, A., Langhoff, J., and Schmidt, G. (1959). Tetranitromethane as a radical
scavenger in radiation chemical studies. J. Phys. Chem., 63, 980.
Advanced Oxidation Processes for Wastewater Treatment 313

Hoffmann, M.R., Martin, S.T., Choi, W.Y., and Bahnemann, D.W. (1995). Environ-
mental applications of semiconductor photocatalysis. Chem. Rev., 95, 69–96.
Hoigné, J. (1998). Chemistry of aqueous ozone and transformation of pollutants by
ozonation and advanced oxidation processes, The Handbook of Environmental
Chemistry, Vol. 5. Springer, Berlin.
Hoigné, J., and Bader, H. (1975). Ozonation of water: Role of hydroxyl radicals as
oxidizing intermediates. Science, 190, 782–784.
Hoigné, J., and Bader, H. (1979). Ozonation of water: Selectivity and rate of oxidation
of solutes. Ozone Sci. Eng., 1, 73–85.
Hoigné, J., and Bader, H. (1983a). Rate constants of reactions of ozone with organic
and inorganic compounds in water. I: Non-dissociating organic compounds.
Water Res., 17, 173–183.
Hoigné, J., and Bader, H. (1983b). Rate constants of reactions of ozone with organic
and inorganic compounds in water. II: Dissociating organic compounds. Water
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Res., 17, 185–194.


Hoigné, J., Bader, H., Haag, W.R., and Staehelin, J. (1985). Rate constants of reac-
tions of ozone with organic and inorganic compounds in water. III. Inorganic
compounds and radicals. Water Res., 19, 993–1004.
Hsiao, Y.L., and Nobe, K. (1993). Oxidative reactions of phenol and chloroben-
zene with in situ electrogenerated Fenton’s reagent. Chem. Eng. Commun.,
126, 97–110.
Hu, J., and Wang, J.L. (2007). Degradation of chlorophenols in aqueous solution by
γ -radiation. Radiat. Phys. Chem., 76, 1489–1492.
Hu, J., Wang, J.L., and Chen, R. (2006). Degradation of 4-chlorophenol in aque-
ous solution by gamma-radiation and ozone oxidation. Sci. China, Ser. B, 49,
186–192.
Hua, I., Höchemer, R.H., and Hoffmann, M.R. (1995). Sonochemical degradation of
p-nitrophenol in a parallel-plate near-field acoustical processor. Environ. Sci.
Technol., 29, 2790–2796.
Huang, H.J., Li, D.Z., Lin, Q., Zhang, W.J., Shao, Y., Chen, Y.B., Sun, M., and Fu, X.Z.
(2009). Efficient degradation of benzene over LaVO4 /TiO2 nanocrystalline het-
erojunction photocatalyst under visible light irradiation. Environ. Sci. Technol.,
43, 4164–4168.
Huang, Y.H., Chou, S.S., Perng, M.G., Huang, G.H., and Cheng, S.S. (1999). Case
study on the bioeffluent of petrochemical wastewater by electro-Fenton method.
Water Sci. Technol., 39, 145–149.
Ikehata, K., and El-Din, M.G. (2005a). Aqueous pesticide degradation by ozonation
and ozone-based advanced oxidation processes: A review (Part I). Ozone Sci.
Eng., 27, 83–114.
Ikehata, K., and El-Din, M.G. (2005b). Aqueous pesticide degradation by ozonation
and ozone-based advanced oxidation processes: A review (Part II). Ozone Sci.
Eng., 27, 173–202.
Ikehata, K., Naghashkar, N.J., and Ei-Din, M.G. (2006). Degradation of aqueous
pharmaceuticals by ozonation and advanced oxidation processes: A review.
Ozone Sci. Eng., 28, 353–414.
314 J. L. Wang and L. J. Xu

Ince, N.H., Tezcanli, G., Belen, R.K., and Apikyan, İ.G. (2001). Ultrasound as a
catalyzer of aqueous reaction systems: The state of the art and environmental
applications. Appl. Catal. B: Environ., 29, 167–176.
Isariebel, Q.P., Carine, J.L., Ulises-Javier, J.H., Anne-Marie, W., and Henri, D.
(2009). Sonolysis of levodopa and paracetamol in aqueous solutions. Ultrason.
Sonochem., 16, 610–616.
Jaeger, C.D., and Bard, A.J. (1979). Spin trapping and electron-spin resonance de-
tection of radical intermediates in the photo-decomposition of water at TiO2
particulate systems. J. Phys. Chem., 83, 3146–3152.
Jiang, J., Pang, S.Y., and Ma, J. (2008a). Comment on “Factors affecting the yield
of oxidants from the reaction of nanoparticulate zero-valent iron and oxygen”.
Environ. Sci. Technol., 42, 5377.
Jiang, J., Pang, S.Y., and Ma, J. (2008b). Comment on “Polyoxometalate-enhanced
oxidation of organic compounds by nanoparticulate zero-valent iron and ferrous
Downloaded by [Tsinghua University] at 19:33 24 September 2012

ion in the presence of oxygen”. Environ. Sci. Technol., 42, 8167–8168.


