You are on page 1of 14

Applied Catalysis B: Environmental 49 (2004) 114

TiO2 -assisted photocatalytic degradation of azo dyes in


aqueous solution: kinetic and mechanistic investigations
A review
Ioannis K. Konstantinou , Triantafyllos A. Albanis
Department of Chemistry, Laboratory of Environmental Technology, University of Ioannina, Ioannina 45110, Greece
Received 5 July 2003; received in revised form 24 November 2003; accepted 24 November 2003

Abstract
The photocatalytic degradation of azo dyes containing different functionalities has been reviewed using TiO2 as photocatalyst in aqueous
solution under solar and UV irradiation. The mechanism of the photodegradation depends on the radiation used. Charge injection mechanism
takes place under visible radiation whereas charge separation occurred under UV light radiation. The process is monitored by following either
the decolorization rate and the formation of its end-products. Kinetic analyses indicate that the photodegradation rates of azo dyes can usually
be approximated as pseudo-first-order kinetics for both degradation mechanisms, according to the LangmuirHinshelwood model. The degradation of dyes depend on several parameters such as pH, catalyst concentration, substrate concentration and the presence of electron acceptors
such as hydrogen peroxide and ammonium persulphate besides molecular oxygen. The presence of other substances such as inorganic ions,
humic acids and solvents commonly found in textile effluents is also discussed. The photocatalyzed degradation of pesticides does not occur
instantaneously to form carbon dioxide, but through the formation of long-lived intermediate species. Thus, the study focuses also on the
determination of the nature of the principal organic intermediates and the evolution of the mineralization as well as on the degradation pathways followed during the process. Major identified intermediates are hydroxylated derivatives, aromatic amines, naphthoquinone, phenolic
compounds and several organic acids. By-products evaluation and toxicity measurements are the key-actions in order to assess the overall
process.
2003 Elsevier B.V. All rights reserved.
Keywords: Azo dyes; Photocatalytic degradation processes; Operational parameters; Transformation products

1. Introduction
Textile dyes and other industrial dyestuffs constitute one
of the largest group of organic compounds that represent an
increasing environmental danger. About 120% of the total
world production of dyes is lost during the dyeing process
and is released in the textile effluents [14]. The release of
those colored waste waters in the environment is a considerable source of non aesthetic pollution and eutrophication
and can originate dangerous byproducts through oxidation,
hydrolysis, or other chemical reactions taking place in the
wastewater phase [58].

Corresponding author. Tel.: +30-26510-98363;


fax: +30-26510-98795.
E-mail address: iokonst@cc.uoi.gr (I.K. Konstantinou).

0926-3373/$ see front matter 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcatb.2003.11.010

Decolorization of dye effluents has therefore received increasing attention. For the removal of dye pollutants, traditional physical techniques (adsorption on activated carbon,
ultrafiltration, reverse osmosis, coagulation by chemical
agents, ion exchange on synthetic adsorbent resins, etc.) can
generally be used efficiently [912]. Nevertheless, they are
non-destructive, since they just transfer organic compounds
from water to another phase, thus causing secondary pollution. Consequently, regeneration of the adsorbent materials
and post-treatment of solid-wastes, which are expensive
operations, are needed [12,13]. Due to the large degree
of aromatics present in dye molecules and the stability of
modern dyes, conventional biological treatment methods
are ineffective for decolorization and degradation [1417].
Furthermore, the majority of dyes is only adsorbed on the
sludge and is not degraded [18]. Chlorination and ozonation are also being used for the removal of certain dyes but

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

at slower rates as they have often high operating costs and


limited effect on carbon content [12,1922].
These are the reasons why advanced oxidation processes
(AOPs) have been growing during the last decade since they
are able to deal with the problem of dye destruction in aqueous systems. AOPs were based on the generation of very
reactive species such as hydroxy radicals ( OH) that oxidize a broad range of pollutants quickly and non selectively.
AOPs such as Fenton and photo-Fenton catalytic reactions
[2327], H2 O2 /UV processes [28,29] and TiO2 mediated
photo-catalysis [11,3033] have been studied under a broad
range of experimental conditions in order to reduce the color
and organic load of dye containing effluent waste waters.
Among AOPs, heterogeneous photocatalysis using TiO2
as photo-catalyst appears as the most emerging destructive
technology [4,3439]. The key advantage of the former is its
inherent destructive nature: it does not involve mass transfer; it can be carried out under ambient conditions (atmospheric oxygen is used as oxidant) and may lead to complete
mineralization of organic carbon into CO2 . Moreover, TiO2
photocatalyst is largely available, inexpensive, non-toxic and
show relatively high chemical stability. Finally, TiO2 photocatalytic process is receiving increasing attention because
of its low cost when using sunlight as the source of irradiation. The utilization of combined photocatalysis and solar
technologies may be developed to a useful process for the
reduction of water pollution by dying compounds because
of the mild conditions required and their efficiency in the
mineralization [7,4044].
The application of photocatalytic procedures for remediation of textile wastewater is rather limited to few investigations [4549]. There are many studies dealing with the
photocatalytic decolorization of specific textile dyes from
different chemical categories, and most of them including
a detailed examination of the so-called primary process under different working conditions [7,11,32,41,42,5057]. On
the contrary, little information is available on the reaction
mechanisms involved in the photocatalytic degradation of
dyes and on the identification of major transient intermediates which have been more recently recognized as very
important aspects of these processes, especially in view of
their practical applications [6,34,35,44,58]. Thus, information about real mineralization of the dye or decreases in
toxicity are scarce and therefore our attention has been also
focused on the reaction types and mechanisms, based on
the identification of the transformation products. Moreover,
the effect of common dyebath constituents on the photocatalytic treatment efficiency is also discussed in order to examine the application of the photocatalytic degradation on
real wastewater effluents.
Of the dyes available on the market today, approximately
5070% are azo compounds followed by the anthraquinone
group. Azo dyes can be divided into monoazo, diazo, triazo
classes according to the presence of one or more azo bonds
(N=N) and are found in various categories, i.e. acid, basic, direct, disperse, azoic and pigments [3,59]. Some azo

dyes and their dye precursors have been shown to be or are


suspected to be human carcinogens as they form toxic aromatic amines [44,60,61]. Therefore azo dyes are pollutants
of high environmental impact and were selected as the most
relevant group of dyes concerning their degradation using
TiO2 assisted photocatalysis.
To our knowledge there is not a review dealing with the
photocatalytic degradation of dyes although that there are
some reviews concerning the photocatalytic degradation of
other pollutants such as pesticides [37,6264]. This review
intend to assist workers involved in azo dyes photocatalytic
treatment using TiO2 by: (a) compiling data on the degree
and on the factors influencing dye photodegradation, and
(b) Summarizing and discussing data on the mineralization
degree, the intermediates and reaction mechanisms followed
during the process. The azo dyes were classified in terms of
the characteristic structural groups.
2. Experimental procedures
2.1. Photocatalytic degradation mechanisms
2.1.1. Photocatalytic oxidation
The detailed mechanism of the process has been discussed previously in the literature [4,6,58,6568] and will
be only briefly summarized here. It is well established that
conduction band electrons (e ) and valence band holes (h+ )
are generated when aqueous TiO2 suspension is irradiated
with light energy greater than its band gap energy (Eg ,
3.2 eV). The photogenerated electrons could reduce the dye
or react with electron acceptors such as O2 adsorbed on the
Ti(III)-surface or dissolved in water, reducing it to superoxide radical anion O2 . The photogenerated holes can oxidize the organic molecule to form R+ , or react with OH or
H2 O oxidizing them into OH radicals. Together with other
highly oxidant species (peroxide radicals) they are reported
to be responsible for the heterogeneous TiO2 photodecomposition of organic substrates as dyes. According to this,
the relevant reactions at the semiconductor surface causing
the degradation of dyes can be expressed as follows:
TiO2 + hv(UV) TiO2 (eCB + hVB + )
+

TiO2 (hVB ) + H2 O TiO2 + H + OH


+

TiO2 (hVB ) + OH TiO2

+ OH

(1)
(2)
(3)

TiO2 (eCB ) + O2 TiO2 + O2

(4)

O2 + H+ HO2

(5)

Dye + OH degradation products

(6)

Dye + hVB oxidation products

(7)

Dye + eCB reduction products

(8)

