You are on page 1of 9

Applied Clay Science 9192 (2014) 4654

Contents lists available at ScienceDirect

Applied Clay Science


journal homepage: www.elsevier.com/locate/clay

Research paper

Fe-clay-plate as a heterogeneous catalyst in photo-Fenton oxidation of


phenol as probe molecule for water treatment
Haithem Bel Hadjltaief a, Patrick Da Costa b,, Patricia Beaunier c,d, Mara Elena Glvez e, Mourad Ben Zina a
a

Laboratoire Eau, Energie et Environnement (LR3E), Code: AD-10-02, Ecole Nationale d'Ingnieurs de Sfax, Universit de Sfax, B.P1173.W.3038 Sfax, Tunisia
Sorbonne Universits, UPMC Paris 6, Institut Jean Le Rond d'Alembert, UMR CNRS 7190, 2 Place de la Gare de Ceinture, 78210 Saint Cyr L'cole, France
UPMC, Univ Paris 06, UMR 7197, Laboratoire Ractivit de Surface, Le Raphal, 3 rue Galile, 94200 Ivry, France
d
CNRS, UMR 7197, Laboratoire Ractivit de Surface, Le Raphal, 3 rue Galile, 94200 Ivry, France
e
ETH Zurich, Institute of Energy Technology, ML J 13, Sonneggstr. 3, CH-8092 Zurich, Switzerland
b
c

a r t i c l e

i n f o

Article history:
Received 15 May 2013
Received in revised form 16 January 2014
Accepted 31 January 2014
Available online 3 March 2014
Keywords:
Natural clay
Iron catalyst
Heterogeneous catalyst
Phenol
Water treatment
Photo-Fenton

a b s t r a c t
A novel heterogeneous photo-Fenton plate catalyst was prepared by immobilizing iron species on the surface of
natural Tunisian clay. The activity of this structured catalyst was assayed in the degradation of phenol under UV
irradiation at two different wavelengths (245 nm, UVC, and 365 nm, UVA, radiation). Phenol removal rate from
the aqueous solution always increased in the presence of the Fe-plate catalyst, even under dark-Fenton conditions and for both 254 and 365 nm UV radiation, conrming the efciency of the prepared catalytic system in
the Fenton process. HPLC analysis conrmed a phenol degradation mechanism towards an almost complete mineralization of the organic compound. An apparent activation energy of 48.7 kJ/mol was calculated from the removal experiments performed at different reaction temperatures in the presence of the Fe-plate catalyst.
Catalytic activity remains almost unaltered after ve consecutive reaction cycles re-using the same Fe-plate. Catalyst stability was conrmed by means of TEMEDX analysis, pointing to this novel plate catalyst as a promising
option as a Fenton heterogeneous catalyst for the mineralization of organic compounds in wastewaters.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Industrial, agricultural, and domestic wastes have contributed to the
contamination of water sources with several organic compounds, frequently toxic and non-biodegradable (Jin et al., 2010; Savage and
Diallo, 2005). Such wastewaters have become a major social and economic problem, as modern health-quality standards and environmental
regulations are becoming gradually more restrictive. Among the compounds contained in such wastewaters, phenol is considered as one of
the most toxic pollutants, harmful to human health and to water life
(ATSDR, 2008; Busca et al., 2008). It is moreover classied as a teratogenic and carcinogenic agent. Thus, phenol is listed in water hazard
class 2 in several countries. Biodegradability is only 90% in surface
waters after seven days, and the aquatic toxicity of phenol (LC50) is
12 mg L1 (Daphnia magna, 48 h). In EU countries, the maximum concentration of phenol allowed in drinking water is 0.5 mg L1 (Weber
et al., 2008).
The group of technologies globally known as advanced oxidation
processes (AOPs), is characterized by their high removal efciencies of
refractory organic pollutants, i.e. the ones are difcult to mineralize,
that is, totally oxidize. They are based on the generation of reactive radicals, such as hydroxyls, which are able to oxidize the organic pollutants
Corresponding author. Tel.: +33 1 30 85 48 65; fax: +33 1 30 85 48 99.
E-mail address: patrick.da_costa@upmc.fr (P. Da Costa).

http://dx.doi.org/10.1016/j.clay.2014.01.020
0169-1317/ 2014 Elsevier B.V. All rights reserved.

up to either their mineralization, or the generation of easily biodegradable small molecules, at near-ambient temperature and atmospheric
pressure (Kavitha and Palanivelu, 2004; Martins and Quinta-Ferreira,
2011).
Among the different AOPs, Fenton and photo-Fenton oxidation processes are environmentally friendly, since they do not involve the use of
harmful chemical reagents. Besides such methods are easy to handle
and can be operated using quite uncomplicated reactor designs. Homogeneous-Fenton reaction has been considered in the last decades as one
of the most efcient routes for the treatment of water polluted with recalcitrant chemicals (Bobu et al., 2006; Chen et al., 2010; GarridoRamrez et al., 2010; Herney-Ramrez et al., 2011; Jin et al., 2010;
Liotta et al., 2009; Pera-Titus et al., 2004; Saracco et al., 2001; Wang,
2008). The Fenton process is based on an electron transfer between hydrogen peroxide and a homogeneous metal in solution, generally iron
(II) (Fe2+), resulting in the formation of hydroxyl radicals (Pera-Titus
et al., 2004; Wang, 2008):
2

Fe

H2 O2 Fe

OH OH Fenton reaction:

Fe3+ ions are subsequently regenerated to active Fe2+, with H2O2:


