You are on page 1of 9

Environmental Engineering and Management Journal, December 2002, Vol.1, No.

4, 567-575
www.omicron.ch.tuiasi.ro/EEMJ/

“Gh. Asachi” Technical University of Iasi, Romania

_____________________________________________________________________

IRON CONTAINING PILLARED BENTONITES AS


HETEROGENEOUS FENTON-TYPE CATALYSTS FOR
THE OXIDATION OF PHENOL IN WASTEWATERS

Cezar Catrinescu1*, Carmen Teodosiu1, Matei Macoveanu1, Jocelyne


Miehe-Brendle2, Ronan Le Dred2
1
Department of Environmental Engineering, Technical University Iasi-Faculty of
Industrial Chemistry, Bd. D. Mangeron 71A, 6600 Iasi, Romania; 2Laboratoire de
Matériaux Minéraux, CNRS UMR 7016, 3, Rue Alfred Werner, Mulhouse-CEDEX,
France
_______________________________________________________________________
Abstract

This study presents an evaluation of the catalytic performances of some iron containing
clay-based catalysts for the catalytic wet hydrogen peroxide oxidation of phenolic
aqueous wastes. The raw material for the catalyst, a Romanian bentonite, was used to
prepare Al and mixed (Al-Fe) pillared clays using the powder method. Furthermore, the
raw clay and the Al pillared clay were ion exchanged in order to obtain active Fenton-
type catalysts. Selected catalysts were characterized by DRX, BET and chemical
analysis techniques. All the tests were performed on a laboratory scale set-up. Although,
the iron exchanged raw clay exhibits the higher activity in phenol removal, this catalyst
is not stable against leaching. Both pillared clays are highly active in phenol removal,
allowing total elimination of phenol, but the pillared clay with mixed (Al-Fe) pillars is
more stable against iron leaching than the iron exchanged catalysts. However, at an
initial pH of 5.0 this catalyst looses most of its activity.

Keywords: catalytic wet peroxide oxidation, phenolic wastewater, clay, PILCs, Fenton-
type oxidation

1. Introduction

Catalytic wet peroxide oxidation is seen as a promising wastewater


treatment technology. This process is efficient for the removal of priority
pollutants from wastewaters with medium to high organic load (TOC =1-1000
mg C l-1), evacuated at low flow rates (up to 25 m3/h) as demonstrated by Centi
(2002). In the classical version of this process, Fenton’s reagent (Fe2+/Fe3+/H2O2)
is used to generate highly reactive hydroxyl radicals from hydrogen peroxide, in
mild reaction conditions. The process may be applied to wastewaters, sludges, or
contaminated soils, having as effects :

1
* Coresponding author. Phone/Fax: +40-32-237594; Department of Environmental Engineering,
Faculty of Industrial Chemistry, Bd. D. Mangeron 71A, 6600 Iasi, Romania. E-mail:
ccatrine@ch.tuiasi.ro
Catrinescu et al./Environmental Engineering and Management Journal 1 (2002), 4, 567-575

• The destruction organic pollutants


• Toxicity reduction biodegradability improvement (due to the BOD and
COD removal)
• Odor and color removal
Even though these systems offer a cost effective source of HO•, there are
two major drawbacks that limit the industrial application of this technology: (1)
the tight range of pH (e.g., pH = 2.0-4.0) in which the reaction proceeds and (2)
the need for recovering the precipitated catalyst after the treatment. The resulted
sludge may contain organic substances, as well as heavy metals and has to be
further treated, increasing thus the overall costs.
To circumvent these drawbacks, the use of heterogeneous catalysts has
been proposed. The idea behind this choice was to combine the efficiency of the
homogeneous Fenton-type processes and the convenience of heterogeneous
catalysts.
Recently, very interesting preliminary results have been reported on the
catalytic activity of:
- iron containing zeolite ZSM-5 (Fajerwerg and Debellefontaine, 1996;
Fajerwerg et al., 1997), copper containing pillared clays (Barrault et al.,
1998; Frini et al., 1997) and iron containing pillared clays (Barrault et al.,
2000; Catrinescu, 2002) for phenol oxidation;
- transition metal containing zeolites ZSM-5 and Y (Centi et al., 2000;
Larachi et al. 1998) for carboxylic acids oxidation;
- iron containing ZSM-5 zeolite (Centi et al., 2001) for p-cumaric acid
oxidation;
- iron containing USY zeolite (Catrinescu, 2002) for reactive azo dyes
oxidation.
In contrast to early studies in this field using iron ions supported on
alumina (Al-Hayek and Doré,1990) the above mentioned authors concluded that
it may be possible to develop a stable heterogeneous catalyst for the CWPO of
organic refractory compounds.