The resulting OH radical, being a very strong oxidizing


agent (standard redox potential +2.8 V) can oxidize most of

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

azo dyes to the mineral end-products. Substrates not reactive toward hydroxyl radicals are degraded employing TiO2
photocatalysis with rates of decay highly influenced by the
semiconductor valence band edge position [69]. The role of
reductive pathways (Eq. (8)) in heterogeneous photocatalysis has been envisaged also in the degradation of several
dyes but in a minor extent than oxidation [58,70].
2.1.2. Photosensitized oxidation
The mechanism of photosensitized oxidation (called
also photo-assisted degradation) by visible radiation ( >
420 nm) is different from the pathway implicated under
UV light radiation. In the former case the mechanism suggests that excitation of the adsorbed dye takes place by
visible light to appropriate singlet or triplet states, subsequently followed by electron injection from the excited dye
molecule onto the conduction band of the TiO2 particles,
whereas the dye is converted to the cationic dye radicals
(Dye+ ) that undergoes degradation to yield products as
follows [32,34,50,51,67,68,7072]:
Dye + hv(VIS) 1 Dye or 3 Dye
1

Dye or 3 Dye + TiO2 Dye+ + TiO2 (eCB )

(9)
(10)

TiO2 (eCB ) + O2 O2 + TiO2

(11)

Dye+ degradation products

(12)

The cationic dye radicals readily reacts with hydroxyl ions


undergoing oxidation via reactions 13 and 14 or interacts
effectively with O2 , HO2 or HO species to generate
intermediates that ultimately lead to CO2 (Eqs. (15)(19)).
Dye+ + OH Dye + HO

(13)

Dye + 2HO H2 O + oxidation products

(14)

O2 + H HO2

(15)

HO2 + H+ +TiO2 (eCB ) H2 O2 + TiO2

(16)

H2 O2 + TiO2 (eCB ) HO + HO + TiO2

(17)

Dye+ +O2 DO2 degradation products

(18)

Dye+ +HO2 (or HO ) degradation products

(19)

In experiments that are carried out using sunlight or simulated sunlight (laboratory experiments) it is suggested that
both photooxidation or photosensitizing mechanism occurred during the irradiation and both TiO2 and the light
source are necessary for the reaction to occur. In the photocatalytic oxidation, TiO2 has to be irradiated and excited
in a near-UV energy to induce charge separation. On the
other hand, dyes rather TiO2 are excited by visible light
followed by electron injection onto TiO2 conduction band,
which leads to photosensitized oxidation. It is difficult to
conclude whether the photocatalytic oxidation is superior
to the photosensitizing oxidation mechanism, but the photosensitizing mechanism will help to improve the overall

efficiency and make the photobleaching of dyes using solar


light more feasible [51].
2.2. Primary substrate disappearance
Several experimental results indicated that the destruction rates of photocatalytic oxidation of various dyes over
illuminated TiO2 fitted the LangmuirHinshelwood (LH)
kinetics model [4,9,65,7375]:
r=

dC
kKC
=
dt
1 + KC

(20)

where r is the oxidation rate of the reactant (mg/l min), C


the concentration of the reactant (mg/l), t the illumination
time, k the reaction rate constant (mg/l min), and K is the
adsorption coefficient of the reactant (l/mg).
When the chemical concentration Co is a millimolar solution (Co small) the equation can be simplified to an apparent
first-order equation [4,9,37]:
 
Co
k
Ln
t
(21)
= kKt = kapp. t or Ct = Co eapp.
C
A plot of ln Co /C versus time represents a straight line, the
slope of which upon linear regression equals the apparent
first-order rate constant kapp .
Generally first-order kinetics are appropriate for the
entire concentration range up to few ppm and several
studies were reasonably well fitted by this kinetic model
[8,38,43,44,51,54,55,58,60,76]. The LH model was established to describe the dependence of the observed reaction
rate on the initial solute concentrations.
It has been agreed, with minor doubts that the expression for the rate of photomineralization of organic
substrates such dyes with irradiated TiO2 follows the
LangmuirHinshelwood (LH) law for the four possible situations; (a) the reaction takes place between two adsorbed
substances, (b) the reaction occurs between a radical in solution and an adsorbed substrate molecule, (c) the reaction
takes place between a radical linked to the surface and a
substrate molecule in solution, and (d) the reaction occurs
with the both of species being in solution. In all cases, the
expression for the rate equation is similar to that derived
from the LH model, which has been useful in modeling
the process, although it is not possible to find out whether
the process takes place on the surface in the solution or at
the interface [6].
It is likely that sorption of the dye is an important parameter in determining photocatalytic degradation rates. All
isotherms showed L-shape curves according to the classification of Giles et al. [77] that means there is no strong competition between the water and the dye molecules to occupy
the TiO2 surface sites. The adsorption isotherms fit well to
Langmuirian type implying a monolayer adsorption model
[4,73,75,76].
The color removal of the dye solution was determined
usually with the absorbance value at the maximum of the

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

absorption spectrum for every dye by monitoring UV-Vis


spectrum in 200800 nm zone using a spectrophotometer
[41,51,72,78]. Alternatively, the disappearance of dye was
monitored by high performance liquid chromatography
equipped with a UV diode array detector [42,58,79,80] or
MS detector [3].
2.3. Factors influencing the photocatalytic degradation
2.3.1. Effect of initial dye concentration
It is important both from a mechanistic and from an application point of view to study the dependence of the photocatalytic reaction rate on the substrate concentration. It is
generally noted that the degradation rate increases with the
increase in dye concentration to a certain level and a further
increase in dye concentration leads to decrease the degradation rate of the dye [8,42]. The rate of degradation relates
to the probability of OH radicals formation on the catalyst surface and to the probability of OH radicals reacting
with dye molecules. As the initial concentrations of the dye
increase the probability of reaction between dye molecules
and oxidizing species also increases, leading to an enhancement in the decolorization rate. On the contrary, the degradation efficiency of the dye decreases as the dye concentration
increases further. The presumed reason is that at high dye
concentrations the generation of OH radicals on the surface of catalyst is reduced since the active sites are covered
by dye ions. Another possible cause for such results is the
UV-screening effect of the dye itself. At a high dye concentration, a significant amount of UV may be absorbed by the
dye molecules rather than the TiO2 particles and that reduces
the efficiency of the catalytic reaction because the concentrations of OH and O2 decrease [9,52,53,68,74,81,82].
The major portion of degradation occurs in the region near
to the irradiated side (termed as reaction zone) where the irradiation intensity is much higher than in the other side [83].
Thus at higher dye concentration, degradation decreases at
sufficiently long distances from the light source or the reaction zone due to the retardation in the penetration of light.
Hence, it is concluded that as initial concentration of the dye
increases, the requirement of catalyst surface needed for the
degradation also increases [7].
2.3.2. Effect of TiO2 loading
Whether in static, slurry or dynamic flow reactors the initial reaction rates were found to be directly proportional to
catalyst concentration indicating the heterogeneous regime.
However, it was observed that above a certain level of
concentration the reaction rate even decreases and becomes
independent of the catalyst concentration. Most of studies
reported enhanced degradation rates for catalyst loading
up to 400500 mg/l [8,42,52,53,55,84,85]. Only a slight
enhancement or decrease was observed when TiO2 concentration further increased up to 2000 mg/l. This can be rationalized in terms of availability of active sites on TiO2 surface
and the light penetration of photoactivating light into the

suspension. The availability of active sites increases with the


suspension of catalyst loading, but the light penetration, and
hence, the photoactivated volume of the suspension shrinks.
Moreover, the decrease in the percentage of degradation at
higher catalyst loading may be due to deactivation of activated molecules by collision with ground state molecules
[7]. Agglomeration and sedimentation of the TiO2 particles were observed elsewhere when 2000 mg/l of TiO2 was
added to the dye solution [52]. In such a condition, part of
the catalyst surface probably became unavailable for photon
absorption and dye adsorption, thus bringing little stimulation to the catalytic reaction. On the contrary, continuous
increase of the photocatalytic degradation rate of Reactive
Black 5 was found up to 3500 mg/l TiO2 [74]. The crucial
concentration depends on the geometry, the working conditions of the photoreactor and the type of UV-lamp (power,
wavelength). The optimum amount of TiO2 has to be added
in order to avoid unnecessary excess catalyst and also to ensure total absorption of light photons for efficient photomineralization. This optimum loading of photocatalyst is found
to be dependent on the initial solute concentration [62].
2.3.3. Effect of pH
The interpretation of pH effects on the efficiency of dye
photodegradation process is a very difficult task because of
its multiple roles. First, is related to the ionization state of
the surface according to the following reactions,
TiOH + H+ TiOH2 +

(22)