3

Fe

H2 O2 Fe

HO2 H



2
Fe
regeneration :

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

Unfortunately, this last reaction, Eq. (2), proceeds considerably


more slowly than Eq. (1), i.e. k2 = 0.02 mol1 dm3 s 1 vs. k1 =
58 mol1 dm3 s1 (Bai et al., 2013). Therefore, Fe2+ ions are quickly
consumed but slowly regenerated and the resulting low concentration
of Fe2+ makes the overall Fenton reaction slow down.
A combination of UV irradiation and Fenton process, namely photoFenton method, has been developed in order to improve the efciency
in the oxidation of the most refractory organic compounds. In this process, oxidation rate increases, resulting in a higher degree of mineralization, due to the enhanced production of OH radicals. Hydroxyl radicals
are generated due to both the photo-decomposition of hydrogen peroxide and its iron catalyzed decomposition:
H2 O2 UV OH OH
2

Fe

H2 O2 Fe

Photolysis of H2 O2 :

OH OH

PhotoFenton reaction:

However, the homogeneous Fenton process has a signicant disadvantage, for it needs up to 5080 ppm Fe in solution. This is well
above the limits set by EU directives, which allow a maximum of
2 ppm Fe in treated water to be discharged directly into the environment (Walling, 1975). Due to this requirement, the application of the
homogeneous photo-Fenton treatment of large water efuents may
produce considerable amounts of sludge in the nal neutralization
step (Sabhi and Kiwi, 2001). Thus, replacement of the homogeneous
catalysts with heterogeneous catalysts where the active metal can be incorporated into a support stands out as a promising alternative.
Although the use of a heterogeneous catalyst may result in lower oxidation rates than in homogeneous conditions, due to diffusion resistances
of the reactants into the pore and products out of the pore, this problem
can be minimized or completely solved by means of the proper choice of
a support of adequate surface area and pore size distribution. Different
catalyst supports such as synthetic and natural zeolites, clay and pillared
clays, polymers, silica, carbon or resins have been considered as possible
heterogeneous supports (Garrido-Ramrez et al., 2010; Liotta et al.,
2009; Liu et al., 2009; Navalon et al., 2010; Polaert et al., 2002; Santos
et al., 2006; Zazo et al., 2006).
Pillared interlayered clays (PILCs) containing iron oxide pillars (FePILCs) are known as promising heterogeneous Fenton catalysts
(Catrinescu et al., 2011; Herney-Ramrez et al., 2011; Tabet et al.,
2006) and photo-Fenton catalysts (Chen and Zhu, 2007; Iurascu et al.,
2009) for the degradation of organic pollutants in wastewaters, combining a good catalytic activity with high stability against iron leaching.
Moreover, clays have been used for the immobilization of iron species
like aqua-complexes or oxides (Du et al., 2009; Garrido-Ramrez et al.,
2010; Luo et al., 2009). So far, the latter materials are used in slurry
photo-catalytic systems that face reactor design problems associated
with the light penetration into the bulk of treated water and the separation and recovering of heterogeneous photo-catalysts at the end of the
treatment. Guo and Al-Dahhan (2005) have previously reported a

47

ow packed-bed reactor containing a pellet-conformed pillared clay


catalyst, but for phenol wet air oxidation. Three-phase trickle beds and
column reactors are quite common designs for this particular application (Habtu et al., 2011). However, photo-Fenton processes infer additional and singular requirements from the reactor design point of view
that make the application of such reactor concepts unfeasible. In fact,
to the best of our knowledge, only one recent work focuses on the preparation of a non-dispersed heterogeneous catalyst, presented as an ironcontaining natural bentonite (clay) plate for photo-Fenton catalysis for
resorcinol degradation (Gonzlez-Bahamn et al., 2011).
The present study focuses on the preparation of a stable platestructured heterogeneous photo-Fenton catalyst based on natural
Tunisian clay. The activity of this Fe-clay plate catalyst was assayed in
the photo-Fenton degradation of phenol, under different experimental
conditions, considering as well its stability on a prolonged operation
time.
2. Materials and methods
2.1. Catalyst preparation
2.1.1. Purication of the natural clay
Natural clay from the Medenine region (Tunisia) was used as raw
material. The clay was puried by means of careful aqueous dispersion
and decantation. The fraction with a particle size smaller than 2 m
was selected, dispersed in 1 M NaCl solution and stirred at room temperature for 12 h. The supernatant was removed after settling. This procedure was repeated 3 times. After complete exchange, the Na-clay was
separated by centrifugation, washed with distilled water, and nally dialyzed to eliminate the excess of chloride ions (conrmed by the AgNO3
test (Darder et al., 2005)). The solid was then dried at 60 C, ground to
100 mesh, and kept in a sealed vessel.
2.1.2. Synthesis of Fe-immobilized clay plate catalyst
Na-clay was combined with water and sand (Na-clay/sand weight
ratio of 1:1). The mix was molded forming circular plates ( =
7.8 cm, 0.4 cm thickness) which were dried at room temperature
(2530 C), in order to avoid the formation of ssures. The dried solid
was then calcined at 250 C for 4 h. For the deposition of an iron oxide
active phase, the plate was immersed in an aqueous solution of
Fe(NO3)3 (5 g L1), heated at 60 C and kept for 2 h under magnetic stirring. This procedure was repeated 3 times. The Fe-Clay plate catalyst
was then dried was and subsequently calcined at 350 C for 4 h. An
image of the catalyst is presented in Fig. 1.
2.2. Physico-chemical characterization
The chemical composition and structural features of the natural clay
was analyzed by means of X-ray uorescence (XRF, ARL 9800 XP
spectrometer), powder X-ray diffraction (XRD Philips PW 1710

Fig. 1. The Fe-clay plate catalyst.