Jonsson, A.M., Hallquist, M., and Ljungström, E. (2008). Influence of OH scavenger
on the water effect on secondary organic aerosol formation from ozonolysis of
limonene, 3-carene, and α-pinene. Environ. Sci. Technol., 42, 5938–5944.
Kallel, M., Belaid, C., Boussahel, R., Ksibi, M., Montiel, A., and Elleuch, B. (2009).
Olive mill wastewater degradation by Fenton oxidation with zero-valent iron
and hydrogen peroxide. J. Hazard. Mater., 163, 550–554.
Kasprzyk-Hordern, B., Ziółek, M., and Nawrocki, J. (2003). Catalytic ozonation and
methods of enhancing molecular ozone reactions in water treatment. Appl.
Catal. B: Environ., 46, 639–669.
Katsoyiannis, I.A., Ruettimann, T., and Hug, S.J. (2008). pH dependence of Fenton
reagent generation and As(III) oxidation and removal by corrosion of zero
valent iron in aerated water. Environ. Sci. Technol., 42, 7424–7430.
Katsoyiannis, I.A., Ruettimann, T., and Hug, S.J. (2009). Response to comment on “pH
dependence of Fenton reagent generation and As(III) oxidation and removal
by corrosion of zero valent iron in aerated water”. Environ. Sci. Technol., 43,
3980–3981.
Keenan, C.R., and Sedlak, D.L. (2008a). Comment on “Factors affecting the yield
of oxidants from the reaction of nanoparticulate zero-valent iron and oxygen”:
Response. Environ. Sci. Technol., 42, 5378–5378.
Keenan, C.R., and Sedlak, D.L. (2008b). Factors affecting the yield of oxidants from
the reaction of nanoparticulate zero-valent iron and oxygen. Environ. Sci. Tech-
nol., 42, 1262–1267.
Keenan, C.R., and Sedlak, D.L. (2008c). Ligand-enhanced reactive oxidant generation
by nanoparticulate zero-valent iron and oxygen. Environ. Sci. Technol., 42,
6936–6941.
Khataee, A.R., Vatanpour, V., and Ghadim, A.R.A. (2009). Decolorization of C.I.
Acid Blue 9 solution by UV/Nano-TiO2 , Fenton, Fenton-like, electro-Fenton
and electrocoagulation processes: A comparative study. J. Hazard. Mater., 161,
1225–1233.
Kibanova, D., Cervini-Silva, J., and Destaillats, H. (2009). Efficiency of clay-TiO2
nanocomposites on the photocatalytic elimination of a model hydrophobic air
pollutant. Environ. Sci. Technol., 43, 1500–1506.
Advanced Oxidation Processes for Wastewater Treatment 315

Kidak, R., and Ince, N.H. (2007). Catalysis of advanced oxidation reactions by ultra-
sound: A case study with phenol. J. Hazard. Mater., 146, 630–635.
Kim, C., Kim, J.T., Kim, K.S., Jeong, S., Kim, H.Y., and Han, Y.S. (2009). Immo-
bilization of TiO2 on an ITO substrate to facilitate the photoelectrochemical
degradation of an organic dye pollutant. Electrochim. Acta, 54, 5715–5720.
Kishimoto, N., Morita, Y., Tsuno, H., Oomura, T., and Mizutani, H. (2005). Ad-
vanced oxidation effect of ozonation combined with electrolysis. Water Res.,
39, 4661–4672.
Konstantinou, I.K., and Albanis, T.A. (2004). TiO2 -assisted photocatalytic degradation
of azo dyes in aqueous solution: Kinetic and mechanistic investigations: A
review. Appl. Catal. B: Environ., 49, 1–14.
Kubesch, K., Zona, R., Solar, S., and Gehringer, P. (2005). Degradation of catechol
by ionizing radiation, ozone and the combined process ozone-electron-beam.
Radiat. Phys. Chem., 72, 447–453.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Kusic, H., Koprivanac, N., and Bozic, A.L. (2006). Minimization of organic pollu-
tant content in aqueous solution by means of AOPs: UV- and ozone-based
technologies. Chem. Eng. J., 123, 127–137.
Kwon, S., Fan, M., Cooper, A.T., and Yang, H.Q. (2008). Photocatalytic applications
of micro- and nano-TiO2 in environmental engineering. Crit. Rev. Environ. Sci.
Technol., 38, 197–226.
Lapertot, M., Pulgarı́n, C., Fernández-Ibáñez, P., Maldonado, M.I., Pérez-Estrada, L.,
Oller, I., Gernjak, W., and Malato, S. (2006). Enhancing biodegradability of pri-
ority substances (pesticides) by solar photo-Fenton. Water Res., 40, 1086–1094.
Lau, T.K., Chu, W., and Graham, N. (2007). Reaction pathways and kinetics of
butylated hydroxyanisole with UV, ozonation, and UV/O3 processes. Water
Res., 41, 765–774.
Lee, C.H., Keenan, C.R., and Sedlak, D.L. (2008a). Polyoxometalate-enhanced oxi-
dation of organic compounds by nanoparticulate zero-valent iron and ferrous
ion in the presence of oxygen. Environ. Sci. Technol., 42, 4921–4926.
Lee, C.H., Keenan, C.R., and Sedlak, D.L. (2008b). Response to comment on
“Polyoxometalate-enhanced oxidation of organic compounds by nanopartic-
ulate zero-valent iron and ferrous ion in the presence of oxygen”. Environ. Sci.
Technol., 42, 8169.
Lee, C.H., and Sedlak, D.L. (2008). Enhanced formation of oxidants from bimetallic
nickel-iron nanoparticles in the presence of oxygen. Environ. Sci. Technol., 42,
8528–8533.
Lee, C.H., and Yoon, J. (2004). Application of photoactivated periodate to the decol-
orization of reactive dye: Reaction parameters and mechanism. J. Photochem.
Photobiol. A: Chem., 165, 35–41.
Legrini, O., Oliveros, E., and Braun, A.M. (1993). Photochemical processes for water
treatment. Chem. Rev., 93, 671–698.
Lesko, T., Colussi, A.J., and Hoffmann, M.R. (2006). Sonochemical decomposition of
phenol: Evidence for a synergistic effect of ozone and ultrasound for the elimi-
nation of total organic carbon from water. Environ. Sci. Technol., 40, 6818–6823.
Ley, S.V., and Low, C.M.R. (1989). Ultrasound in synthesis, Sgringer-Verlag: New
York.
316 J. L. Wang and L. J. Xu

Li, D.P., and Qu, J.H. (2009). The progress of catalytic technologies in water purifi-
cation: A review. J. Environ. Sci.-China, 21, 713–719.
Li, H.Y., Qu, J.H., and Liu, H.J. (2006). Removal of a type of endocrine disruptors: di-
n-butyl phthalate from water by ozonation. J. Environ. Sci.-China, 18, 845–851.
Li, K., and Crittenden, J. (2009). Computerized pathway elucidation for hydroxyl
radical-induced chain reaction mechanisms in aqueous phase advanced oxida-
tion processes. Environ. Sci. Technol., 43, 2831–2837.
Li, Q., Page, M.A., Mariñas, B.J., and Shang, J.K. (2008). Treatment of coliphage MS2
with palladium-modified nitrogen-doped titanium oxide photocatalyst illumi-
nated by visible light. Environ. Sci. Technol., 42, 6148–6153.
Li, X.Z., Chen, C.C., and Zhao, J.C. (2001). Mechanism of photodecomposition of
H2 O2 on TiO2 surfaces under visible light irradiation. Langmuir, 17, 4118–4122.
Lin, K.J., Cooper, W.J., Nickelsen, M.G., Kurucz, C.N., and Waite, T.D. (1995). De-
composition of aqueous solutions of phenol using high energy electron beam
Downloaded by [Tsinghua University] at 19:33 24 September 2012

irradiation: A large scale study. Appl. Radiat. Isot., 46, 1307–1316.