TiOH + OH TiO + H2 O

(23)

as well as to that of reactant dyes and products such as acids


and amines. pH changes can thus influence the adsorption of
dye molecules onto the TiO2 surfaces, an important step for
the photocatalytic oxidation to take place [86]. Bahnemann
et al. [87] have already reviewed that acid-base properties
of the metal oxide surfaces can have considerable implications upon their photocatalytic activity. The point of zero
charge (pzc) of the TiO2 (Degussa P25) is at pH 6.8 [88].
Thus, the TiO2 surface is positively charged in acidic media
(pH < 6.8), whereas it is negatively charged under alkaline
conditions (pH > 6.8).
Second, hydroxyl radicals can be formed by the reaction
between hydroxide ions and positive holes. The positive
holes are considered as the major oxidation species at low
pH whereas hydroxyl radicals are considered as the predominant species at neutral or high pH levels [78,89]. It was
stated that in alkaline solution OH are easier to be generated by oxidizing more hydroxide ions available on TiO2
surface, thus the efficiency of the process is logically enhanced [54,55,65,90,91]. Similar results are reported in the
photocatalysed degradation of acidic azo dyes and triazine
containing azo dyes [7,9,52,74,92,93], although it should
be noted that in alkaline solution there is a Coulombic
repulsion between the negative charged surface of photocatalyst and the hydroxide anions. This fact could prevent the

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

formation of OH and thus decrease the photoxidation. Very


high pHs have been found favorable even when anionic azo
dyes should hamper adsorption on the negatively charged
surface [81]. At low pH, reduction by electrons in conduction band may play a very important role in the degradation
of dyes due to the reductive cleavage of azo bonds.
Third, the TiO2 particles tend to agglomerate under acidic
condition and the surface area available for dye adsorption
and photon absorption would be reduced [86]. Hence, pH
plays an important role both in the characteristics of textile
waters and in the reaction mechanisms that can contribute
to dye degradation, namely, hydroxyl radical attack, direct
oxidation by the positive hole and direct reduction by the
electron in the conducting band.
The degradation rate of some azo dyes increases with decrease in pH as reported elsewhere [42,67,94,95]. At pH <
6, a strong adsorption of the dye on the TiO2 particles is
observed as a result of the electrostatic attraction of the positively charged TiO2 with the dye. At pH > 6.8 as dye
molecules are negatively charged in alkaline media, their
adsorption is also expected to be affected by an increase in
the density of TiO groups on the semiconductor surface.
Thus, due to Coulombic repulsion the dyes are scarcely adsorbed [44,65,76]. For the above reasons the photocatalytic
activity of anionic dyes (mainly sulphonated dyes) reached a
maximum in acidic conditions followed by a decrease in the
pH range 711 [42,53,58,68,75,76]. Moreover, the higher
degradation rate at acid pH is seen also for Vis/TiO2 experiments due to the efficient electron-transfer process due to
strong surface complex bond formation. This effect is less
marked in neutral/basic pH solutions [67].
On the contrary, different optimal pHs (67) have been observed for the photocatalytic degradation of other azo dyes,
and a decrease of degradation in both acidic and alkaline
pH was reported [82,96]. The inhibitory effect seems to be
more pronounced in the alkaline range (pH = 1113). At
high pH values the hydroxyl radicals are rapidly scavenged
and they do not have the opportunity to react with dyes [97].
An additional explanation for the pH effects can be related
with changes in the specification of the dye. That is, protonation or deprotonation of the dye can change its adsorption
characteristics and redox activity [7].
Since the influence of the pH is dependent on dye type
and on properties of TiO2 surface his effect on the photocatalytic efficiency must be accurately checked before any
application.
2.3.4. Effect of light intensity and irradiation time
Ollis et al. [98] reviewed the studies reported for the effect
of light intensity on the kinetics of the photocatalysis process
and stated that (i) at low light intensities (020 mW/cm2 ),
the rate would increase linearly with increasing light intensity (first order), (ii) at intermediate light intensities beyond a certain value (approximately 25 mW/cm2 ) [62], the
rate would depend on the square root of the light intensity
(half order), and (iii) at high light intensities the rate is in-

dependent of light intensity. This is likely because at low


light intensity reactions involving electronhole formation
are predominant and electronhole recombination is negligible. However, at increased light intensity electronhole
pair separation competes with recombination, thereby causing lower effect on the reaction rate. In the studies reviewed
here, the enhancement of the rate of decolorization as the
light intensity increased was also observed [7,42,52,74,75].
It is evident that the percentage of decolorization and photodegradation increases with increase in irradiation time. The
reaction rate decreases with irradiation time since it follows
apparent first-order kinetics and additionally a competition
for degradation may occur between the reactant and the intermediate products. The slow kinetics of dyes degradation
after certain time limit is due to: (a) the difficulty in converting the N-atoms of dye into oxidized nitrogen compounds
[99], (b) the slow reaction of short chain aliphatics with OH
radicals [100], and (c) the short life-time of photocatalyst
because of active sites deactivation by strong by-products
deposition (carbon etc.).
2.3.5. Effect of oxidants
It was observed that H2 O2 and S2 O8 2 addition was beneficial for the photoxidation of the dyes of different chemical
groups included azo dyes [6,8,43,52]. The reactive radical
intermediates ( SO4 and OH) formed from these oxidants
by reactions with the photogenerated electrons can exert a
dual function: as strong oxidant themselves and as electron
scavengers, thus inhibiting the electronhole recombination
at the semiconductor surface [101] according to the following equations:
H2 O2 + O2 OH + OH + O2

(24)

H2 O2 + hv 2 OH

(25)

H2 O2 + eCB OH + OH

(26)

S2 O8 2 + eCB SO4 2 + SO4

(27)

SO4 + H2 O SO4 2 + OH + H+

(28)

Moreover, the solution phase may at times be oxygen


starved, because of either oxygen consumption or slow oxygen mass transfer. Peroxide addition thereby increases the
rate towards what it would have been an adequate oxygen
supply. The presence of S2 O8 2 positively influences the
mineralization rate, despite the decreasing of pH as the oxidant properties of the system probably prevail on the effect
of pH reduction. On the contrary, as far as the substrate is
concerned, the faster degradation rate can be due to both
the decrease of the pH and the oxidant action of S2 O8 2
[43].
However, H2 O2 can also become a scavenger of valence
band holes and OH, when present at high concentration,
[68,102104]:
H2 O2 + 2hVB + O2 + 2H+

(29)

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

H2 O2 + OH H2 O + HO2

(30)

HCO3 + OH H2 O + CO3

(35)

HO2 + OH H2 O + O2

(31)

SO4 2 + h+ SO4

(36)

SO4 2 + OH SO4 + OH

(37)

As both hVB + and OH are strong oxidants for dyes, the


photocatalytic oxidation will be inhibited when H2 O2 level
gets too high. Furthermore, H2 O2 can be adsorbed onto TiO2
particles to modify their surfaces and subsequently decrease
its catalytic activity.
Since the influence of the above additives, in particular
H2 O2 , has been in some cases controversial and it appeared
dependent on the substrate type and on various experimental
parameters [105] the usefulness of which must be accurately
checked before their application [43].
2.3.6. Effect of humic acids, natural occurring ions and
solvents
The occurrence of dissolved inorganic ions is rather
common in dye-containing industrial wastewater. Often,
wastewater contains a mixture of pollutants, organic solvents
as well as dissolved organic matter and humic substances,
if mixed with other waste streams. These substances may
compete for the active sites on the TiO2 surface or deactivate the photocatalyst and, subsequently, decrease the
degradation rate of the target dyes. Alternatively, they may
act as light screens, thus reducing the photon receiving
efficiency.
The Vis/TiO2 photocatalytic degradation of different
classes of dyes is reported to be retarded by many commonly used industrial solvents and acids, as well as by
many naturally abundant mineral species and dissolved organic matter [99]. The retardation by humic substances may
be by the combination effects of light attenuation, competition for active sites and surface deactivation [106108].
Finally, various solvents such as acetonitrile and ethanol
were found to have a significant retardation effect on the
photobleaching of dyes even at low concentrations [68,106]
as it is also stated for phenols and aromatic products [109].
Of the anionic species studied (HCl, NaCl, NaNO3 , HNO3 ,
H3 PO4 and NaHCO3 ), HCl exhibited the strongest inhibition effect followed by H3 PO4 [68,106]. Inhibition effects
of anions can be explained as the reaction of positive
holes and hydroxyl radical with anions, that behaved as h+
and OH scavengers (Eqs. (32)(37)) resulting prolonged
color removal. Probably the adsorbed anions compete
with dye for the photo-oxidizing species on the surface
and preventing the photocatalytic degradation of the dyes
[87,93,110]. Formation of inorganic radical anions (e.g.
Cl , NO3 ) under these circumstances is possible to occur
[111].
Cl + hVB + Cl or Cl + OH ClOH

(32)

NO3 + hVB + NO3 or


NO3 + OH NO3 + OH

(33)

CO3 2 + OH OH + CO3

(34)