48

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

diffractometer, K, 40 kV/40 mA, with a scanning rate of 2 per min)


and infrared spectroscopy (IR, Digilab Excalibur FTS 3000 spectrometer;). Loss on ignition of the clay was determined after calcination at
1000 C, until mass change was no longer observed.
The morphology of plates was studied using scanning electronic microscopy (SEM, Hitachi SU-70). This equipment has an Oxford X-Max
50 mm2 X-ray spectroscopy system through dispersive energy (EDX),
which enabled qualitative evaluation of chemical composition. High
resolution transmission electron microscopy (HRTEM) images were acquired on a JEOL JEM 2011 equipped with LaB6 lament and operating
at 200 kV. The images were collected with a 4008 2672 pixel CCD
camera (Gatan Orius SC1000) coupled with the DIGITAL MICROGRAPH
software. Chemical analyses were obtained by a EDX microanalyzer
(PGT IMIX PC) mounted on the microscope. The plates were grinded,
dispersed in ethanol and sonicated. A drop of the dispersion was deposited on a carbon-coated copper grid for the TEM observations.
2.3. Activity tests
The photo-catalytic oxidation experiments were carried out
in a 250 mL Pyrex open vessel, placed on a magnetic stirrer and
under 2 parallel UV-lamps lamps (2 15 at 254 and 365 nm with
930/1350 W cm 2 ). The distance between the solution and the
UV source was kept constant at 15 cm, in all experiments. A schematic of this experimental installation is presented in Fig. 2.
After stabilization of the stirring speed (350 rpm) and pH (3.5), the
Fe-clay plate catalyst was introduced into the vessel containing
100 mL of the aqueous solution of phenol (prepared from analytical
grade phenol 99%, Merck). Then 8 mL of a 1000 mg L1 H2O2 solution,
prepared from H2O2 Merck reagent, was added to the reaction vessel.
The addition of H2O2 was considered as the initial time for reaction.
The solution was subsequently stirred for 3 h. During reaction, liquid aliquots were retrieved from the vessel at selected time intervals. Residual H2O2 in these samples was immediately quenched with MnO2
(Merck), in order to avoid the occurrence of dark Fenton reaction
through the possible presence of leached iron. Before analysis, liquid
was ltered using PTFE lters (0.45 m). Nevertheless, let us remark
here that the liquid solutions after reactions were analyzed by means
of ICP-OES. No Fe in these solutions was detected. Thus, we can assume
that the concentration of Fe ions being under the detection limit of the
apparatus, that is, well belongs to the range of ppm.
The phenol concentration in the solution was analyzed by means of
gas chromatography (GC), in an Agilent 2025 GC equipped with a
Zebron capillary column ZB-5MSi (30 m 0.32 mm 0.25 m) and

Table 1
Chemical composition of natural clay.
Oxides (%)
SiO2

Al2O3

CaO

Fe2O3

MgO

K2O

TiO2

SiO2/Al2O3

48.2

22.3

6.7

17.5

1.7

1.5

1.1

2.2

a ame ionization detector (FID). Phenol removal efciency was calculated as follows:
% 100  C0 Ct =C0 :

Where C0 and Ct (mg L1) are the liquid-phase concentration of the


phenol at the initial and any time t respectively, measured by means
of GC.
In order to complete the information on the phenol oxidation
process, additional analysis of the liquid samples extracted from the
reaction vessel was performed by means of high-pressure liquid chromatography (HPLC, VARIAN (pump and detector) equipped with electronic injector JASCO and a pursuit 5 C18 150 4.6 mm column, with
a detection wavelength of 254 nm under water and acetonitrile and a
total ow rate of 0.5 mL min1).
3. Results and discussion
3.1. Natural clay and catalyst characterization
Table 1 shows the chemical composition of the natural clay, as analyzed by means of XRF. The main constituents are silica, alumina, iron,
calcium. Moreover, the absence of a correlation between SiO2 and
Al2O3 contents indicates that the excess of SiO2 is due to the presence
of quartz, as conrmed by means of XRD. In fact, the powder X-ray
diffraction pattern for this clay, presented in Fig. 3, evidences that
main crystalline phases are quartz (26.7), kaolinite (22.8) and illite
(12.6). On the Fe-impregnated plate after calcination, the iron oxide
phases observed are related to phase -Fe2O3, hematite, with diffraction
peaks appearing at 2 = 32.2, 35.7 and 36.4 (Ayodele and Hameed,
2013; Carriazo, 2012; Nogueira et al., 2011).
The FT-IR absorption spectra, Fig. 4, show the presence of OH
stretching bands at the vicinity of the 3500 cm 1 domain, as well as
the Si\O stretching bands near 1000 cm 1 (Balan et al., 2001). The
Q

Q: quartz
An: anhydrid compouned(CaSO4 )
H: hematite
D: dickite
G: gluconite
I: illite
K: kaolinite

10

15

20

CaSO 4

An

Raw clay

H H

25

Calcined clay

Q
H

H HH

30

35

40

45

50

55

60

Tetha (degr)
Fig. 2. Experimental set-up used in for the phenol degradation experiments (1: aquarium
mirror, 2: magnetic stirrer, 3: plate catalyst, 4: open Pyrex vessel, and 5: UV lamps).

Fig. 3. Powder X-ray diffraction pattern for the raw natural clay and the Fe impregnated
clay after calcination.