Linsebigler, A.L., Lu, G.Q., and Yates, J.T. (1995). Photocatalysis on TiO2 surfaces:
Principles, mechanisms, and selected results. Chem. Rev., 95, 735–758.
Liu, H., Wang, C., Li, X.Z., Xuan, X.L., Jiang, C.C., and Cui, H.N. (2007). A novel
electro-Fenton process for water treatment: Reaction-controlled pH adjustment
and performance assessment. Environ. Sci. Technol., 41, 2937–2942.
Liu, Y., Liu, H.L., Ma, J., and Wang, X. (2009). Comparison of degradation mecha-
nism of electrochemical oxidation of di- and tri-nitrophenols on Bi-doped lead
dioxide electrode: Effect of the molecular structure. Appl. Catal. B: Environ., 91,
284–299.
Logager, T., Holcman, J., Sehested, K., and Pedersen, T. (1992). Oxidation of ferrous
ions by ozone in acidic solutions. Inorg. Chem., 31, 3523–3529.
Ma, H.J., Wang, M., Yang, R.Y., Wang, W.F., Zhao, J., Shen, Z.Q., and Yao, S.D.
(2007). Radiation degradation of Congo Red in aqueous solution. Chemosphere,
68, 1098–1104.
Ma, J., and Graham, N.J.D. (1997). Preliminary investigation of manganese-catalyzed
ozonation for the destruction of atrazine. Ozone Sci. Eng., 19, 227–240.
Ma, J.H., Song, W.J., Chen, C.C., Ma, W.H., Zhao, J.C., and Tang, Y.L. (2005). Fenton
degradation of organic compounds promoted by dyes under visible irradiation.
Environ. Sci. Technol., 39, 5810–5815.
Mack, J., and Bolton, J.R. (1999). Photochemistry of nitrite and nitrate in aqueous
solution: A review. J. Photochem. Photobiol. A: Chem., 128, 1–13.
Malpass, G.R.P., Miwa, D.W., Miwa, A.C.P., Machado, S.A.S., and Motheo, A.J.
(2007). Photo-assisted electrochemical oxidation of atrazine on a commercial
Ti/Ru0.3 Ti0.7 O2 DSA electrode. Environ. Sci. Technol., 41, 7120–7125.
Martin, S.T., Herrmann, H., Choi, W.Y., and Hoffmann, M.R. (1994a). Time-resolved
microwave conductivity. 1. TiO2 photoreactivity and size quantization. J. Chem.
Soc., Faraday Trans., 90, 3315–3322.
Martin, S.T., Herrmann, H., and Hoffmann, M.R. (1994b). Time-resolved microwave
conductivity. 2. Quantum-sized TiO2 and the effect of adsorbates and light-
intensity on charge-carrier dynamics. J. Chem. Soc., Faraday Trans., 90,
3323–3330.
Advanced Oxidation Processes for Wastewater Treatment 317

Martin, S.T., Lee, A.T., and Hoffmann, M.R. (1995). Chemical mechanism of inorganic
oxidants in the TiO2 /UV process: Increased rates of degradation of chlorinated
hydrocarbons. Environ. Sci. Technol., 29, 2567–2573.
Martı́nez-Huitle, C.A., and Brillas, E. (2009). Decontamination of wastewaters con-
taining synthetic organic dyes by electrochemical methods: A general review.
Appl. Catal. B: Environ., 87, 105–145.
Martı́nez-Huitle, C.A., De Battisti, A., Ferro, S., Reyna, S., Cerro-López, M., and Quiro,
M.A. (2008). Removal of the pesticide methamidophos from aqueous solutions
by electrooxidation using Pb/PbO2 , Ti/SnO2 , and Si/BDD electrodes. Environ.
Sci. Technol., 42, 6929–6935.
Martı́nez-Huitle, C.A., and Ferro, S. (2006). Electrochemical oxidation of organic
pollutants for the wastewater treatment: Direct and indirect processes. Chem.
Soc. Rev., 35, 1324–1340.
Masarwa, A., Rachmilovich-Calis, S., Meyerstein, N., and Meyerstein, D. (2005). Oxi-
Downloaded by [Tsinghua University] at 19:33 24 September 2012

dation of organic substrates in aerated aqueous solutions by the Fenton reagent.


Coord. Chem. Rev., 249, 1937–1943.
Mason, T.J., and Lorimer, J.P. (1988). Sonochemistry: Theory, applications and uses
of ultrasound in chemistry, New York: Ellis Horwood Ltd.
Matthews, R.W. (1984). Hydroxylation reactions induced by near-ultraviolet pho-
tocatalysis of aqueous titanium dioxide suspensions. J. Chem. Soc., Faraday
Trans. I, 80, 457–471.
Mazille, F., Schoettl, T., and Pulgarin, C. (2009). Synergistic effect of TiO2 and iron
oxide supported on fluorocarbon films. Part 1: Effect of preparation parameters
on photocatalytic degradation of organic pollutant at neutral pH. Appl. Catal.
B: Environ., 89, 635–644.
McKay, D.J., and Wright, J.S. (1998). How long can you make an oxygen chain? J.
Am. Chem. Soc., 120, 1003–1013.
Mead, E.L., Sutherland, R.G., and Verrall, R.E. (1976). The effect of ultrasound on
water in the presence of dissolved gases. Can. J. Chem./Rev. Can. Chim., 54,
1114–1120.
Melero, J.A., Martı́nez, F., Botas, J.A., Molina, R., and Pariente, M.I. (2009). Heteroge-
neous catalytic wet peroxide oxidation systems for the treatment of an industrial
pharmaceutical wastewater. Water Res., 43, 4010–4018.
Mezyk, S.P., Jones, J., Cooper, W.J., Tobien, T., Nickelsen, M.G., Adams, J.W., O’Shea,
K.E., Bartels, D.M., Wishart, J.F., Tornatore, P.M., Newman, K.S., Gregoire, K.,
and Weidman, D.J. (2004). Radiation chemistry of methyl tert-butyl ether in
aqueous solution. Environ. Sci. Technol., 38, 3994–4001.
Michaud, P.A., Mahé, E., Haenni, W., Perret, A., and Comninellis, C. (2000). Prepara-
tion of peroxodisulfuric acid using boron-doped diamond thin film electrodes.
Electrochem. Solid-State Lett., 3, 77–79.
Michaud, P.A., Panizza, M., Ouattara, L., Diaco, T., Foti, G., and Comninellis, C.
(2003). Electrochemical oxidation of water on synthetic boron-doped diamond
thin film anodes. J. Appl. Electrochem., 33, 151–154.
Mills, A., and LeHunte, S. (1997). An overview of semiconductor photocatalysis. J.
Photochem. Photobiol. A: Chem., 108, 1–35.
318 J. L. Wang and L. J. Xu