Although the reactivity of these radicals may be considered,


they are not as reactive as h+ and OH [112] and thus, the
observed retardation effect is still thought to be the strong
adsorption of the anions on the TiO2 surface [110].
The effect of several types of metal ions (Cu2+ , Zn2+ ,
3+
Fe , Al3+ and Cd2+ ) on the photodegradation of non-azo
dyes in TiO2 aqueous dispersions under visible light illumination, has been investigated by Chen et al. [113]. They
have concluded that Cu2+ and Fe3+ ions have a strong
suppressing effect on the photodegradation of all three
dyes examined, by altering the interfacial electron-transfer
pathway under visible light irradiation. They noted that the
addition of Cu2+ and Fe3+ decreases the reduction of O2 by
the conduction electrons, subsequently blocks the formation
of reactive oxygen species (O2 / OOH, OH) and hence
suppresses the photodegradation of dyes under visible irradiation. However, other metal ions such as Zn2+ , Cd2+
and Al3+ affect the photoreaction only slightly through an
alteration of the adsorption of dyes.
On the basis of hydroxyl radical formation through photocatalytic reactions of Fe3+ ions and the products of their
hydrolysis in aqueous solutions [114] is assumed that, the
presence of Fe3+ in the reaction environment, together with
TiO2 , should increase the rate of the photocatalytic processes. An increased degradation rate was observed in the
photocatalytic degradation of azo dye acid red 1 in TiO2
suspensions containing Fe(III) aquo ions (105 to 104 M)
[115]. This beneficial behavior was attributed to the increased amount of dye adsorbed on the iron(III)-modified
TiO2 surface and this was further confirmed by the fact
that iron species such as Fe2+ not adsorbed on the semiconductor had no kinetic effects. The beneficial effect of
Fe3+ ions was also found on the photocatalytic degradation of rhodamine B in aqueous TiO2 suspensions
[33].
Baran et al. [116] studied the photocatalytic degradation
of several anionic and cationic azo dyes in the presence
of TiO2 and FeCl3 . They have found that Fe3+ ions have
a catalytic influence on the decolorization of the studied
anionic dyes but an inhibiting influence on the decolorization
of the cationic dyes. In conclusion, the role of Fe3+ ions
on the photocatalytic degradation of several dyes shows a
controversial behavior depending on the physico-chemical
properties of dyes. Thus, in the presence of these ions the
specific azo dye degradation should be considered in order
to determine the treatment efficiency.
The photocatalytic decolorization of the triazine azo dye
MX-5B was reported to increase slightly in the presence of
1 M of Cu2+ and Ni2+ at pH = 2.4 [112]. Their reduced
forms could trap holes and that explains the decrease of the
e /h+ recombination rate and a higher production of OH.

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

Excess of Cu2+ and Ni2+ led to short-circuiting reactions,


which created a cyclic process without generating active
OH and retarded the reaction. However, at pH 10.8 the
photodegradation of MX-5B was completely inhibited by
the trace quantities of Cu2+ and Ni2+ . The deposition of
NiO2 on the surface of TiO2 was found to deactivate the
photocatalyst [112].
An understanding of the retardation effects not only aids in
assessing the feasibility of using photocatalytic oxidations to
treat wastewater, but also allows a thoughtful photocatalytic
oxidation design.
2.4. Photocatalytic mineralization of dyes
2.4.1. Analysis of the end products
In order to assess the degree of mineralization reached
during the photocatalytic treatment the formation of CO2
and inorganic ions [6,32,34,44,58,91,117], is generally determined. However, in the presence of real wastewaters the
monitoring of inorganic ions and CO2 gives only a global
estimation on the well functioning of the treatment, but does
not provide information on the real decay of the contaminant. In such cases the determination of total organic carbon
(TOC) and/or the measurement of the chemical oxygen
demand (COD) or the biological oxygen demand (BOD)
of the irradiated solution is generally used for monitoring
the mineralization of the dye [7,32,42,52,58,76,80,118]. In
general, at low reactant levels or for compounds which do
not form important intermediates, complete mineralization
and reactant disappearance proceed with similar half lives,
but at higher reactant levels where important intermediates occur, mineralization is slower than the degradation
of the parent compound. Until now, total mineralization
has been observed for the photacatalytic degradation of
most of the azo dyes even at longer irradiation periods
[42,44,58,74,76,119]. Only in the case of triazine containing dyes, the mineralization was not complete due to the
high stability of triazine nucleus and the stable cyanuric
acid was formed, as in the case of s-triazine herbicides
[120], which fortunately is not toxic [41,52,80,121].
Usually COD or TOC values decrease with increase in
irradiation time whereas the amount of NH4 + and NO3
ions increase with increase in irradiation time. However, the
formation of Cl and SO4 2 increases initially and subsequently remains unchanged. COD or TOC curves have an
exponential or sigmoidal shape. The sigma-shaped curves
indicating that is related to the formation of relative tolerant
by-products [44,52,118]. This pattern means that during the
first steps of the process where the solution is still colored
there is only a small decrease of the parameter measured
(TOC or COD or BOD) due to the fact that dye molecules
are decomposed to lower molecular weight compounds and
the resulting intermediates still contribute to the COD of
the solution. After the decolorization of the solution the
COD decreases sharply (the linear segment of the S shaped
curve) reaching a plateau that corresponds to the oxidation

of most stable compounds indicating that almost complete


mineralization of intermediates has occurred.
For chlorinated dye molecules, Cl ions are easily released in the solution and are the first of the ions appearing during the photocatalytic degradation [42,43,80]. This
could be interesting in a process, where photocatalysis would
be associated with a biological treatment which is generally not efficient for chlorinated compounds. Nitrogen is
mineralized into NH4 + , NO3 and N2 . The proportion depends mainly on the initial oxidation degree of nitrogen, the
substrate structure and on irradiation time [122124]. By
comparing the initial rates, NH4 + appears as the primary
product with respect to NO3 in the case of amine compounds. The nitrogen atoms in the amino-groups of the dyes
can lead to NH4 + ions by successive attacks by hydrogen
species
R-NH2 + H R + NH3

(38)

NH3 + H+ NH4 +

(39)

The total amount of nitrogen-containing ions present in


the solution at the end of the experiments is usually lower
than that expected from stoichiometry indicating that
N-containing species remain adsorbed in the photocatalyst
surface or most probably, that significant quantities of N2
and/or NH3 have been produced and transferred to the
gas-phase. [42,44,76]. The formation of N2 in azo dyes
can be accounted for by the same processes responsible for
NH4 + formation:
R-N = N-R + H R-N = N + R H

(40)

R-N = N R + N N

(41)

When nitrogen is present in the 3 state as in amino groups


or in pyrazoline ring, it spontaneously evolves as NH4 +
cations with the same oxidation degree, before being subsequently and slowly oxidized into nitrate [58]. In the azo
bonds each nitrogen atom is in its +1 oxidation degree. This
oxidation degree favors the evolution of gaseous dinitrogen
by the two step reduction process expressed previously. N2
evolution constitutes the ideal case for a decontamination reaction involving totally innocuous nitrogen-containing final
product.
The dyes containing sulfur atoms are mineralized into sulfate ions [43,80]. In all the studies the formation of SO4 2
was always observed and in most cases its stoichiometric
formation was found in the final steps of the photoreaction
when organic intermediates still were present [43,52,80].
The reported initial slopes are positive indicating that SO4 2
ions are initial products, directly resulting from the initial
attack on the sulfonyl group. Release of sulphate ions upon
dye degradation was a little slower than decolorization but
much faster than TOC loss. Non-stoichiometric formation
of sulphate ions is usually explained by a strong adsorption
on the photocatalyst surface [44,125,126]. This strong adsorption could partially inhibit the reaction rate which, however, remains acceptable [111,127]. Generally, it is found

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

that nitrate anions have little effect on the kinetics of reaction whereas sulfate, chloride and phosphate ions, especially
at concentrations of greater than 103 mol dm3 , can reduce
the rate by 2070% due to the competitive adsorption at the
photoactivated reaction sites [111]. The release of SO4 2
can be accounted by an initial attack by a photo-induced
OH radical:
OH + p+ OH

(42)

R-SO3 + OH R-OH + SO3

(43)

SO3 + OH SO4 2 + H

(44)