2. Materials and methods

2.1.Catalyst synthesis and characterization

The host clay used to prepare the catalysts was a natural Romanian
bentonite (Mt) from Valea Chioarului deposit. This clay has been thoroughly
described and characterized in a previous study (Azzouz et al., 1996).
Iron exchanged clay (Fe-Mt) was prepared by a classical ion-exchange
method, at 80 °C in molar solution of Fe(NO3)3.
The Al pillared clay (Al-PILC) catalyst was prepared by pillaring the host
clay using the powder method, as described in detail in a previous paper (Miehe
et al., 1996). Introduction of Fe3+ was performed by a cation doping technique
(Zhu and Lu, 1998). Al-PILC was suspended in 1M NaCl solution at a solid to
liquid ratio of 1:100(w/v). The pH of this suspension was adjusted to 9.0, by
titrating dilute solution of NaOH. The dispersion was stirred overnight, than the
568
Iron containing pillared bentonites as heterogeneous catalysts

solid was separated by filtration, washed with demineralized water and dried at
60 °C. The Na+ doped sample (Na-Al-PILC) was then exchanged three times, at
80 °C, with a 0.1 M aqueous solution of Fe(NO3)3, thus allowing the
introduction of Fe3+ into the catalyst layers (Fe-(Al-PILC)).
For the preparation of mixed Al-Fe pillared clay ((Al-Fe)-PILC), the
pillaring solution was prepared by titration of a 0.1 M Al3+/Fe3+ cationic solution
with 0.2 M NaOH. The cationic solution contained 0.18 and 0.02 mol l-1 of
AlCl3 and FeCl3, respectively. The NaOH solution was slowly added to the
cationic solution, at 60 °C, until the hydrolysis molar ratio (OH/Al+Fe) was 2.
The dry clay powder (2g) was directly dispersed into the pillaring solution. The
intercalation was performed at 50 °C with stirring, for 24 h. Then the catalyst
was separated by filtration, washed with demineralized water, dried overnight at
60 °C and finally calcined at 550 °C for 2h.
The X-ray diffraction patterns were recorded on a Philips PW1800
diffractometer. Specific surface area was determined by N2 adsorption at –196
ºC (BET method) using a Micromeritics ASAP 2010 instrument.

2.2. Phenol oxidation

In order to evaluate phenol removal by catalytic WPO, synthetic water


samples were prepared and analyzed using an experimental set-up as described
below.
Since usually natural and synthetic clay-based catalysts could present
adsorption activity in aqueous medium that would compete with oxidation for
phenol removal, a thorough control of the adsorption activity was performed
prior to oxidation.
Phenol oxidation was carried out in a thermostated glass reactor of 250 ml
equipped with a magnetic stirrer, a reflux condenser and a pH electrode. The
solid catalyst was introduced into 100 ml of an aqueous phenol solution (250 mg
l-1), under continuous stirring. After the stabilization of the temperature at the
selected value and the correction of pH, the solution was analyzed to confirm the
absence of adsorption by the pillared clay. Then a solution of 6 % H2O2 was
added to achieve the selected H2O2/phenol ratio, these being the basic conditions
for experimental tests. Samples of the reaction medium were withdrawn at
regular intervals, and then the reaction was blocked by raising the pH to 9-10,
adding MnO2 and allowing the samples to sit overnight.
Phenols concentrations was determined spectrophotometrically; chemical
oxygen demand (COD) values were obtained by closed reflux colorimetric
method and the content of iron ions in the solution was determined
spectrophotometrically as Fe(1,10-phen)32+ using a HACH DR 2000
spectrophotometer. (APHO, 1992).