Moriwaki, H., Takagi, Y., Tanaka, M., Tsuruho, K., Okitsu, K., and Maeda, Y. (2005).
Sonochemical decomposition of perfluorooctane sulfonate and perfluorooc-
tanoic acid. Environ. Sci. Technol., 39, 3388–3392.
Mozumder, A. (1999). Fundamentals of radiation chemistry, 1st edition, Academic
Press, New York, USA.
Muthuvel, I., and Swaminathan, M. (2008). Highly solar active Fe(III) immobilised
alumina for the degradation of Acid Violet 7. Sol. Energy Mater. Sol. Cells, 92,
857–863.
Nélieu, S., Perreau, F., Bonnemoy, F., Ollitrault, M., Azam, D., Lagadic, L., Bohatier,
J., and Einhorn, J. (2009). Sunlight nitrate-induced photodegradation of chloro-
toluron: Evidence of the process in aquatic mesocosms. Environ. Sci. Technol.,
43, 3148–3154.
Nemes, A., Fábián, I., and Gordon, G. (2000a). Experimental aspects of mechanistic
studies on aqueous ozone decomposition in alkaline solution. Ozone Sci. Eng.,
Downloaded by [Tsinghua University] at 19:33 24 September 2012

22, 287–304.
Nemes, A., Fábián, I., and Gordon, G. (2000b). The kinetics and mechanism of aque-
ous ozone decomposition in alkaline solution. Inorg. React. Mech., 2, 327–341.
Neppolian, B., Doronila, A., Grieser, F., and Ashokkumar, M. (2009). Simple and
efficient sonochemical method for the oxidation of arsenic(III) to arsenic(V).
Environ. Sci. Technol., 43, 6793–6798.
Neyens, E., and Baeyens, J. (2003). A review of classic Fenton’s peroxidation as an
advanced oxidation technique. J. Hazard. Mater., 98, 33–50.
Noltingk, B.E., and Neppiras, E.A. (1950). Cavitation produced by ultrasonics. Proc.
Phys. Soc. B, 63, 674–685.
Noubactep, C. (2009). Comment on “pH dependence of Fenton reagent generation
and As(III) oxidation and removal by corrosion of zero valent iron in aerated
water”. Environ. Sci. Technol., 43, 233.
Nowell, L.H., and Hoigné, J. (1992a). Photolysis of aqueous chlorine at sunlight and
ultraviolet wavelengths. I. Degradation rates. Water Res., 26, 593–598.
Nowell, L.H., and Hoigné, J. (1992b). Photolysis of aqueous chlorine at sunlight
and ultraviolet wavelengths. II. Hydroxyl radical production. Water Res., 26,
599–605.
Ntampegliotis, K., Riga, A., Karayannis, V., Bontozoglou, V., and Papapolymerou, G.
(2006). Decolorization kinetics of Procion H-exl dyes from textile dyeing using
Fenton-like reactions. J. Hazard. Mater., 136, 75–84.
Okamoto, K., Yamamoto, Y., Tanaka, H., and Itaya, A. (1985a). Kinetics of hetero-
geneous photocatalytic decomposition of phenol over anatase TiO2 powder.
Bull. Chem. Soc. Jpn., 58, 2023–2028.
Okamoto, K., Yamamoto, Y., Tanaka, H., Tanaka, M., and Itaya, A. (1985b). Het-
erogeneous photocatalytic decomposition of phenol over TiO2 powder. Bull.
Chem. Soc. Jpn., 58, 2015–2022.
Oliver, B.G., and Carey, J.H. (1977). Photochemical production of chlorinated or-
ganics in aqueous solutions containing chlorine. Environ. Sci. Technol., 11,
893–895.
Ormad, P., Cortés, S., Puig, A., and Ovelleiro, J.L. (1997). Degradation of organochlo-
ride compounds by O3 and O3 /H2 O2 . Water Res., 31, 2387–2391.
Advanced Oxidation Processes for Wastewater Treatment 319

Pálfi, T., Takács, E., and Wojnárovits, L. (2007). Degradation of H-acid and its deriva-
tive in aqueous solution by ionising radiation. Water Res., 41, 2533–2540.
Panizza, M., and Cerisola, G. (2005). Application of diamond electrodes to electro-
chemical processes. Electrochim. Acta, 51, 191–199.
Panizza, M., Delucchi, M., and Cerisola, G. (2005). Electrochemical degradation of
anionic surfactants. J. Appl. Electrochem., 35, 357–361.
Park, W., Hwang, M.H., Kim, T.H., Lee, M.J., and Kim, I.S. (2009). Enhancement
in characteristics of sewage sludge and anaerobic treatability by electron beam
pre-treatment. Radiat. Phys. Chem., 78, 124–129.
Peller, J., Wiest, O., and Kamat, P.V. (2001). Sonolysis of 2,4-dichlorophenoxyacetic
acid in aqueous solutions. Evidence for ·OH-radical-mediated degradation. J.
Phys. Chem. A, 105, 3176–3181.
Pera-Titus, M., Garcı́a-Molina, V., Baños, M.A., Giménez, J., and Esplugas, S. (2004).
Degradation of chlorophenols by means of advanced oxidation processes: a
Downloaded by [Tsinghua University] at 19:33 24 September 2012

general review. Appl. Catal. B: Environ., 47, 219–256.