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

-OH stretching

Si-O
stretching

Si-O-Mg
carbonates

Si-O-Al

49

Table 1, a feature being relatively typical of Tunisian clays. However,


iron content increase from 2.8 to 6.1 at.% conrms the efciency of the
ion-exchange treatment, which leads to the deposition of an approx.
3.5 wt.%. TEM observation of the clay plate. Fig. 6a conrms the existence of a large variety of constituents with different morphologies
and chemical compositions. Fig. 6b shows an illite particle near an aggregate of magnetite crystals. The EDX spectrum on the illite site,
Fig. 6e, evidences the presence of a small quantity of iron inside the particle. Moreover, the study of the Fe-plate catalyst, i.e. Fig. 6f, proved that
Fe species are effectively deposited on the surface of the clay.
3.2. Catalytic performance of Fe-plates and kinetic study

Fig. 4. FTIR spectra of the raw natural clay.

band appearing at 1428 cm1 corresponds to or is indicative of the presence of carbonates, i.e. calcite (CaCO3) or dolomite (Ca, Mg(CO3)2), conrmed by the presence of CaO and MgO revealed by the chemical
analysis by means of XRF. Additionally, bands at 472 and 533 cm1
can be assigned to Si\O\Mg and Si\O\Al, respectively.
SEM images acquired for the already conformed plates, before and
after iron loading, are shown in Fig. 5a and c, and b and d, respectively.
Images evidence a relatively noticeable change in the material morphology upon iron addition. The clay plate surface, see Fig. 5a, appears more
uniform and at than after iron ion-exchange, Fig. 5b. A closer look,
Fig. 5d, evidences some particle fragmentation and the alignment of
the material in the form of cross-linked layers, as a consequence of
iron loading. SEMEDX analysis results are shown in Table 2. Initial
iron content in the clay was as well evidenced by means of XRF analysis,

Fig. 7 shows the measured phenol removal efciency () as a function of reaction time, for the several experiments performed in the presence of H2O2, Fe-plate catalysts, H2O2 and Fe-plate catalyst, H2O2 and UV
irradiation at 254 nm, H2O2 and UV irradiation at 365 nm, H2O2 + UV
254 nm and Fe-plate catalysts, and H2O2 + UV 365 nm and Fe-plate catalyst. In spite of its expected low oxidation potential (Zhou et al., 2011),
it can be observed that phenol can be already oxidized in the only presence of hydrogen peroxide, reaching a maximal conversion of 8% after
150 min of time-on-stream, and remaining almost constant in the
next 30 min of further reaction. The experimental run performed in
the presence of the Fe-plate catalyst, in the absence of H2O2 or UV irradiation, evidences a maximal phenol conversion of around 2225%,
which becomes almost steady after 50 min of reaction time. This phenol
removal efciency can be ascribed to the adsorption of phenol on the
catalyst surface. For the sake of comparison, a non-impregnated plate
has been tested and only a 15% removal efciency had been obtained
after 120 min of time on stream. Phenol degradation extent increases
substantially if H2O2 is added to the reaction mixture. In this case, a removal efciency of 68% is achieved after 180 min of reaction, which indicates that Fenton reaction takes place resulting in the enhanced
formation of radicals that are involved in the oxidation of the organic
compound. Reaction may be hindered to some point by the presence
of originally inactive Fe species (Aleksi et al., 2010), or by the

a)

b)

c)

d)

Fig. 5. SEM images for the clay plates a) and c) before; and b) and d) after Fe impregnation.

50

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

in Fig. 7 evidence that phenol oxidation proceeds faster under UV irradiation at a wavelength of 254 nm than when that of 365 nm was
employed. The phenol removal efciency of 100% was attained after
less than 60 min in the rst case; whereas it takes 120 min to completely
degrade phenol in the second case. This is rst of all due to fact that
UV254 radiation is absorbed more effectively by the different iron species present in the system, as well as by phenol (Legrini et al., 1993).
Moreover, it is well known that the photo-hydrolysis of H2O2, Eq. (3),
proceeds faster and more effectively under UV radiation at wavelengths
lower than 320 nm (Chen and Zhu, 2007; Legrini et al., 1993; Walling,
1975). Therefore, a higher extent of radical formation through this pathway can be expected, resulting in faster oxidation of the organic compound. It is worth noting as well the differences in phenol removal
efciency as a function of reaction time measured under UV irradiation
and upon H2O2 addition, in the absence of the Fe-plate catalyst. Enhanced photolysis of H2O2 under UV254 irradiation results in phenol removal efciencies which are all the time higher than those measured in
the presence of the Fe-plate catalysts irradiated with UV365. In fact 100%

Table 2
EDX of the samples.
C

Na

Mg

Al

Si

Ca

Fe

3.2
2.7

59.1
61.5

0.8
1

2.1
1.5

8.6
6.6

20.8
18.9

0.3
0.1

2.0
1.5

2.8
6.1

Atomic %
Plate
Fe-plate

essentially lower rate of Fe2+ regeneration reaction, Eq. (2), in comparison to Fenton radical generation, Eq. (1). UV irradiation is an important
key for achieving higher phenol removal efciencies. When reaction is
performed in the presence of the Fe-plate catalyst, upon H2O2 addition
and UV irradiation, 100% phenol removal can be nally attained. This
fact proves the well know higher effectiveness of the photo-Fenton process in comparison to dark Fenton reaction, which is due to the enhanced formation of radicals. Yield, however, is strongly dependent on
the wavelength of the UV radiation employed. The results presented

b)
a)

c)

d)

e)

f)

Fig. 6. TEM images for the a) and b) clay plates; c) Fe-clay plate and d) Fe-clay plate after 5 consecutive cycles of phenol oxidation. EDX spectra of e) clay-plate and f) Fe-clay plate.