569
Catrinescu et al./Environmental Engineering and Management Journal 1 (2002), 4, 567-575

3. Results and discussion

3.1. Catalyst synthesis and characterization

The results of the chemical analysis of the catalysts are presented in


Table 1. In order to correctly correlate these data with the catalytic activity, it
must be pointed out that both Mt and Al-PILC contain some structural iron.
However, these cations are located in framework positions and therefore it is
unlikely to be highly active, as catalytic centers, in the oxidation process.

Table 1. Chemical composition of Mt, Al-PILC and (Al-Fe)-PILC

% Mt Al-PILC (Al-Fe)-PILC
SiO2 77.27 71.56 70.55
Al2O3 15.38 22.9 21.86
MgO 2.28 2.21 2.15
Fe2O3 1.2 1.94 4.21
Na2O 0.64 0.5 0.33
K2O 2.28 0.63 0.67
CaO 0.95 0.26 0.23

Table 2 presents the total iron content (from the parent clay and that
introduced by different methods) of the studied samples. While Mt contains only
structural iron, Fe-Mt and Fe-(Al-PILC) samples also contain Fe3+ located in
ion-exchange positions, where they compensate the negative electric charge of
the structure. In the case of mixed oxide pillar (Al-Fe-PILC) it is accepted that,
a fraction of iron is located in the octahedral sites of Al13 pillars, by isomorphic
substitution. The remaining part of iron is dispersed out of the pillars, in the
lamellar structure of the material.

Table 2. Total iron content of the studied samples

Sample Total iron content, %


Mt 0.79
Fe-Mt 2.15
Fe-(Al-PILC) 2.0
(Al-Fe)-PILC 2.9

Table 3 lists the basal spacing and surface area of PILCs with single and
mixed pillars after calcination at 550 ºC. As a reference, the basal spacing of the
raw clay, at room temperature, is presented as well. The starting clay has a low
thermal stability and if heated at 300 ºC exhibits a surface area of 17.3 m2 g-1
and a basal spacing of about 10 Å. It is known that the thickness of the natural
clay layers is 9.6 Å. Depending on the type of hydroxycation used as pillar
precursor, the pillared material will posses an increased interlayer spacing. For
Al-PILC an intense and sharp d001 peak is obtained, indicating a high layer
stacking order along c- axis. In the case of mixed oxide pillar, the d001 peak
become broader and lower, indicating a lower extent of ordered pillars (Fig. 1).
570
Iron containing pillared bentonites as heterogeneous catalysts

Table 3. XRD basal spacing and BET surface area of Mt, Al-PILC and (Al-Fe)-PILC

XRD (001) basal spacing (Å) Surface area BET


(m2 g-1)
Mta 13.0 46
Al-PILCb 16.07 238.6
(Al-Fe)-PILCb 15.26 230.8
a
– at 25 ºC; b- after calcination at 550 ºC

Fig. 1. X-ray diffraction patterns of PILCs with single and mixed oxide pillars

As can be seen from Table 3 the pillaring process leads to an increase in


the surface area with respect to the raw clay. Fig. 2 displays the nitrogen
adsorption isotherms for pillared clays. They correspond to Type II in Brunauer,
Deming, Deming and Teller classification, with hysteresis loops of Type H3 in
IUPAC classification. These behavior is consistent with the expected structure
of the materials prepared by expanding a layer structure.
Al-PILC has a surface area of 238.6 m2 g-1 whereas for Al-Fe-PILC the
surface area decreases to 230 m2 g-1. The interpretation of these results is
however complicated since, as presented earlier, some precipitated
oxyhydroxide Fe species are retained by the clay and the internal pore surface
area is difficult to define.

Fig. 2. Nitrogen adsorption isotherms for PILCs with single and mixed oxide pillars

571
Catrinescu et al./Environmental Engineering and Management Journal 1 (2002), 4, 567-575

3.2. Catalytic tests

3.2.1 Phenol oxidation in the presence of homogeneous Fe3+/H2O2 system


Firstly, the influence of the pH on the homogeneous Fenton-type process
was studied. The results are presented in Fig. 3.