The attack of sulfonate groups would be favored if the


molecule is adsorbed with its SO3 group orientated to the
surface [76]. The hydrogen atom generated can react with
other radicals or with a neutral functional group such as an
amino group.
2.4.2. Nature and evolution of organic intermediates
Cost effective treatment to complete pesticide mineralization is usually not feasible and by-products generation
appears to be unavoidable with photocatalytic degradation process. Kinetics of formation and decomposition of
the intermediates are needed and identification of these
byproducts needs to be established in order to (1) determine which specific compounds will appear in the effluent
(2) increase our knowledge on the degradation pathways in
order to reveal which step is crucial for the global reaction
of the process. Identification of by-products is one of the
keys to maximizing the overall process efficiency. Since
hydroxyl radicals react non-selectively, various by-products
are formed at low concentration levels. Various analytical
techniques such as high performance liquid chromatography
(HPLC) [65,67,128], gas chromatographymass spectrometry (GCMS) [34,44,65,68], liquid chromatographymass
spectrometry (LCMS) [4,6,128], 1 H NMR [34,65,72], diffuse reflectance FT-IR [44,72,79,80,129] and electron spin
resonance (ESR) [33,34] were used for the determination
of organic intermediates.
The azo dyes that have been already studied for their
TiO2 -mediated photocatalytic degradation are summarized in Table 1. Generally, the sites near the azo bond
(CN=Nbond) is the attacked area in the photocatalytic
degradation process, whilst the TiO2 photocatalytic destruction of the CN= bond and NN bonds leads to fading
of the dyes [32]. Aromatic intermediates were identified
for most dyes. They are either aromatic amine or phenolic compounds. The formation of aminobenzenesulfonate
suggests the reductive cleavage of the azo group prior to
the opening of the aromatic ring [58,96]. The formation
of aromatic amines has also been reported in the natural
aerobic reduction of azo dye [130]. On the other hand,
the formation of phenolic compounds as intermediates is
commonly observed in the photocatalytic degradation of
other aromatic compounds [37,58]. Several organic acids
were found as aliphatic intermediates. Main products were

formic and acetic acids. Other organic acids detected were


oxalic, glycolic, glyoxylic and malonic acids. The formation
of these acids could correspond to the opening of aromatic
and naphthalene rings followed by a sequence of oxidation
steps which leads to progressively lower molecular weight
acids and the evolution of CO2 . Formation of CO2 takes
place via decarboxylation of carboxylic acids according to
the photo-Kolbe reaction [126]:
R-COO + h+ R + CO2

(45)

For most of the intermediate compounds the maximum


concentration of formic acid was larger than that of acetic
acid, as formic acid is more degradable by photocatalytic
processes than acetic acid [44,58]. Typical bell-shaped
curves were obtained in some studies for the aromatic and
naphthalene intermediates and the resulted acids [44,58] as
in the case of other pollutants as pesticides [37].
The possibility of generating molecular fragments during the photocatalyzed degradation that can be more toxic
than the parent compound [131] make also toxicity measurements obligatory part of the experiments. Moreover,
if only partial degradation is envisaged, toxicity assessment of treated water becomes necessary. Toxicity analyses
of the phototreated solutions of pesticides indicate the
toxicity of the sum of compounds formed and not only
those that have been identified. This parameter is very
important for the treatment operation. The toxicity of all
transient intermediates was compared to that observed for
the parent compound. A variety of toxicity measurement
systems exist, including those based on bacteria and algae, animal cells, swell mammals fish fly and zooplankton
[132,133]. Two of them, Microtox method and the inhibition of Escherichia coli bacterial respiration, were usually
applied in studies dealing with photocatalytic degradation
of dyes.
With Microtox method, the comparison was carried out
by monitoring the bioluminescence of the bacteria Vibrio fisheri as a function of illumination time [74,134]. It
has been used extensively to assess the toxicity of a wide
range of aquatic and terrestrial pollutants. This assay appears to be the best available method with high sensitivity
and reliability for field monitoring and toxicity screening of industrial effluents. Alternatively the inhibition in
bacterial respiration (e.g. Escherichia coli) was followed
[46,119,135]. This test has been reported to be well adapted
to toxicity screening of effluents, being both easy to use and
capable of providing reproducible and repeatable results
[132].
The resulted EC50 values after the decolorization of the
solution from the triazine dyes MX-5B and K-2G indicating
that the intermediates produced from photocatalytic degradation were not toxic [52,80]. On the contrary, the presence
of residual toxicity together with the permanence of about
20% of TOC content during the photocatalytic degradation
of Remazol Black B (Reactive Black 5) [119] and other azo
dyes [117] indicate that the molecular fragments produced

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

Table 1
Summary of TiO2 -mediated photocatalytic degradation studies of azo dyes
Azo dyes

Light irradiation

Transformation products

References

UV
Vis
Vis
Simulated solar
Solar

[50,55,58,70,79,136,139,140]
[71,137]
[50,116]
[44,67]
[43]

Acid orange 20

UV-Vis

[12]

Acid orange 52

UV
UV-Vis
Simulated solar
Solar

+
+

[55,76]
[65]
[128]
[43]

Tartazine,
Acid yellow 17
New coccine
Orange G
Basic orange 66
Acid red 27, acid red 33,
Allura red AC
Acid red 14
Acid red 1
Basic yellow 15
Reactive yellow 17
Cationic blue X-GRL
Basic blue 41
Acid yellow 23
Acid red 3B
Red acid G
Methyl red
30 H/K azo dyes

UV
UV
UV
UV/Solar
Vis
Vis
Vis
UV
Vis
UV
Solar/UV
UV-Vis
Vis
Vis
UV
UV-Vis
Solar/UV
UV

[58,139]
[58]
[58]
[76,93]
[116]
[51]
[51]
[68]
[115]
[9]
[7]
[112,118]
[116,142]
[116]
[54,56]
[50]
[76,93]
[66,141]

UV
UV

[9]
[56]

Reactive black 5

UV-Vis
UV

[74]
[85,119]

Reactive blue 221


Acid brown 14
Direct blue 1

UV
Solar
UV

[105]
[42]
[70]

Congo red

UV
Solar

[58,73]
[93]

Acid black 1

UV-Vis
UV

[53]
[58]

Naphthol blue black


Solophenyl green BLE,
Direct blue 160
Lanasol blue
Safira HEXL

Vis
UV
UV
Vis
UV

[129]
[60]
[9]
[57]
[75]

UV
UV
UV
UV
UV-Vis
Vis
Solar/UV

+/

[80]
[52,80,112]
[118]
[118]
[73]
[116]
[7]

Monoazo dyes
Acid orange 7

Di- and triazo dyes


Reactive red 120
Direct fast scarlet 4BS

Triazine-containing azo dyes


Reactive brilliant red K-2G
Procion red MX-5B
Reactive yellow KD-3G
Reactive red 15, reactive red 24
Reactive brilliant red X3B
Reactive red 45
Reactive blue 4

10

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

at higher photochemical reaction times are toxic. In the case


of anthraquinone dye Remazol brilliant Blue R the toxicity
is completely removed at the first reaction times, but almost
totally recovered with the progress of the photochemical
process [119].
2.4.2.1. Monoazo dyes. Acid orange 7 (AO 7) is the most
studied compound among the azo dyes as far as its photocatalytic degradation under several experimental conditions.
The degradation pathways and the formation of by-products
is also fully described [44,50,55,58,67,70,71,79,136140]
thus, AO 7 could be used as a model compound for oxidative degradation studies of azo dyes. Oxidative attack
of an azo dye from the phenyl azonaphthol family as AO
7 leads to benzene sulfonate and naphthoquinone as primary degradation products. Vinodgopal et al. [71] reported
the formation of four by-products (benzene sulphonic
acid, sulphoanilic acid, 1,4-naphthoquinone and phthalic
acid) and Bauer et al. [137] have identified in addition
quinone and 4-hydroxybenzene sulphonic acid during the
first steps of Vis/TiO2 photosensitized degradation of AO
7. The former products were also identified by Stylidi
et al. [44], which studied the complete degradation of AO
7 under solar light irradiation. Twenty-two transformation
products were identified in total, including 2-naphthol,
2-hydroxy-1,4-naphthoquinone, smaller aromatic intermediates such as pthalic acid and phtalimide and aliphatic
acids such as fumaric, succinic, maleic and malonic acids.
The lowest molecular weight compounds detected in that
study, are oxalic, acetic and formic acids. The photoxidation of AO 7 para isomer (acid orange 20) was also studied
under Vis/TiO2 [12,137]. Based on the previous identified
by-products the major degradation pathways for AO 7 are
shown in Fig. 1.
From the aminoazobenzene sub-category of monoazo
dyes, acid orange 52 (AO 52, namely also methyl orange)
is studied in details by various advanced oxidation processes UV/H2 O2 , UV/TiO2 , Vis/TiO2 , solar light/TiO2
[43,55,65,76,128]. Up to 18 intermediates were identified
including aniline, N,N-dimethyl aniline, hydroxy anilines,
hydroxy analogues of AO 52, phenols, quinone, benzene
sulphonic acid, demethylated analogues of AO 52, and various aliphatic and carboxylic acids. Spadaro et al. [117]
proposed that oxidation of aminoazobenzene dyes proceeds
by the addition of an hydroxy radical to the carbon atom
bearing the azo bond, followed by the breaking of the resulting adduct. The products such as benzenesulfonic acid,
N,N-dimethylaniline and 4-hydroxy-N,N-dimethylaniline
could arise from such reactions. The electron withdrawing
sulphonate group inhibits reactivity towards OH of the
ring that carried it, thus the ring with the amino group
is the first target for the hydroxy radicals [65]. The addition of OH on the carbon atom bearing the sulphonate
group and the subsequent elimination of SO3 is an non
probable pathway due to the electron withdrawing effect
of sulphonate group and steric hindrance. On the contrary,