Peternel, I., Koprivanac, N., and Kusic, H. (2006). UV-based processes for reactive
azo dye mineralization. Water Res., 40, 525–532.
Pétrier, C., Lamy, M.F., Francony, A., Benahcene, A., David, B., Renaudin, V., and
Gondrexon, N. (1994). Sonochemical degradation of phenol in dilute aqueous
solutions: Comparison of the reaction rates at 20 and 487 kHz. J. Phys. Chem.,
98, 10514–10520.
Pi, Y.Z., Zhang, L.S., and Wang, J.L. (2007). The formation and influence of hydro-
gen peroxide during ozonation of para-chlorophenol. J. Hazard. Mater., 141,
707–712.
Piera, E., Calpe, J.C., Brillas, E., Domènech, X., and Peral, J. (2000). 2,4-
Dichlorophenoxyacetic acid degradation by catalyzed ozonation: TiO2 /UVA/O3
and Fe(II)/UVA/O3 systems. Appl. Catal. B: Environ., 27, 169–177.
Pignatello, J.J., Liu, D., and Huston, P. (1999). Evidence for an additional oxidant in
the photoassisted Fenton reaction. Environ. Sci. Technol., 33, 1832–1839.
Pignatello, J.J., Oliveros, E., and MacKay, A. (2006). Advanced oxidation processes
for organic contaminant destruction based on the Fenton reaction and related
chemistry. Crit. Rev. Environ. Sci. Technol., 36, 1–84.
Pines, D.S., and Reckhow, D.A. (2002). Effect of dissolved cobalt(II) on the ozonation
of oxalic acid. Environ. Sci. Technol., 36, 4046–4051.
Pirgalioğlu, S., and Özbelge, T.A. (2009). Comparison of non-catalytic and catalytic
ozonation processes of three different aqueous single dye solutions with respect
to powder copper sulfide catalyst. Appl. Catal., A, 363, 157–163.
Popov, P., and Getoff, N. (2004). Ozonolysis and combination of ozonolysis and
radiolysis of aqueous fluorene. Radiat. Phys. Chem., 69, 311–315.
Poulios, I., and Tsachpinis, I. (1999). Photodegradation of the textile dye Reactive
Black 5 in the presence of semiconducting oxides. J. Chem. Technol. Biotechnol.,
74, 349–357.
Pradhan, A.A., and Gogate, P.R. (2010). Degradation of p-nitrophenol using acoustic
cavitation and Fenton chemistry. J. Hazard. Mater., 173, 517–522.
Pratap, K., and Lemley, A.T. (1994). Electrochemical peroxide treatment of aqueous
herbicide solutions. J. Agric. Food. Chem., 42, 209–215.
320 J. L. Wang and L. J. Xu

Rakitskaya, T.L., Ennan, A.A., Granatyuk, I.V., Bandurko, A.Y., Balavoine, G.G.A.,
Geletii, Y.V., and Paina, V.Y. (1999). Kinetics and mechanism of low-
temperature ozone decomposition by Co-ions adsorbed on silica. Catal. Today,
53, 715–723.
Rao, Y.F., and Chu, W. (2009). Reaction mechanism of linuron degradation in TiO2
suspension under visible light irradiation with the assistance of H2 O2 . Environ.
Sci. Technol., 43, 6183–6189.
Rauf, M.A., and Ashraf, S.S. (2009). Radiation induced degradation of dyes–An
overview. J. Hazard. Mater., 166, 6–16.
Rauf, M.A., Bukallah, S.B., Hamadi, A., Sulaiman, A., and Hammadi, F. (2007). The
effect of operational parameters on the photoinduced decoloration of dyes
using a hybrid catalyst V2 O5 /TiO2 . Chem. Eng. J., 129, 167–172.
Rayleigh. (1917). On the pressure developed in a liquid during the collapse of a
spherical cavity. Philos. Mag., 34, 94–98.
Downloaded by [Tsinghua University] at 19:33 24 September 2012

Reisz, E., Schmidt, W., Schuchmann, H.P., and von Sonntag, C. (2003). Photolysis
of ozone in aqueous solutions in the presence of tertiary butanol. Environ. Sci.
Technol., 37, 1941–1948.
Rengifo-Herrera, J.A., Pierzchala, K., Sienkiewicz, A., Forró, L., Kiwi, J., and Pulgarin,
C. (2009). Abatement of organics and Escherichia coli by N, S co-doped TiO2
under UV and visible light. Implications of the formation of singlet oxygen (1O2 )
under visible light. Appl. Catal. B: Environ., 88, 398–406.
Rosenfeldt, E.J., Linden, K.G., Canonica, S., and von Gunten, U. (2006). Comparison
of the efficiency of ·OH radical formation during ozonation and the advanced
oxidation processes O3 /H2 O2 and UV/H2 O2 . Water Res., 40, 3695–3704.
Rothenberger, G., Moser, J., Grätzel, M., Serpone, N., and Sharma, D.K. (1985).
Charge carrier trapping and recombination dynamics in small semiconductor
particles. J. Am. Chem. Soc., 107, 8054–8059.
Rothschild, W.G., and Allen, A.O. (1958). Studies in the radiolysis of ferrous sulfate
solutions: III. Air-free solutions at higher pH. Radiat. Res., 8, 101–110.
Rupa, A.V., Manikandan, D., Divakar, D., and Sivakumar, T. (2007). Effect of de-
position of Ag on TiO2 nanoparticles on the photodegradation of Reactive
Yellow-17. J. Hazard. Mater., 147, 906–913.
Safarzadeh-Amiri, A., Bolton, R.J., and Cater, S.R. (1996). The use of iron in advanced
oxidation processes. J. Adv. Oxid. Technol., 1, 18.
San, N., Hatipoğlu, A., Koçtürk, G., and Çinar, Z. (2001). Prediction of primary
intermediates and the photodegradation kinetics of 3-aminophenol in aqueous
TiO2 suspensions. J. Photochem. Photobiol. A: Chem., 139, 225–232.
Sánchez-Polo, M., López-Peñalver, J., Prados-Joya, G., Ferro-Garcı́a, M.A., and
Rivera-Utrilla, J. (2009). Gamma irradiation of pharmaceutical compounds,
nitroimidazoles, as a new alternative for water treatment. Water Res., 43,
4028–4036.
Sanchez-Prado, L., Barro, R., Garcia-Jares, C., Llompart, M., Lores, M., Petrakis, C.,
Kalogerakis, N., Mantzavinos, D., and Psillakis, E. (2008). Sonochemical degra-
dation of triclosan in water and wastewater. Ultrason. Sonochem., 15, 689–694.
Satyawali, Y., and Balakrishnan, M. (2008). Wastewater treatment in molasses-based
alcohol distilleries for COD and color removal: A review. J. Environ. Manage.,
86, 481–497.
Advanced Oxidation Processes for Wastewater Treatment 321