100
Phenol
80
COD
%, Removal

60 250 mg l-1 Phenol, 50 °C, 20 mg l-


1
Fe3+ 37.23 mmol l-1 H2O2
40
20
0
1.5 2.5 3.5 4.5 5.5 6.5
pH

Fig. 3. Influence of the pH on phenol oxidation by homogeneous Fenton-type system

As can be seen from Fig. 3, for the homogeneous Fenton-type process,


the optimum pH seems to be around 3. Below this pH value, the concentration of
stable Fe3+ species increases at the expense of the active Fe2+ species. For
neutral/basic media, the precipitation of ferric hydroxides takes place and the
parasitic decomposition of H2O2 to H2O and O2 becomes preponderant. The need
for costly acidification of the initial wastewater is one of the major obstacles
towards an industrial application of the process (Centi et al., 2000; Centi et al.,
2001).

3.2.2.Phenol oxidation over solid catalysts


Fig. 4 presents the results of the degradation of phenol over various
solid catalysts.

100 Mt
Phenol removal, %

Fe-Mt
80
Fe-(Al-PILC)
60
(Al-Fe-PILC)
40
20 250 mg l-1 Phenol,
0 pH=3.5;
0 30 60 90 120
50 °C, 1 g l-1 catalyst
37.23 mmol l-1 H2O2
Time, min

Fig. 4. Phenol removal over different solid catalysts

The most active catalysts are Fe-Mt and Fe-(Al-PILC) that contain Fe3+
ions introduced by ion exchange. The catalysts that contain mixed oxide pillars
572
Iron containing pillared bentonites as heterogeneous catalysts

is also active, allowing complete elimination of phenol after 2 h of reaction. As


expected, the raw clay was quite inactive in the catalytic wet peroxide oxidation
process.
Nevertheless, there are still many problems in obtaining adequate an
proof that a supposed heterogeneous catalyst is actually operating in a truly
heterogeneous manner. The conventional practice of recycling the catalyst
without observing a significant loss of activity is not a proof of heterogeneity
(Rafelt and Clark, 2000). The examination of the reaction mixture for leached
catalyst is essential to confirm heterogeneity. Consequently, the amount of Fe
cations in solution after each experiment was determined, after the hot filtration
of the catalyst. (Table 4)

Table 4. Amount of iron in solution after a 2 h experiment

Type of catalyst Mt Fe-Mt Fe-(Al-PILC) (AL-Fe-PILC)


Iron leached, ppm - 12 1.3 0.2

In order to check if the leached fraction iron is responsible for the


catalytic activity, the catalyst was filtered at the reaction temperature, after 30
min of reaction, and the filtrate was tested for activity. Reaction conditions were
the same as in Fig. 4. The results are presented in Fig. 5.
100
90 Fe-Mt
Filtr
80
Phenol removal, %

Fe-(Al-PILC)
70
60 (Al-Fe-PILC)
50
40
30
20
10
0
0 30 60 90 120
Time, min
Fig. 5. Influence of the leached iron on phenol removal; reaction conditions as in Fig. 4

It is obvious that the iron ions leached from Fe-Mt catalyst are able to
further catalyze the process and to obtain a complete removal of phenol. The
catalysts based on pillared clays are more stable against leaching, Fe ions being
strongly attached to the catalyst surface. It is interesting to note the high stability
of the mixed oxide pillared clay (Al-Fe-PILC). In this catalyst, there is a strong
interaction between the iron species and aluminum in the pillars and the other
iron fraction, located outside pillars, has been proved to be also highly stable
against leaching. This feature seems to be very important in obtaining promising
catalysts for heterogeneous Fenton-type processes.

573
Catrinescu et al./Environmental Engineering and Management Journal 1 (2002), 4, 567-575

From the above mentioned catalysts, only the pillared clays were
selected for further experiments, due to their promising behavior. For these
catalysts, two experiments were carried out at two initial pH values (3.3 and 5.0)
of the phenolic wastewaters. The results are presented in Fig. 6.