HO
+Na O 3 S

HO 3 S

OH

H
NH 2

OH
OH
+ N2

HO 3 S

O
OH
O
HO 3 S

OH

2-

SO 4

C OOH
C OOH

O
O
O

Benzoic acid derivatives

CO 2 + H 2 0

Aliphatic
+
Carboxylic acids

Fig. 1. Major photocatalytic pathways of acid orange 7, a representative azo dye for phenyl-azonaphthol chemical group, based on the
identification of by-products from previous reported degradation studies
[12,44,67,71,137].

hydroxy-analogue derivatives of AO 52 are identified [65].


Similar reaction was observed also during the degradation of aminoazobenzene acid orange 5 (AO 5) and the
hydroxyazo dye AO 7 [138] but is presumed not to be
the major oxidation way. The N(CH3 )2 substituent group
is also an important site of attack [65,128] and thus the
demethylated analogues could be formed by such reactions. The full characterization of the resulted degradation
products is given in Baiocchi et al. [128]. According to the
previous identified by-products of AO 52 the major degradation pathways for aminoazobenzene dyes are shown in
Fig. 2.
Bezenesulfonic acid and phenols were identified also as
intermediates for the photooxidation of other monoazo dyes
such as Tartazine, Acid yellow 17, New Coccine and Orange G [58,76,139]. Zhang and Tian [66,141] studied the
effect of substituents and the spectral characterization of excited states on the TiO2 photocatalytic degradation of 30
H/K acid based azo dyes. They have concluded that the substituent dependent intramolecular resonance energy in the
conjugated molecules of o-arylazonaphthols determines the
light fastness of the dyes and that the triplet state of the dyes

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

e
ring hydroxylated
dye derivatives

.OH

.OH
R1

+Na O3S
c

N N
a1

ring hydroxylated
dye derivatives

a1

a2

ring hydroxylated
dye derivatives

R2
b

a2
R1
N
R2

.OH

a1

+ N2 +

HO3S

11

Dealkylated Dye
derivatives

.OH

R1
HO

N
R2

a2

H
HO3S

OH

N
R2

a1
NH2

.OH
HO

NH2

.OH
OH
2SO4 +

HO

OH

+
+ NH4 + NO3
Carboxylic
+
Aliphatic
Acids

CO2 + H2O

Fig. 2. Major photocatalytic degradation pathways for aminoazobenzene dyes based on the identification of by-products from previous reported degradation
studies of acid orange 52 [65,128].

does not react with OH and O2 . Finally, several studies


have been reported for the TiO2 photocatalytic degradation
of other monoazo dyes (Table 1) compiling data only on the
decolorization of the solution and the monitoring of the end
products (TOC or COD and inorganic ions) under various
operational conditions.
2.4.2.2. Di- and triazo dyes. Similar results were obtained
for some di- and triazo dyes. Phenol, and 4-nitro-2-hydroxy
phenol were identified as intermediate products for the
TiO2 photocatalytic degradation of Congo Red and Acid
Black 1, respectively [53,58,75,93]. 1,2 Naphthaquinone
were identified for Naphthol Blue Black azo dye [129].
Much work has been done on the operational conditions
and the parameters influencing the photocatalytic degradation of di and triazo dyes be several researchers (Table 1)
[9,42,56,57,70,74,75,82,119].
In some studies comparative photocatalytic degradation rates between azo dyes from different groups are
reported. Monoazo dyes are easier oxidized than diazo
dyes which may in turn be easier than triazo dyes as long
as auxochrome groups are similar [9,58]. This trend is
also observed in ozonation of azo dyes [143]. Dyes of
the -naphthol type degrade faster than dyes derived from

N,N-dimethylaniline and dyes containing the sulphonic acid


group degrade faster than those with the carboxylic group
[55].
2.4.2.3. Triazine containing azo dyes. Several studies reported the photocatalytic oxidation of triazine azo dyes in
various operational conditions [7,52,73,80,118] but only one
of them reported the photocatalytic pathways and the formation of transformation products [80]. The photocatalytic
degradation of triazine-containing azo dyes, Procion Red
MX-5B and Reactive Brilliant Red K-2G was investigated
in aqueous TiO2 dispersions [80]. The whole photocatalytic
degradation proceeds in three steps. In the first step, the more
active bonds were hydroxylated. These included the CN
bond linked to the benzene ring or the naphthalene ring and
the CS bond of sulfonate group linked to the naphthalene
ring or the benzene ring, to form organic acids with or without hydroxyl group and the related ions (SO4 2 and NH4 + ).
For the second step, the groups linked to the triazine ring
were replaced by hydroxyl group to yield the well-known
cyanuric acid, as in the case of s-triazines herbicides, and
the related ions (NO3 and Cl ). At the same time the aromatic acids produced from the first step were subsequently
hydroxylated and led to the cleavage of aromatic ring to form

12

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

aliphatic acids. The third step involved the further oxidation


of the aliphatic acids to produce CO2 and water [80].

3. Conclusions
Effective destruction of azo dyes belonging to different chemical groups is possible by photocatalysis in the
presence of TiO2 suspensions and UV, Vis or solar light.
The kinetics of the photocatalytic oxidation follows a
LangmuirHinshelwood model and depends on several
factors such as, dye concentration, mass of catalyst, wavelength, radiant flux and the addition of oxidants or the
presence of natural occurring substances (humic substances
and/or inorganic ions). Since the influence of the above factors has been in some cases controversial and it appeared
dependent on substrate type and on various experimental
parameters their impact on the dye degradation must be
accurately checked before their application.
The great part of the studies on the photocatalytic degradation of azo dyes relies only on the monitoring of solution
decolorization, TOC or COD and inorganic ions. Monitoring the disappearance rate of the target dyes is not the most
appropriate parameter to classify the efficiency of this process. Only few studies reported thorough mechanisms with
detailed reaction steps of the different pathways leading to
several photoproducts. Lot of work has to be done also in
the quantification of these intermediates. Quantitative data
about the relative importance of the different routes of degradation is lacking and will help further in the mechanistic
studies of the photocatalytic reactions. Cost-effective treatments to complete compound mineralization are usually not
practicable, therefore, by product evaluation is one of the
keys to optimize each treatment and to maximize the overall process. Toxicity tests of the treated water will gather
also useful information about the practical application of the
photocatalytic process. The better understanding of the photocatalytic process and the operative conditions could give
great opportunities for its application for the destruction of
environmental organic contaminants.

References
[1] H. Zollinger (Ed.), Color Chemistry: Synthesis, Properties and Applications of Organic Dyes and Pigments, second revised ed., VCH,
1991.
[2] J. Weber, V.C. Stickney, Wat. Res. 27 (1993) 63.
[3] C. Rfols, D. Barcel, J. Chromatogr. A 777 (1997) 177.
[4] A. Houas, H. Lachheb, M. Ksibi, E. Elaloui, C. Guillard, J.M.
Hermann, Appl. Catal. B: Environ. 31 (2001) 145.
[5] U. Pagga, D. Bruan, Chemosphere 15 (1986) 479.
[6] A. Bianco-Prevot, C. Baiocchi, M.C. Brussino, E. Pramauro, P.
Savarino, V. Augugliaro, G. Marci, L. Palmisano, Environ. Sci.
Technol. 35 (2001) 971.
[7] B. Neppolian, H.C. Choi, S. Sakthivel, B. Arabindoo, V. Murugesan,
Chemosphere 46 (2002) 1173.
[8] M. Saquib, M. Muneer, Dyes Pigments 56 (2003) 3749.

[9] W.Z. Tang, H. An, Chemosphere 31 (1995) 4157.