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

51

H2O2 + Fe-plate + UV254

100

initial
15 min
30 min
45 min
120 min

H2O2 + Fe-plate + UV365


H2O2 + UV365

(%)

H2O2 + UV254

60

H2O2 + Fe-plate

40

Intensity (a.u.)

80

Fe-plate

20
H2O2

0
0

20

40

60

80

100

120

140

160

180

Reaction time (min)

Retention time (min)

Fig. 7. Phenol removal efciency measured during experimental runs under different reaction conditions.

Fig. 9. HPLC chromatograms for the solutions extracted from the reaction vessel at different reaction times, during the photo-Fenton phenol removal experiment in the presence
of the Fe-plate upon H2O2 addition and UV254 irradiation.

removal efciency can be reached in this case, after 120 min of reaction,
even in the absence of a catalyst.
Color change in the reaction solution during the degradation
of phenol has been reported by Mijangos et al. (2006), who indicated
that some highly colored intermediate compounds may include
p-benzoquinone (yellow), o-benzoquinone (red), and hydroquinone
(colorless), and the mixed solution of all intermediate compounds revealed a brown color. Moreover, these colored intermediates are more
refractory and difcult to oxidize. Their increased stability is due to
the conjugated carbonyl groups contained in their chemical structure
(Mijangos et al., 2006). Generally, they possess higher toxicity than
phenol itself (Yalfani et al., 2009). Fig. 8 reports a picture of each of
the corresponding nal solution after 120 min of reaction. Incomplete
phenol oxidation in the case of the reaction performed upon H2O2 addition either in the presence of the Fe-plate (non-irradiated), Fig. 8 solution (a), or under UV365 irradiation (non-catalytic), Fig. 8 solution (b),
results in fact in a certain brownish color, in agreement with the observations of Mijangos et al. (2006), pointing to the presence of such intermediates, i.e. p-benzoquinone and/or o-benzoquinone. On the other
hand, the solutions for those experimental runs resulting in 100% phenol removal efciency, UV365 + H2O2 + Fe-plate, UV254 + H2O2 and
UV254 + H2O2 + Fe-plate, Fig. 8 solutions (c), (d) and (e) respectively,
appear colorless.

The mechanism of phenol degradation in water solutions has been


quite extensively studied (Davlin and Harris, 1984; Duprez et al.,
1996; Santos et al., 2006; Soria-Snchez et al., 2011). There is a general
agreement that phenol oxidation starts with the hydroxylation of the
molecule leading either to hydroquinone or catechol, which are subsequently oxidized to p-benzoquinone and o-benzoquinone, respectively.
Therefore, the reaction is proceeding in two parallel pathways with maleic acid as common intermediate. Maleic acid can be then directly oxidized to CO2 or via oxalic and formic acids which are partially oxidized
to acetic acid, considered as the most refractory product of phenol degradation. Normally, 90% of phenol is oxidized into CO2 and 10% remains
as acetic acid in the media (Duprez et al., 1996).
Fig. 9 shows the HPLC chromatograms acquired for aliquots taken
out of the reaction vessel at different reaction times, during phenol degradation experiment upon H2O2 addition in the presence of the Fe-plate
catalyst and UV irradiation of 254 nm wavelength. Phenol peak appears
at a retention time of 5.53 min. After some minutes of reaction, the intensity of this phenol peak diminishes, whereas the intensity of the
peak appearing at retention times immediately near to that of phenol
increases. This peak can be assigned to the presence of hydrobenzoic
acid, which has been considered as well as an intermediate in the phenol oxidation process (Eftaxias et al., 2006; Quintanilla et al., 2006). The
HPLC chromatogram obtained for the solution after 30 min of reaction

Fig. 8. Aliquots of the reaction solutions after 120 min of reaction time (a) H2O2 + Fe-plate, (b) H2O2 + UV365, c) H2O2 + UV365 + Fe-plate, d) H2O2 + UV254 and e) H2O2 + UV254 + Fe-plate.

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

time evidences a higher phenol oxidation extent, reected in the markedly lower intensity of the peak at 5.5 min, whereas a new peak appears
between those corresponding to fumaric and maleic acids, which can be
due to the formation of acetic acid. Already after 45 min of reaction time,
the phenol peak completely disappears from the HPLC chromatogram.
After 120 min of reaction time, only small amounts of the low molecular
weight acids are present.

100
80

(%)

52

60
40

3.3. Inuence of phenol concentration and reaction temperature


It is of practical interest to investigate the effect of the initial pollutant phenol concentration on the removal efciency that can be
attained, in order to evaluate the behavior of this particular water treatment system. Fig. 10 shows the inuence of various initial phenol concentrations on the removal efciency measured in the photo-Fenton
reaction, in the presence of the Fe-plate catalysts and under UV254 irradiation. The rate of phenol degradation decreases for increasing concentration of this organic compound. In this sense, 100% removal efciency
was achieved after 17 min of reaction for an initial concentration of
25 mg L1, 32 min for an initial concentration of 75 mg L1, 60 min
for 150 mg L 1 and 143 min for 200 mg L1. In the heterogeneous
photo-Fenton process, the reaction occurs at the surface of Fe-plates between the OH radicals generated at the active sites and phenol molecules adsorbed on the surface. Thus, when phenol concentration is
high enough, the number of active sites available decreases due to competitive adsorption of the phenol molecules on the catalytic surface
(Chen et al., 2008). In addition, the intermediate products of phenol oxidation might also compete for adsorption sites with phenol molecules,
which may block their interaction with the Fe(II)/Fe(III) active phase
(Chen et al., 2008).
The inuence of reaction temperature on the photo-Fenton degradation of phenol in the presence of the prepared Fe-plates catalyst was as
well studied and the results obtained are reported in Fig. 11. Temperature substantially inuences the rate of the phenol degradation reaction.
For example, after 20 min of reaction, degradation efciency increases
from about 50% at 20 C to almost 100% at 50 C. The increase in temperature enhances the rate of hydroxylation at the catalyst active sites
resulting in an increased production of OH radicals. Moreover, a higher
reaction temperature can provide more energy for the reactants to overcome the activation energy barrier (Xu et al., 2008). This excess energy
leads to higher collision frequency between the OH radicals and phenol

20C
40C
50C

20
0
0

10

20

30

40

50

60

Reaction time (min)


Fig. 11. Phenol removal efciency as a function of reaction time measured during
the H2O2 + UV254 + Fe-plate experiment at different temperatures (20, 40 and 50 C).