100 100
Phenol removal, %

Phenol removal, %
80 pH=3.5 80
pH=5.0
60 60

40 40
pH=3.5
20 20
pH=5.0
0 0
0 50 100 0 50 100
Time, min Time, min

Fig. 6. The effect of pH on phenol removal on Al-Fe-PILC (a) and Fe-(Al-PILC) (b)
250 mg l-1 Phenol; 50 °C, 1 g l-1 catalyst 37.23 mmol l-1 H2O2

For AL-Fe-PILC catalyst, the activity was much higher at low pH value
than that at pH = 5.0. On the contrary, the iron exchanged Al-PILC displays
comparable activities at these two initial pH values (pH = 3.0 and 5.0). This
behavior could be due to the different nature of iron in the solid catalysts. For
Fe-(Al-PILC) iron ions are located in the Al-PILC cavities, in ion exchange
positions, held mainly by electrostatic forces, but able to move relatively free, as
in solutions. Their precipitation, as ferric hydroxides, seems to be shielded by
the negative charged surface of the Al-PILC, therefore being able to promote the
oxidation process even at pH=5.0. The iron species in Al-Fe-PILC, either
located in the pillar or precipitated as iron oxides on the catalyst surface, are not
able to establish an effective redox system with H2O2, at pH=5.0.

4. Conclusions

This study presents the results obtained during the catalytic oxidation
with hydrogen peroxide of an aqueous solution of phenol on clay-based catalysts
containing iron. The catalysts prepared by ion-exchange techniques are very
active in removing phenol, but the leached fraction of iron from the catalyst is
not negligible. The catalyst containing mixed Al-Fe pillars exhibits an
interesting activity, allowing the total elimination of phenol in mild reaction
conditions, at low pH values (up to 3.5) and is remarkably stable against
leaching. However, the activity of this catalyst strongly decreases at pH 5.0,
while for the iron exchanged Al-PILC the decrease in activity is less
pronounced. Due to its stability and catalytic activity the mixed Al-Fe pillared

574
Iron containing pillared bentonites as heterogeneous catalysts

clays could be one of the most promising catalysts for the catalytic wet peroxide
oxidation of refractory organic pollutants from wastewaters.

References

Al-Hayek N., Doré M., (1990), Water Research, 24, 973-982.


American Public Health Organisation, (1992), Standard Methods for the Examination of
Water and Wastewater, 18th edition.
Azzouz A., Dumitriu E., Hulea V., Catrinescu C., Carja G., (1996), Progress in
Catalysis, 5, 9-18.
Barrault J., Bouchoule C., Echachoui K., Frini-Srasra N., Trablesi M., Bergaya F.,
(1998), Appl. Catal. B: Environmental, 15, 269-274.
Barrault J., Abdellaoui M., Bouchoule C., Majeste A., Tatibouet J. M., Louloudi A.,
Papayannakos N., Gangas N. H., (2000), Appl. Catal. B: Environmental 27, L225-
L230.
Catrinescu C., Teodosiu C., Macoveanu M., Miehe-Brendle J., Le Dred R., (2002),
Water Research, in press.
Catrinescu C., Neamtu M., Macoveanu M., Yediler A., Kettrup A.,(2002) Environmental
Engineering and Management Journal, 1, 177-186.
Centi G., Perathoner S., Torre, T., Verduna, M.G., (2000), Catal. Today, 55, 61-69.
Centi G., Perathoner S., and Romeo G., (2001), Studies in Surface Science and
Catalysis. In: Zeolites and Mesoporous Materials at the Dawn of the 21st Century,
Galarneau A., Di Renzo F., Fajula F., Vedrine J. (Eds.), Vol. 135, Elsevier,
Amsterdam, 181.
Centi G., (2002), State of the Art in Technology for Water Depollution, EMG on
Cleaner Technologies for Sustainable Chemistry, Trieste, 29-30 April, 2002.
Fajerwerg, K., Debellefontaine, H., (1996), Appl. Catal. B: Environmental, 10, L229-
L235.
Fajerwerg K., Foussard J. N., Perrard A., Debellefontaine H., (1997), Wat. Sci. Tech.,
35, 103-110.
Frini N., Crespin M., Trabelsi M., Messad D., Van Damme H., Bergaya F., (1997),
Applied Clay Science, 12, 281-292.
Larachi F., Levesque S., Sayari A., (1998), J. Chem. Biotechnol., 73, 127-130.
Miehe J., Le Dred R., Saehr D., Baron J., (1996), Synthesis of Porous Materials, Zeolites
and Nanostructures, Occelli M. L., Kessler H., (Eds.), Marcel Dekker, New York,
491.
Rafelt J.S., Clark J.H., (2000), Catal. Today, 57, 33-44.
Zhu H. Y., Lu G. Q., (1998), Journal of Porous Materials, 5, 227-239.

575

You might also like