[10] V. Meshko, L. Markovska, M. Mincheva, A.E. Rodrigues, Wat. Res.
35 (2001) 3357.
[11] W.S. Kuo, P.H. Ho, Chemosphere 45 (2001) 77.
[12] C. Galindo, P. Jacques, A. Kalt, Chemosphere 45 (2001) 997.
[13] P. Cooper, J. Soc. Dyers. Colour. 109 (1993) 97.
[14] S.S. Patil, V.M. Shinde, Environ. Sci. Technol. 22 (1988) 1160.
[15] A.T. More, A. Vira, S. Fogel, Environ. Sci. Technol. 23 (1989) 403.
[16] M.C. Venceslau, S. Tom, J.J. Simon, Environ. Technol. 15 (1994)
917.
[17] I. Arslan, I.A. Balcioglou, Dyes Pigments 43 (1999) 95.
[18] U. Pagga, K. Taeger, Wat. Res. 28 (1994) 1051.
[19] H.L. Sheng, M.L. Chi, Wat. Res. 27 (12) (1993) 1743.
[20] S.H. Lin, W.Y. Liu, J. Environ. Eng. 120 (1994) 437.
[21] F. Strickland, S. Perkins, Text. Chem. Color. 5 (1995) 11.
[22] Y.M. Slokar, A.M. Le Marechal, Dyes Pigments 37 (1998) 335.
[23] W.G. Kuo, Wat. Res. 26 (1992) 881.
[24] E. Balanosky, J. Fernadez, J. Kiwi, A. Lopez, Wat. Sci. Technol.
40 (1999) 417.
[25] W. Feng, D. Nansheng, Z. Yuegang, Chemosphere 39 (1999) 2079.
[26] C. Morrison, J. Bandara, J. Kiwi, J. Adv. Oxid. Technol. 1 (1996)
160.
[27] S.F. Kang, C.H. Liao, S.T. Po, Chemosphere 41 (2000) 1287.
[28] I. Arslan, T. Akmehmet, T. Tuhkamen, Environ. Technol. 20 (1999)
921.
[29] N.H. Ince, D.T. Gonenc, Environ. Technol. 18 (1997) 179.
[30] K. Vinodgopal, P. Kamat Chemtech. 4 (1996) 18.
[31] C. Lizama, M.C. Yeber, J. Freer, J. Baeza, H.D. Mansilla, Wat. Sci.
Technol. 44 (2001) 197.
[32] F. Zhang, J. Zhao, T. Shen, H. Hidaka, E. Pelizzetti, N. Serpone,
Appl. Catal. B: Environ. 15 (1998) 147.
[33] P. Qu, J. Zhao, T. Shen, H. Hidaka, J. Mol. Catal. A: Chem. 129
(1998) 257.
[34] G. Liu, T. Wu, J. Zhao, H. Hidaka, N. Serpone, Environ. Sci.
Technol. 33 (1999) 2081.
[35] J. Zhao, T. Wu, K. Wu, K. Oikawa, H. Hidaka, N. Serpone, Environ.
Sci. Technol. 32 (1998) 2394.
[36] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem.
Rev. 95 (1995) 69.
[37] I.K. Konstantinou, T.A. Albanis, Appl. Catal. B: Environ. 42 (2003)
319.
[38] A.L. Linsebigler, L. Guangquan, J.T. Yates, Chem. Rev. 95 (1995)
735.
[39] P. Reeves, R. Ohlhausen, D. Sloan, K. Pamplin, T. Scoggins, Solar
Energy 48 (1992) 413.
[40] X.Z. Li, M. Zang, Wat. Sci. Technol. 34 (9) (1996) 49.
[41] Y. Wang, Wat. Res. 34 (2000) 990.
[42] S. Sakthivel, B. Neppolian, M.V. Shankar, B. Arabindoo, M.
Palanichamy, V. Murugesan, Solar Energy Mater. Solar Cells 77
(2003) 65.
[43] V. Augugliaro, C. Baiocchi, A. Bianco-Prevot, E. Garcia-Lopez, V.
Loddo, S. Malato, G. Marci, L. Palmisano, M. Pazzi, E. Pramauro,
Chemosphere 49 (2002) 1223.
[44] M. Stylidi, D.I. Kondarides, X.E. Verykios, Appl. Catal. B: Environ.
40 (2003) 271.
[45] I.A. Balcioglou, I. Arslan, Environ. Pollut. 103 (1998) 261.
[46] S. Gomes de Moraes, R. Sanches Freire, N. Duran, Chemosphere
40 (2000) 369.
[47] I.A. Balcioglou, I. Arslan, Environ. Technol. 18 (1997) 1053.
[48] P. Peralta-Zamora, S. Gomes de Moraes, R. Pelegrini, M. Freire
Jr., J. Reyes, H. Mansilla, N. Duran, Chemosphere 36 (1998) 2119.
[49] I.A. Alaton, I.A. Balcioglou, D.W. Bahnemann, Wat. Res. 36 (2002)
1143.
[50] F. Chen, Y. Xie, J. Zhao, G. Lu, Chemosphere 44 (2001) 1159.
[51] G.A. Epling, C. Lin, Chemosphere 46 (2002) 561.
[52] C.M. So, M.Y. Cheng, J.C. Yu, P.K. Wong, Chemosphere 46 (2002)
905.

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114


[53] J. Grzechulska, A.W. Morawski, Appl. Catal. B: Environ. 36 (2002)
45.
[54] Z. Shourong, H. Qingguo, Z. Jun, W. Bingkun, J. Photochem.
Photobiol. A: Chem. 108 (1997) 235.
[55] M.S.T. Conalves, A.M.F. Oliveira-Campos, M.M.S. Pinto, P.M.S.
Plasencia, M.J.R.P. Queiroz, Chemosphere 39 (1999) 781.
[56] C. Zhu, L. Wang, L. Kong, X. Yang, L. Wang, S. Zheng, F. Chen,
F. MaiZhi, H. Zong Chemosphere 41 (2000) 303.
[57] R. Surez-Parra, I. Hernadez-Perez, M.E. Rincon, S. Lopez-Ayala,
M.C. Roldan-Ahumada, Solar Energy Mat. Solar Cells 76 (2003)
189.
[58] K. Tanaka, K. Padermpole, T. Hisanaga, Wat. Res. 34 (2000) 327.
[59] P.C. Vandevivere, R. Bianchi, W. Verstraete, J. Chem. Technol.
Biotechnol. 72 (1998) 289.
[60] C. Gomes da Silva, J.L. Faria, J. Photochem. Photobiol. A: Chem.
155 (2003) 133.
[61] M.A. Brown, S.C. De Vito, Crit. Rev. Environ. Sci. Technol. 23
(1993) 249.
[62] J.M. Hermann, Catal. Today 53 (1999) 115.
[63] H.D. Burrows, M. Canle, J.A. Santaballa, S. Steenken, J. Photyochem. Photobiol. B: Biol. 67 (2002) 71.
[64] S. Chiron, A. Fernndez-Alba, A. Rodriguez, E.C. Calvo, Wat. Res.
34 (2000) 366.
[65] C. Galindo, P. Jacques, A. Kalt, J. Photochem Photobiol. A: Chem.
130 (2000) 35.
[66] H. Zhan, H. Tian, Dyes Pigments 37 (1998) 231.
[67] J. Bandara, J.A. Mielczarski, J. Kiwi, Langmuir 15 (1999) 7680.
[68] N. Daneshvar, D. Salari, A.R. Khataee, J. Photochem. Photobiol.
A: Chem. 157 (2003) 111.
[69] V. Maurino, C. Minero, E. Pelizzetti, M. Vincenti, Coll. Surf. A
151 (1999) 329.
[70] K. Vinodgopal, I. Bedja, S. Hotchandani, P.V. Kamat, Langmuir 10
(1994) 1767.
[71] K. Vinodgopal, D. Wynkoop, P. Kamat, Environ. Sci. Technol. 30
(1996) 1660.
[72] C. Galindo, P. Jacques, A. Kalt, J. Photochem. Photobiol. A: Chem.
141 (2001) 47.
[73] Y. Xu, C.H. Langford, Langmuir 17 (2001) 897.
[74] L.B. Reutergarth, M. Iangphasuk, Chemosphere 35 (1997) 585.
[75] T. Sauer, G.C. Neto, H.J. Jose, R.F.P.M. Moreira, J. Photochem.
Photobiol. A: Chem. 149 (2002) 147.
[76] H. Lachheb, E. Puzenat, A. Houas, M. Ksibi, E. Elaoui, G. Guillard,
J.M. Hermann, Appl. Catal. B: Environ. 39 (2002) 75.
[77] C.H. Giles, A.P. Da Silva, I.A. Easton, J. Colloid. Interf. Sci. 47
(1974) 766.
[78] W.Z. Tang, C.P. Huang, Wat. Res. 29 (1995) 745.
[79] L. Lucarelli, V. Nadtochenko, J. Kiwi, Langmuir 16 (2000) 1102.
[80] C. Hu, J.C. Yu, Z. Hao, P.K. Wong, Appl. Catal. B: Environ. 42
(2003) 47.
[81] A. Mills, R.H. Davis, D. Worsley, Chem. Soc. Rev. 22 (1993) 417.
[82] I. Poulios, I. Aetopoulou, Environ. Technol. 20 (1999) 479.
[83] L. Zhang, C.Y. Liu, X.M. Ren, J. Photochem. Photobiol. A 85
(1995) 239.
[84] R.J. Davis, J.L. Gainer, G. ONeal, I. Wu, Wat. Environ. Res. 66
(1994) 50.
[85] B. Zielinska, J. Grzechulska, B. Grzmil, A.W. Morawski, Appl.
Catal. B: Environ. 35 (2001) L1L7.
[86] M.A. Fox, M.T. Dulay, Chem. Rev. 93 (1993) 341.
[87] D.W. Bahnemann, J. Cunningham, M.A. Fox, E. Pelizzetti, P.
Pichat, N. Serpone, in: R.G. Zepp, G.R. Heltz, D.G. Crosby (Eds.),
Aquatic Surface Photochemistry, Lewis Publishers, Boca Raton,
1994, p. 261.
[88] I. Poulios, I. Tsachpinis, J. Chem. Technol. Biotechnol. 71 (1999)
349.
[89] S. Tunesi, M. Anderson, J. Phys. Chem. 95 (1991) 3399.
[90] A. Sharma, P. Rao, R.P. Mathur, S.C. Ametha, J. Photochem. Photobiol. A 86 (1995) 97.