(or intermediate) molecules which eventually results in faster degradation (Sun et al., 2007).
The apparent activation energy Ea, for the degradation of phenol by
photo-Fenton oxidation can be calculated from the corresponding
Arrhenius equation:
k AexpEa =RT:

An activation energy value of 48.7 kJ mol1 was obtained. Generally, the reaction activation energy of ordinary thermal reactions is usually between 60 and 250 kJ mol1 (Chen and Zhu, 2007). With respect to
similar catalytic processes, Shukla et al. reported values of around 61.7
75.5 kJ/mol for Co/SiO2 catalysts (Shukla et al., 2011), and around
67.4 kJ/mol for Co/SBA-15 (P. Shukla et al., 2010; P.R. Shukla et al.,
2010), whereas activation energy was found to be slightly lower but
still around 60 kJ/mol for Co supported on activated carbon (P. Shukla
et al., 2010; P.R. Shukla et al., 2010). This result implies that the degradation of phenol in an aqueous solution by the photo-Fenton oxidation
process in the presence of the Fe-plate catalyst prepared requires lower
activation energy than the average of the degradation systems, and
points to this material and derived structured catalyst as a perspective
and effective option for the decontamination of wastewater containing
such organic pollutant.

100
3.4. Stability of the catalytic system

(%)

80
60
40

25 mg/L
75 mg/L
150 mg/L
200 mg/L

20
0
0

20

40

60

80

100

120

140

160

180

Reaction time (min)


Fig. 10. Phenol removal efciency as a function of reaction time measured during the H2O2 +
UV254 + Fe-plate experiment for different initial phenol concentrations (25, 75, 150 and
200 mg L1).

In order to evaluate catalyst stability and the possible change in its


activity or in its physic-chemical features, successive removal experiments were performed. The Fe-plate catalyst was therefore used in
ve consecutive cycles, i.e. 1 h of operation using fresh phenol solutions,
under UV254 irradiation. After each run, the Fe-plate catalyst was removed from the reaction vessel, carefully washed with distilled water
and dried at 60 C for 12 h. Regarding the changes in activity observed
with subsequent reaction cycles, a slight decay of about 10% in the removal efciency attained after 1 h of reaction time was measured but
only after the 4th consecutive reuse of the catalyst. Moreover, no weight
loss of the Fe-plate catalyst was measured after the ve reaction cycles.
TEM observations conrmed this stability. After successive phenol degradation experiments, the Fe-plate surface was not modied. The catalyst surface presented still the same aspect than the one observed in
Fig. 6d. Moreover, EDX analysis conrmed the same quantitative presence of the Fe species.

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

4. Conclusions
A Fe-clay plate catalyst was prepared using Tunisian clay as a
starting material. Physico-chemical characterization of the catalyst evidenced successful immobilization of the Fe-active phase on the clay
matrix.
The activity of this catalyst was assayed in the heterogeneous
photo-Fenton oxidation of a probe molecule: phenol in aqueous solution which is representative of water contaminant. Phenol removal efciency of 100% was attained after less than 60 min of reaction in
the presence of the prepared Fe-plate catalyst and under UV irradiation of 254 nm wavelength. For equivalent reaction conditions, i.e.
H2O2 addition, UV irradiation and wavelength, and after a certain reaction time, phenol removal efciency measured was all the time
higher in the presence of the Fe-plate catalyst. In other words, the
rate of phenol degradation was higher in the presence of such catalytic system, proving its efciency as a Fenton catalyst even under
the less favorable reaction conditions. Moreover, HPLC analysis of
samples periodically extracted from the reaction vessel conrmed
phenol degradation following the reaction pathways described in
literature, and almost reaching complete mineralization of the organic compound, with a small amount of the more refractory
short-chain acids remaining in the solution after 120 min of reaction
time.
Negligible loss of activity was observed after ve consecutive
reaction cycles performed re-using the same Fe-plate catalyst.
TEM observation of the catalyst, as well as its EDX analysis,
evidenced no visible modication of surface morphology
and chemical composition upon consecutive reaction cycles,
conrming the good stability of the Fe-plate heterogeneous
catalytic system.