Sehested, K., Corfitzen, H., Holcman, J., and Hart, E.J. (1998). On the mechanism of
the decomposition of acidic O3 solutions, thermally or H2 O2 -initiated. J. Phys.
Chem. A, 102, 2667–2672.
Sehested, K., Holcman, J., and Hart, E.J. (1983). Rate constants and products of the
reactions of eaq −, O2 −, and H with ozone in aqueous solutions. J. Phys. Chem.,
87, 1951–1954.
Serpone, N., and Colarusso, P. (1994). Sonochemistry. 1. Effects of ultrasounds on
heterogeneous chemical reactions: A useful tool to generate radicals and to
examine reaction mechanisms. Res. Chem. Intermed., 20, 635–679.
Serpone, N., Terzian, R., Hidaka, H., and Pelizzetti, E. (1994). Ultrasonic induced
dehalogenation and oxidation of 2-, 3-, and 4-chlorophenol in air-equilibrated
aqueous media. Similarities with irradiated semiconductor particulates. J. Phys.
Chem., 98, 2634–2640.
Serrano, K., Michaud, P.A., Comninellis, C., and Savall, A. (2002). Electrochemical
Downloaded by [Tsinghua University] at 19:33 24 September 2012

preparation of peroxodisulfuric acid using boron doped diamond thin film


electrodes. Electrochim. Acta, 48, 431–436.
Shankar, M.V., Nélieu, S., Kerhoas, L., and Einhorn, J. (2008). Natural sunlight
NO3 −/NO2 −-induced photo-degradation of phenylurea herbicides in water.
Chemosphere, 71, 1461–1468.
Shin, S., Yoon, H., and Jang, J. (2008). Polymer-encapsulated iron oxide nanoparti-
cles as highly efficient Fenton catalysts. Catal. Commun., 10, 178–182.
Siedlecka, E.M., Mrozik, W., Kaczyński, Z., and Stepnowski, P. (2008). Degradation
of 1-butyl-3-methylimidazolium chloride ionic liquid in a Fenton-like system. J.
Hazard. Mater., 154, 893–900.
Sirtori, C., Zapata, A., Oller, I., Gernjak, W., Agüera, A., and Malato, S. (2009). Solar
photo-Fenton as finishing step for biological treatment of a pharmaceutical
wastewater. Environ. Sci. Technol., 43, 1185–1191.
Sivasankar, T., and Moholkar, V.S. (2009a). Mechanistic approach to intensification
of sonochemical degradation of phenol. Chem. Eng. J., 149, 57–69.
Sivasankar, T., and Moholkar, V.S. (2009b). Physical insights into the sonochemical
degradation of recalcitrant organic pollutants with cavitation bubble dynamics.
Ultrason. Sonochem., 16, 769–781.
Skoumal, M., Rodrı́guez, R.M., Cabot, P.L., Centellas, F., Garrido, J.A., Arias, C.,
and Brillas, E. (2009). Electro-Fenton, UVA photoelectro-Fenton and solar
photoelectro-Fenton degradation of the drug ibuprofen in acid aqueous medium
using platinum and boron-doped diamond anodes. Electrochim. Acta, 54,
2077–2085.
Song, S., He, Z.Q., Qiu, J.P., Xu, L.J., and Chen, J.M. (2007a). Ozone assisted elec-
trocoagulation for decolorization of C.I. Reactive Black 5 in aqueous solution:
An investigation of the effect of operational parameters. Sep. Purif. Technol.,
55, 238–245.
Song, S., Tu, J.J., Xu, L.J., Xu, X., He, Z.Q., Qiu, J.P., Ni, J.G., and Chen, J.M.
(2008a). Preparation of a titanium dioxide photocatalyst codoped with cerium
and iodine and its performance in the degradation of oxalic acid. Chemosphere,
73, 1401–1406.
322 J. L. Wang and L. J. Xu

Song, S., Xu, L.J., He, Z.Q., Chen, J.M., Xiao, X.Z., and Yan, B. (2007b). Mechanism
of the photocatalytic degradation of C.I. Reactive Black 5 at pH 12.0 using
SrTiO3 /CeO2 as the catalyst. Environ. Sci. Technol., 41, 5846–5853.
Song, S., Xu, L.J., He, Z.Q., Ying, H.P., Chen, J.M., Xiao, X.Z., and Yan, B. (2008b).
Photocatalytic degradation of C.I. Direct Red 23 in aqueous solutions under UV
irradiation using SrTiO3 /CeO2 composite as the catalyst. J. Hazard. Mater., 152,
1301–1308.
Song, S., Xu, X., Xu, L.J., He, Z.Q., Ying, H.P., Chen, J.M., and Yan, B. (2008c).
Mineralization of CI Reactive Yellow 145 in aqueous solution by ultraviolet-
enhanced ozonation. Ind. Eng. Chem. Res., 47, 1386–1391.
Song, S., Ying, H.P., He, Z.Q., and Chen, J.M. (2007c). Mechanism of decolorization
and degradation of CI Direct Red 23 by ozonation combined with sonolysis.
Chemosphere, 66, 1782–1788.
Song, S., Zhan, L.Y., He, Z.Q., Lin, L.L., Tu, J.J., Zhang, Z.H., Chen, J.M., and Xu,
Downloaded by [Tsinghua University] at 19:33 24 September 2012