13

[91] S. Sakthivel, B. Neppolian, M. Palanichamy, B. Arabindoo, V.


Murugesan, Indian J. Chem. Tech. 6 (1999) 161.
[92] K. Hustert, R.G. Zepp, Chemosphere 24 (1992) 335.
[93] C. Guillard, H. Lachheb, A. Houas, M. Ksibi, E. Elaloui, J.M.
Hermann, J. Photochem. Photobiol. A: Chem. 158 (2003) 27.
[94] S. Naskar, S. Arumugam Pillay, M. Chanda, J. Photochem. Photobiol. A 113 (1998) 254.
[95] B. Neppolian, S. Sakthivel, M. Palanichamy, B. Arabindoo, V.
Murugesan, Stud. Sur. Sci. Cat. 113 (1998) 329.
[96] W.Z. Tang, Z. Zhang, H. An, M.O. Quintana, D.F. Torres, Environ.
Technol. 18 (1997) 1.
[97] A.P. Davis, C.P. Huang, Wat. Sci. Technol. 21 (1990) 455.
[98] D.F. Ollis, E. Pelizzetti, N. Serpone, Environ. Sci. Technol. 25
(1991) 1523.
[99] J. Bandara, V. Nadtochenko, J. Kiwi, C. Pulgarin, Wat. Sci. Technol.
35 (1997) 87.
[100] C. Walling, Acc. Chem. Res. 8 (1975) 125.
[101] C.K. Grtzel, M. Jirousek, M. Grtzel, J. Mol. Catal. 60 (1990) 375.
[102] O. Legrini, E. Oliveros, A.M. Braun, Chem. Rev. 93 (1993) 671.
[103] M.W. Peterson, J.A. Tuner, A.J. Nozik, J. Phys. Chem. 95 (1991)
221.
[104] S. Malato, J. Blanco, C. Richter, B. Braun, M.I. Maldonado, Appl.
Catal. B: Environ. 17 (1998) 347.
[105] A.B. Isil, I. Yuksel, J. Environ. Sci. Health A 31 (1996) 123.
[106] G.A. Epling, C. Lin, Chemosphere 46 (2002) 937.
[107] J. Wiszniowski, D. Robert, J. Surmacz-Gorska, K. Miksch, J.V.
Weber, J. Photochem. Photobiol. A: Chem. 152 (2002) 267.
[108] M. Bekbolet, A.S. Suphandag, C.S. Uyguner, J. Photochem. Photobiol. A: Chem. 148 (2002) 121.
[109] C. Renzi, C. Guillard, J.M. Hermann, P. Pichat, G. Baldi, Chemosphere 35 (1997) 819.
[110] M. Skmen, A. zkan, J. Photochem. Photobiol. A: Chem. 147
(2002) 77.
[111] M. Abdullah, G.K.C. Low, R.W. Matthews, J. Phys. Chem. 94
(1990) 6820.
[112] C. Hu, J.C. Yu, Z. Hao, P.K. Wong, Appl. Catal. B: Environ. 46
(2003) 35.
[113] C. Chen, X. Li, W. Ma, J. Zhao, H. Hidaka, N. Serpone, J. Phys.
Chem. B 106 (2002) 318.
[114] W. Feng, D. Nansheng, Chemosphere 41 (2000) 1137.
[115] M. Mrowetz, E. Selli, J. Photochem. Photobiol. A: Chem., in press.
[116] W. Baran, A. Makowski, W. Wardas, Chemosphere 53 (2003) 87.
[117] J.T. Spadaro, L. Isabelle, V. Renganathan, Environ. Sci. Technol.
28 (1994) 1389.
[118] H. Chun, W. Yizhong, Chemosphere 39 (1999) 2107.
[119] C.A.K. Gouvea, F. Wypych, S.G. Moraes, N. Duran, N. Nagata, P.
Peralta-Zamora, Chemosphere 40 (2000) 433.
[120] C. Minero, V. Maurino, E. Pelizzetti, Res. Chem. Interm. 23 (1997)
291.
[121] C. Hu, Y. Wang, Chemosphere 39 (1999) 2107.
[122] E. Puzenat, H. Lacheb, M. Karkmaz, A. Houas, C. Guillard, J.M.
Hermann, Int. J. Photoenergy 5 (2003) 51.
[123] K. Nohara, H. Hidaka, E. Pelizzetti, N. Serpone, J. Photochem.
Photobiol. A: Chem. 102 (1997) 265.
[124] V. Maurino, C. Minero, E. Pelizzetti, P. Piccinini, N. Serpone, H.
Hidaka, J. Photochem. Photobiol. A: Chem. 109 (1997) 171.
[125] J.M. Hermann, C. Guillard, P. Pichat, Catal. Today 17 (1993) 7.
[126] M. Vauthier, C. Guillard, J.M. Hermann, J. Catal. 201 (2001) 46.
[127] M. Kerzhentsev, C. Guillard, J.M. Herrmann, P. Pichat, Catal. Today
27 (1996) 215.
[128] C. Baiocchi, M.C. Brussino, E. Pramauro, A. Bianco-Prevot, L.
Palmisano, G. Marci, Int. J. Mass Spectrom. 214 (2002) 247.
[129] C. Nasr, K. Vinodgopal, L. Fisher, S. Hotchandani, A.K. Chattopadhyay, P.V. Kamat, J. Phys. Chem. 100 (1996) 8436.
[130] E.J. Weber, R.L. Adams, Environ. Sci. Technol. 29 (1995) 1163.
[131] W.F. Jardim, S.G. Moraes, M.M.K. Takiyama, Wat. Res. 31 (1997)
1728.

14

I.K. Konstantinou, T.A. Albanis / Applied Catalysis B: Environmental 49 (2004) 114

[132] I.E. Tothill, A.P.F. Turner, Tr. Anal. Chem. 15 (1996) 178.
[133] V.B. Manilal, A. Haridas, R. Alexander, G.D. Surender, Wat. Res.
26 (1992) 1035.
[134] S. Ledakowicz, M. Solecka, R. Zylla, J. Biotech. 89 (2001) 175.
[135] I. Poulios, E. Micropoulou, R. Panou, E. Kostopoulou, Appl. Catal.
B: Environ. 41 (2003) 345.
[136] K. Vinodgopal, P. Kamat, Environ. Sci. Technol. 29 (1995) 841.
[137] C. Bauer, P. Jacques, A. Kalt, J. Photochem. Photobiol. A: Chem.
140 (2001) 87.

[138] C. Galindo, P. Jacques, A. Kalt, J. Adv. Oxid. Technol. 4 (1999)


400.
[139] K.V. Subba Rao, B. Lavedrine, P. Boule, J. Environ. Photochem.
Photobiol. A: Chem. 154 (2003) 189.
[140] J. Fernandez, J. Kiwi, C. Lizama, J. Freer, J. Baeza, H.D. Mansilla,
J. Photochem. Photobiol. A: Chem. 151 (2002) 213.
[141] H. Zhan, H. Tian, Dyes Pigments 37 (1998) 249.
[142] E. Stathatos, T. Petrova, P. Lianos, Langmuir 17 (2001) 5025.
[143] H.Y. Shu, C.R. Huang, Chemosphere 29 (1995) 2597.

You might also like