References
Aleksi, M., Kui, H., Koprivanac, N., Leszczynska, D., Boi, A.L., 2010. Heterogeneous
Fenton type processes for the degradation of organic dye pollutant in water the application of zeolite assisted AOPs. Desalination 257, 2229.
ATSDR, 2008. ToxFAQs for Phenol. Agency for Toxic Substances and Disease Registry
(ATSDR) (http://www.atsdr.cdc.gov/).
Ayodele, O.B., Hameed, B.H., 2013. Synthesis of copper pillared bentonite ferrioxalate catalyst for degradation of 4-nitrophenol in visible light assisted Fenton process. J. Ind.
Eng. Chem. 19, 966974.
Bai, C., Xiao, W., Feng, D., Xian, M., Guo, D., Ge, Z., Zhou, Y., 2013. Efcient decolorization of
malachite green in the Fenton reaction catalyzed by [Fe(III)-salen]Cl complex. Chem.
Eng. J. 215216, 227234.
Balan, E., Saitta, A.M., Mauri, F., Calas, G., 2001. First principles modeling of the infrared
spectrum of kaolinite. Am. Mineral. 86, 13211330.
Bobu, M., Wilson, S., Greibrokk, T., Lundanes, E., Siminiceanu, I., 2006. Comparison of advanced oxidation processes and identication of monuron photodegradation products in aqueous solution. Chemosphere 63, 17181727.
Busca, G., Berardinelli, S., Resini, C., Arrighi, L., 2008. Technologies for the removal of phenol from uid streams: a short review of recent developments. J. Hazard. Mater. 160,
265288.
Carriazo, J.G., 2012. Inuence of iron removal on the synthesis of pillared clays: a surface
study by nitrogen adsorption, XRD and EPR. Appl. Clay Sci. 6768, 99105.
Catrinescu, C., Arsene, D., Teodosiu, C., 2011. Catalytic wet hydrogen peroxide oxidation of
para-chlorophenol over Al/Fe pillared clays (AlFePILCs) prepared from different host
clays. Appl. Catal. B Environ. 101, 197205.
Chen, J., Zhu, L., 2007. Heterogeneous UVFenton catalytic degradation of dyestuff in
water with hydroxyl-Fe pillared bentonite. Catal. Today 126, 463470.
Chen, A., Ma, X., Sun, H., 2008. Decolorization of KN-R catalyzed by Fe-containing Y and
ZSM-5 zeolites. J. Hazard. Mater. 156, 451460.
Chen, Q., Wu, P., Dang, Z., Zhu, N., Li, P., Wu, J., Wang, X., 2010. Iron pillared vermiculite as
a heterogeneous photo-Fenton catalyst for photocatalytic degradation of azo dye reactive brilliant orange X-GN. Sep. Purif. Technol. 71, 315323.
Darder, M., Colilla, M., Ruiz-Hitzky, E., 2005. Chitosanclay nanocomposites: application
as electrochemical sensors. Appl. Clay Sci. 28, 199208.
Davlin, H.R., Harris, I.J., 1984. Mechanism of the oxidation of aqueous phenol with dissolved oxygen. Ind. Chem. Res. Fundam. 23, 387392.
Du, W., Sun, Q., Lv, X., Xu, Y., 2009. Enhanced activity of iron oxide dispersed on bentonite
for the catalytic degradation of organic dye under visible light. Catal. Commun. 10,
18541858.
Duprez, D., Delano, F., Barbier Jr., J., Isnard, P., Blanchard, G., 1996. Catalytic oxidation of
organic compounds in aqueous media. Catal. Today 29, 317322.