L.J. (2010). Mechanism of the anodic oxidation of 4-chloro-3-methyl phenol in


aqueous solution using Ti/SnO2 -Sb/PbO2 electrodes. J. Hazard. Mater., 175,
614–621.
Spinks, J.W.T., and Woods, R.J. (1990). An introduction to radiation chemistry, third
ed. Wiley-Interscience, New York.
Staehelin, J., Bühler, R.E., and Hoigné, J. (1984). Ozone decomposition in water
studied by pulse radiolysis. 2. Hydroxyl and hydrogen tetroxide (HO4 ) as chain
intermediates. J. Phys. Chem., 88, 5999–6004.
Staehelin, J., and Hoigné, J. (1982). Decomposition of ozone in water: Rate of ini-
tiation by hydroxide ions and hydrogen peroxide. Environ. Sci. Technol., 16,
676–681.
Staehelin, J., and Hoigné, J. (1985). Decomposition of ozone in water in the presence
of organic solutes acting as promoters and inhibitors of radical chain reactions.
Environ. Sci. Technol., 19, 1206–1213.
Sudoh, M., Kodera, T., Sakai, K., Zhang, J.Q., and Koide, K. (1986). Oxidative degra-
dation aqueous phenol effluent with electrogenerated Fenton’s reagent. J. Chem.
Eng. Jpn., 19, 513–518.
Sun, Y.F., and Pignatello, J.J. (1993). Photochemical reactions involved in the total
mineralization of 2,4-D by Fe3+/H2 O2 /UV. Environ. Sci. Technol., 27, 304–310.
Suslick, K.S. (1989). The chemical effect of ultrasound. Sci. Am., 260, 80–86.
Suslick, K.S. (1990). Sonochemistry. Science, 247, 1439–1445.
Szpyrkowicz, L., Juzzolino, C., Kaul, S.N., Daniele, S., and De Faveri, M.D. (2000).
Electrochemical oxidation of dyeing baths bearing disperse dyes. Ind. Eng.
Chem. Res., 39, 3241–3248.
Tai, C., Gu, X.X., Zou, H., and Guo, Q.H. (2002). A new simple and sensitive fluo-
rometric method for the determination of hydroxyl radical and its application.
Talanta, 58, 661–667.
Tezcanli-Güyer, G., and Ince, N.H. (2004). Individual and combined effects of ul-
trasound, ozone and UV irradiation: a case study with textile dyes. Ultrasonics,
42, 603–609.
Tomiyasu, H., Fukutomi, H., and Gordon, G. (1985). Kinetics and mechanism of
ozone decomposition in basic aqueous solution. Inorg. Chem., 24, 2962–2966.
Advanced Oxidation Processes for Wastewater Treatment 323

Vinu, R., and Madras, G. (2008). Kinetics of simultaneous photocatalytic degradation


of phenolic compounds and reduction of metal ions with nano-TiO2 . Environ.
Sci. Technol., 42, 913–919.
Vogt, R., and Schindler, R.N. (1992). Product channels in the photolysis of HOCl. J.
Photochem. Photobiol. A: Chem., 66, 133–140.
Walling, C. (1975). Fenton’s reagent revisited. Acc. Chem. Res., 8, 125–131.
Wang, H., and Wang, J.L. (2007a). Electrochemical degradation of 4-chlorophenol
using a novel Pd/C gas-diffusion electrode. Appl. Catal. B: Environ., 77, 58–65.
Wang, H., and Wang, J.L. (2008a). The cooperative electrochemical oxidation of
chlorophenols in anode-cathode compartments. J. Hazard. Mater., 154, 44–50.
Wang, H., and Wang, J.L. (2008b). Electrochemical degradation of 2,4-
dichlorophenol on a palladium modified gas-diffusion electrode. Electrochim.
Acta, 53, 6402–6409.
Wang, H., and Wang, J.L. (2009a). Comparative study on electrochemical degradation
Downloaded by [Tsinghua University] at 19:33 24 September 2012

of 2,4-dichlorophenol by different Pd/C gas-diffusion cathodes. Appl. Catal. B:


Environ., 89, 111–117.
Wang, H., and Wang, J.L. (2009b). Electrochemical degradation of pentachlorophe-
nol on a palladium modified gas-diffusion electrode. Water Sci. Technol., 59,
1759–1767.
Wang, J.L., and Wang, J.Z. (2007b). Application of radiation technology to sewage
sludge processing: A review. J. Hazard. Mater., 143, 2–7.
Wang, Y.B., and Hong, C.S. (1999). Effect of hydrogen peroxide, periodate and
persulfate on photocatalysis of 2-chlorobiphenyl in aqueous TiO2 suspensions.
Water Res., 33, 2031–2036.
Wang, Z.H., Ma, W.H., Chen, C.C., and Zhao, J.C. (2008). Photochemical coupling
reactions between Fe(III)/Fe(II), Cr(VI)/Cr(III), and polycarboxylates: Inhibitory
effect of Cr species. Environ. Sci. Technol., 42, 7260–7266.
Warneck, P., and Wurzinger, C. (1988). Product quantum yields for the 305-nm pho-
todecomposition of NO3 − in aqueous solution. J. Phys. Chem., 92, 6278–6283.
Wasiewicz, M., Chmielewski, A.G., and Getoff, N. (2006). Radiation-induced degra-
dation of aqueous 2,3-dihydroxynaphthalene. Radiat. Phys. Chem., 75, 201–209.
Watanabe, K., Menzel, D., Nilius, N., and Freund, H.J. (2006). Photochemistry on
metal nanoparticles. Chem. Rev., 106, 4301–4320.
Watts, M.J., and Linden, K.G. (2007). Chlorine photolysis and subsequent OH radical
production during UV treatment of chlorinated water. Water Res., 41, 2871–2878.
Weavers, L.K., and Hoffmann, M.R. (1998). Sonolytic decomposition of ozone in
aqueous solution: Mass transfer effects. Environ. Sci. Technol., 32, 3941–3947.
Wojnárovits, L., and Takács, E. (2008). Irradiation treatment of azo dye containing
wastewater: An overview. Radiat. Phys. Chem., 77, 225–244.
Wu, C.H. (2009). Photodegradation of C.I. Reactive Red 2 in UV/TiO2 -based systems:
Effects of ultrasound irradiation. J. Hazard. Mater., 167, 434–439.
Wu, C.H., Kuo, C.Y., and Chang, C.L. (2008). Homogeneous catalytic ozonation of
C.I. Reactive Red 2 by metallic ions in a bubble column reactor. J. Hazard.
Mater., 154, 748–755.
Wu, K.Q., Xie, Y.D., Zhao, J.C., and Hidaka, H. (1999a). Photo-Fenton degradation
of a dye under visible light irradiation. J. Mol. Catal. A: Chem., 144, 77–84.
324 J. L. Wang and L. J. Xu