53

Eftaxias, A., Font, J., Fortuny, A., Fabregat, A., Stber, F., 2006. Catalytic wet air oxidation of
phenol over active carbon catalysts. Global kinetic modeling using simulated annealing. Appl. Catal. B Environ. 67, 1223.
Garrido-Ramrez, E.G., Theng, B.K.G., Mora, M.L., 2010. Clays and oxide minerals as
catalysts and nanocatalysts in Fenton-like reactionsa review. Appl. Clay Sci. 47,
182192.
Gonzlez-Bahamn, L.F., Hoyos, D.F., Bentez, N., Pulgarn, C., 2011. New Fe-immobilized
natural bentonite plate used as photo-Fenton catalyst for organic pollutant degradation. Chemosphere 82, 11851189.
Guo, J., Al-Dahhan, M., 2005. Catalytic wet air oxidation of phenol in concurrent
downow and upow packed-bed reactors over pillared clay catalysts. Chem. Eng.
Sci. 60, 735746.
Habtu, N., Font, J., Fortuny, A., Bengoa, C., Fabregat, A., Haure, P., Ayude, A., Stber, F., 2011.
Heat transfer in trickle bed column with constant and modulated feed temperature:
experiments and modeling. Chem. Eng. Sci. 66, 33583368.
Herney-Ramrez, J., Silva, A.M.T., Vicente, M.A., Costa, C.A., Madeira, L.M., 2011. Degradation of acid orange 7 using a saponite-based catalyst in wet hydrogen peroxide
oxidation: kinetic study with the Fermi's equation. Appl. Catal. B Environ. 101,
197205.
Iurascu, B., Siminiceanu, I., Vion, D., Vicente, M.A., Gil, A., 2009. Phenol degradation in
water through a heterogeneous photo-Fenton process catalyzed by Fe-treated
laponite. Water Res. 43, 13131322.
Jin, N.B., Christopher, W.K., Saint, C.C., 2010. Recent developments in photocatalytic water
treatment technology: a review. Water Res. 44, 29973027.
Kavitha, V., Palanivelu, K., 2004. The role of ferrous ion in Fenton and photo-Fenton processes for the degradation of phenol. Chemosphere 55, 12351243.
Legrini, O., Oliveros, E., Braun, A.M., 1993. Photochemical processes for water treatment.
Chem. Rev. 93, 671698.
Liotta, L.F., Gruttadauria, M., Di Carlo, G., Perrini, G., Librando, V., 2009. Heterogeneous catalytic degradation of phenolic substrates: catalysts activity. J. Hazard. Mater. 162,
588606.
Liu, T., You, H., Chen, Q., 2009. Heterogeneous photo-Fenton degradation of polyacrylamide in aqueous solution over Fe(III)SiO2 catalyst. J. Hazard. Mater. 162,
18601865.
Luo, M.L., Bowden, D., Brimblecombe, P., 2009. Catalytic property of FeAl pillared clay for
Fenton oxidation of phenol by H2O2. Appl. Catal. B Environ. 85, 201206.
Martins, R.C., Quinta-Ferreira, R.M., 2011. Remediation of phenolic wastewaters by advanced oxidation processes (AOPs) at ambient conditions: comparative studies.
Chem. Eng. Sci. 66, 32433250.
Mijangos, F., Varona, F., Villota, N., 2006. Changes in solution color during phenol oxidation by Fenton reagent. Environ. Sci. Technol. 40, 55385543.
Navalon, S., Alvaro, M., Garcia, H., 2010. Heterogeneous Fenton catalysts based on clays,
silicas and zeolites. Appl. Catal. B Environ. 99, 126.
Nogueira, F.G.E., Lopes, J.H., Silva, A.C., Lago, R.M., Fabris, J.D., Oliveira, L.C.A., 2011. Catalysts based on clay and iron oxide for oxidation of toluene. Appl. Clay Sci. 51,
385389.
Pera-Titus, M., Garca-Molina, V., Baos, M.A., Gimnez, J., Esplugas, S., 2004. Degradation
of chlorophenols by means of advanced oxidation processes: a general review. Appl.
Catal. B Environ. 47, 219256.
Polaert, I., Wilhelm, A.M., Delmas, H., 2002. Phenol wastewater treatment by a two-step
adsorptionoxidation process on activated carbon. Chem. Eng. Sci. 57, 15851590.
Quintanilla, A., Casas, J.A., Mohedano, A.F., Rodrguez, J.J., 2006. Reaction pathway of the
catalytic wet air oxidation of phenol with a Fe/activated carbon catalyst. Appl.
Catal. B Environ. 67, 206216.
Sabhi, S., Kiwi, J., 2001. Degradation of 2,4-dichlorophenol by immobilized iron catalysts.
Water Res. 35, 19942002.
Santos, A., Yustos, P., Gomis, S., Ruiz, G., Garca-Ochoa, F., 2006. Reaction network and kinetic modeling of wet oxidation of phenol catalyzed by activated carbon. Chem. Eng.
Sci. 61, 24572467.
Saracco, G., Solarino, L., Specchia, V., Maja, M., 2001. Electrolytic abatement of
biorefractory organics by combining bulk and electrode oxidation processes. Chem.
Eng. Sci. 56, 15711578.
Savage, N., Diallo, M.S., 2005. Nanomaterials and water purication: opportunities and
challenges. J. Nanoparticle Res. 7, 331342.
Shukla, P., Wang, S.B., Singh, K., Ang, H.M., Tade, M.O., 2010a. Cobalt exchanged zeolites for
heterogeneous catalytic oxidation of phenol in the presence of peroxymonosulphate.
Appl. Catal. B Environ. 99, 163169.
Shukla, P.R., Wang, S.B., Sun, H.Q., Ang, H.M., Tade, M., 2010b. Activated carbon supported
cobalt catalysts for advanced oxidation of organic contaminants in aqueous solution.
Appl. Catal. B Environ. 100, 529534.
Shukla, P., Sun, H.Q., Wang, S.B., Ang, H.M., Tade, M.O., 2011. Photocatalytic oxidation of
phenolic compounds using zinc oxide and sulphate radicals under articial solar
light. Sep. Purif. Technol. 77, 230236.
Soria-Snchez, M., Maroto-Valiente, A., lvarez-Rodrguez, J., Muoz-Andrs, V., RodrguezRamos, I., Guerrero-Ruz, A., 2011. Carbon nanostructured materials as direct catalysts
for phenol oxidation in aqueous phase. Appl. Catal. B Environ. 104, 101109.
Sun, J.H., Sun, S.P., Wang, G.L., Qiao, L.P., 2007. Degradation of azo dye amido black 10B in
aqueous solution by Fenton oxidation process. Dyes Pigments 74, 647652.
Tabet, D., Saidi, M., Houari, M., Pichat, P., Khalaf, H., 2006. Fe-pillared clay as a Fenton-type
heterogeneous catalyst for cinnamic acid degradation. J. Environ. Manag. 80,
342346.
Walling, C., 1975. Fenton's reagent revisited. Acc. Chem. Res. 8, 125131.
Wang, S.A., 2008. Comparative study of Fenton and Fenton-like reaction kinetics in
decolourisation of wastewater. Dyes Pigments 76, 714720.
Weber, M., Weber, M., Kleine-Boymann, M., 2008. Phenol. Ullmann's Encyclopedia of
Industrial Chemistry. John Wiley & Sons.

54

H. Bel Hadjltaief et al. / Applied Clay Science 9192 (2014) 4654

Xu, H.Y., Prasad, M., Liu, Y., 2008. Schorl: a novel catalyst in mineral-catalyzed Fenton like
system for dyeing wastewater discoloration. J. Hazard. Mater. 165, 11861192.
Yalfani, M.S., Contreras, S., Medina, F., Sueiras, J., 2009. Phenol degradation by Fenton's
process using catalytic in situ generated hydrogen peroxide. Appl. Catal. B Environ.
89, 519526.

Zazo, J.A., Casas, J.A., Mohedano, A.F., Rodrguez, J.J., 2006. Catalytic wet peroxide oxidation of phenol with Fe/active carbon catalyst. Appl. Catal. B Environ. 65,
261268.
Zhou, S., Qian, Z., Tao, S., Xu, J., Xia, C., 2011. Catalytic wet peroxide oxidation of phenol
over CuNiAl hydrotalcite. Appl. Clay Sci. 53, 627633.

You might also like