Wu, T.X., Lin, T., Zhao, J.C., Hidaka, H., and Serpone, N. (1999b). TiO2 -assisted
photodegradation of dyes. 9. Photooxidation of a squarylium cyanine dye in
aqueous dispersions under visible light irradiation. Environ. Sci. Technol., 33,
1379–1387.
Wu, T.X., Liu, G.M., Zhao, J.C., Hidaka, H., and Serpone, N. (1999c). Evidence for
H2 O2 generation during the TiO2 -assisted photodegradation of dyes in aqueous
dispersions under visible light illumination. J. Phys. Chem. B, 103, 4862–4867.
Xiao, H., Liu, R.P., Zhao, X., and Qu, J.H. (2008). Enhanced degradation of 2,4-
dinitrotoluene by ozonation in the presence of manganese(II) and oxalic acid.
J. Mol. Catal. A: Chem., 286, 149–155.
Xing, S.T., Hu, C., Qu, J.H., He, H., and Yang, M. (2008). Characterization and
reactivity of MnOx supported on mesoporous zirconia for herbicide 2,4-D min-
eralization with ozone. Environ. Sci. Technol., 42, 3363–3368.
Xue, J., and Wang, J.L. (2008). Radiolysis of pentachlorophenol (PCP) in aqueous
Downloaded by [Tsinghua University] at 19:33 24 September 2012

solution by gamma radiation. J. Environ. Sci.-China, 20, 1153–1157.


Xue, X.F., Hanna, K., Abdelmoula, M., and Deng, N.S. (2009). Adsorption and oxida-
tion of PCP on the surface of magnetite: Kinetic experiments and spectroscopic
investigations. Appl. Catal. B: Environ., 89, 432–440.
Yang, L., Yu, L.E., and Ray, M.B. (2008). Degradation of paracetamol in aqueous
solutions by TiO2 photocatalysis. Water Res., 42, 3480–3488.
Yang, L.M., Yu, L.E., and Ray, M.B. (2009). Photocatalytic oxidation of paraceta-
mol: Dominant reactants, intermediates, and reaction mechanisms. Environ.
Sci. Technol., 43, 460–465.
Yang, S.Y., Lou, L.P., Wang, K., and Chen, Y.X. (2006). Shift of initial mechanism in
TiO2 -assisted photocatalytic process. Appl. Catal., A, 301, 152–157.
Yoon, S.H., and Lee, J.H. (2005). Oxidation mechanism of As(III) in the UV/TiO2
system: Evidence for a direct hole oxidation mechanism. Environ. Sci. Technol.,
39, 9695–9701.
Yu, S.H., Lee, B.J., Lee, M.J., Cho, I.H., and Chang, S.W. (2008). Decomposition
and mineralization of cefaclor by ionizing radiation: Kinetics and effects of the
radical scavengers. Chemosphere, 71, 2106–2112.
Zhang, S.J., Jiang, H., Li, M.J., Yu, H.Q., Yin, H., and Li, Q.R. (2007). Kinetics and
mechanisms of radiolytic degradation of nitrobenzene in aqueous solutions.
Environ. Sci. Technol., 41, 1977–1982.
Zhang, T., Li, C.J., Ma, J., Tian, H., and Qiang, Z.M. (2008). Surface hydroxyl groups of
synthetic α-FeOOH in promoting ·OH generation from aqueous ozone: Property
and activity relationship. Appl. Catal. B: Environ., 82, 131–137.
Zhang, T., and Ma, J. (2008). Catalytic ozonation of trace nitrobenzene in water with
synthetic goethite. J. Mol. Catal. A: Chem., 279, 82–89.
Zhang, X.W., Lei, L.C., Xia, B., Zhang, Y., and Fu, J.L. (2009). Oxidization of carbon
nanotubes through hydroxyl radical induced by pulsed O2 plasma and its ap-
plication for O2 reduction in electro-Fenton. Electrochim. Acta, 54, 2810–2817.
Zhao, J.C., Chen, C.C., and Ma, W.H. (2005a). Photocatalytic degradation of organic
pollutants under visible light irradiation. Top. Catal., 35, 269–278.
Zhao, J.C., Wu, T.X., Wu, K.Q., Oikawa, K., Hidaka, H., and Serpone, N. (1998).
Photoassisted degradation of dye pollutants. 3. Degradation of the cationic dye
rhodamine B in aqueous anionic surfactant/TiO2 dispersions under visible light
Advanced Oxidation Processes for Wastewater Treatment 325

irradiation: Evidence for the need of substrate adsorption on TiO2 particles.


Environ. Sci. Technol., 32, 2394–2400.
Zhao, L., Ma, J., Sun, Z.Z., and Liu, H.L. (2009a). Mechanism of heterogeneous
catalytic ozonation of nitrobenzene in aqueous solution with modified ceramic
honeycomb. Appl. Catal. B: Environ., 89, 326–334.
Zhao, L., Ma, J., Sun, Z.Z., and Zhai, X.D. (2009b). Preliminary kinetic study on the
degradation of nitrobenzene by modified ceramic honeycomb-catalytic ozona-
tion in aqueous solution. J. Hazard. Mater., 161, 988–994.
Zhao, L., Sun, Z.Z., and Ma, J. (2009c). Novel relationship between hydroxyl radical
initiation and surface group of ceramic honeycomb supported metals for the
catalytic ozonation of nitrobenzene in aqueous solution. Environ. Sci. Technol.,
43, 4157–4163.
Zhao, L., Sun, Z.Z., Ma, J., and Liu, H.L. (2009d). Enhancement mechanism of hetero-
geneous catalytic ozonation by cordierite-supported copper for the degradation
Downloaded by [Tsinghua University] at 19:33 24 September 2012

of nitrobenzene in aqueous solution. Environ. Sci. Technol., 43, 2047–2053.


Zhao, W.R., Wu, Z.B., Shi, H.X., and Wang, D.H. (2005b). UV photodegradation of
azo dye Diacryl red X-GRL. J. Photochem. Photobiol. A: Chem., 171, 97–106.
Zhu, X.P., Tong, M.P., Shi, S.Y., Zhao, H.Z., and Ni, J.R. (2008). Essential explanation
of the strong mineralization performance of boron-doped diamond electrodes.
Environ. Sci. Technol., 42, 4914–4920.

View publication stats

You might also like